0% found this document useful (0 votes)
599 views542 pages

Full

This document provides an overview of spectroscopy techniques. It discusses how spectroscopy works by studying the interaction of electromagnetic radiation with atoms and molecules. Different regions of the electromagnetic spectrum provide different information. Techniques covered include magnetic resonance spectroscopies like NMR and EPR, rotational spectroscopy, vibrational spectroscopy like infrared spectroscopy, electronic spectroscopy including UV-Vis, photoelectron spectroscopy, x-ray spectroscopy, photoacoustic spectroscopy, and Mössbauer spectroscopy.

Uploaded by

Soumajit Das
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
599 views542 pages

Full

This document provides an overview of spectroscopy techniques. It discusses how spectroscopy works by studying the interaction of electromagnetic radiation with atoms and molecules. Different regions of the electromagnetic spectrum provide different information. Techniques covered include magnetic resonance spectroscopies like NMR and EPR, rotational spectroscopy, vibrational spectroscopy like infrared spectroscopy, electronic spectroscopy including UV-Vis, photoelectron spectroscopy, x-ray spectroscopy, photoacoustic spectroscopy, and Mössbauer spectroscopy.

Uploaded by

Soumajit Das
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 542

SPECTROSCOPY

William Reusch
Michigan State University
SECTION OVERVIEW
SPECTROSCOPY
Most of what we know about the structure of atoms and molecules comes from studying their
interaction with light (electromagnetic radiation). Different regions of the electromagnetic spectrum
provide different kinds of information as a result of such interactions.

FUNDAMENTALS OF SPECTROSCOPY
ELECTROMAGNETIC RADIATION
INTRODUCTION TO SPECTROSCOPY
LINESHAPE FUNCTIONS
SELECTION RULES
SELECTION RULES AND TRANSITION MOMENT INTEGRAL
THE POWER OF THE FOURIER TRANSFORM FOR SPECTROSCOPISTS
TIME-RESOLVED VS. FREQUENCY RESOLVED
BACK MATTER
INDEX

MAGNETIC RESONANCE SPECTROSCOPIES


The interaction of an externally applied magnetic field and spins (either nuclear or electron).

ELECTRON PARAMAGNETIC RESONANCE

ENDOR: THEORY
EPR: APPLICATION
EPR: INTRODUCTION
EPR: PARALLEL MODE OPERATION
EPR: THEORY
EPR - INTERPRETATION
HYPERFINE SPLITTING
NUCLEAR MAGNETIC RESONANCE

NMR: BACKGROUND PHYSICS AND MATHEMATICS


NMR: EXPERIMENTAL
2D NMR

2D NMR: INDIRECT DETECTION


2D NMR BACKGROUND
2D NMR BASICS
2D NMR EXPERIMENTS
2D NMR INTRODUCTION
HETERONUCLEAR CORRELATIONS
HOMONUCLEAR CORRELATIONS
CARBON-13 NMR
INTERPRETING C-13 NMR SPECTRA
DIFFUSION ORDERED SPECTROSCOPY (DOSY)
MAGNETIC RESONANCE IMAGING
NMR - INTERPRETATION

INTEGRATION IN NMR
PASCAL’S TRIANGLE
NMR HARDWARE
NMR SPECTROSCOPY IN LAB: COMPLICATIONS
PULSE SEQUENCES
PHASE CYCLING AND COHERENCE TRANSFER PATHWAYS
SOLID STATE EXPERIMENTS

1 7/7/2021
CAR-PURCELL-MEIBOOM-GILL (CPMG) ECHO TRAIN ACQUISITION
MAGIC ANGLE HOPPING (MAH)
SOLID STATE NMR EXPERIMENTAL SETUP
NMR: STRUCTURAL ASSIGNMENT
(N+1) RULE
BACKGROUND TO C-13 NMR
DETERMINE STRUCTURE WITH COMBINED SPECTRA
HIGH RESOLUTION PROTON NMR SPECTRA
INTEGRATION IN PROTON NMR
INTERPRETING C-13 NMR SPECTRA
INTRODUCTION TO PROTON NMR
LOW RESOLUTION PROTON NMR SPECTRA
MORE ABOUT ELECTRONICS
MULTIPLICITY IN PROTON NMR
NMR11. MORE ABOUT MULTIPLICITY
NMR14. MORE PRACTICE WITH NMR SPECTROSCOPY
NMR2. CARBON-13 NMR
NMR3. SYMMETRY IN NMR
NMR4. 13C NMR AND GEOMETRY
NMR5. 13C NMR AND ELECTRONICS
NMR8. CHEMICAL SHIFT IN 1H NMR
NMR APPENDIX. USEFUL CHARTS FOR NMR IDENTIFICATION
NMR - THEORY

BLOCH EQUATIONS
LARMOR PRECESSION
NMR INTERACTIONS
CHEMICAL SHIFT (SHIELDING)
DIPOLAR COUPLING
J-COUPLING (SCALAR)
PASCAL’S TRIANGLE CONSTRUCTION
QUADRUPOLAR COUPLING
QUANTUM MECHANIC TREATMENT
RELAXATION
NMR: KINETICS
NUCLEAR OVERHAUSER EFFECT
SOLOMON EQUATIONS
SPIN-SPIN RELAXATION
SPIN LATTICE RELAXATION
ROTATIONS AND IRREDUCIBLE TENSOR OPERATORS
NMR: INTRODUCTION
BACK MATTER

INDEX

ROTATIONAL SPECTROSCOPY
Rotational spectroscopy is concerned with the measurement of the energies of transitions between quantized rotational states of
molecules in the gas phase. The spectra of polar molecules can be measured in absorption or emission by microwave spectroscopy or by
far infrared spectroscopy.

MICROWAVE ROTATIONAL SPECTROSCOPY


ROTATIONAL SPECTROSCOPY OF DIATOMIC MOLECULES
ROTATION OF LINEAR MOLECULES
ROVIBRATIONAL SPECTROSCOPY

VIBRATIONAL SPECTROSCOPY
Infrared spectroscopy (IR spectroscopy or Vibrational Spectroscopy) is the spectroscopy that deals with the infrared region of the
electromagnetic spectrum, that is light with a longer wavelength and lower frequency than visible light. It covers a range of techniques,
mostly based on absorption spectroscopy.

2 7/7/2021
VIBRATIONAL MODES
COMBINATION BANDS, OVERTONES AND FERMI RESONANCES
INTRODUCTION TO VIBRATIONS
ISOTOPE EFFECTS IN VIBRATIONAL SPECTROSCOPY
MODE ANALYSIS
NORMAL MODES
NUMBER OF VIBRATIONAL MODES IN A MOLECULE
SYMMETRY ADAPTED LINEAR COMBINATIONS
BACK MATTER

INDEX
INFRARED SPECTROSCOPY
HOW AN FTIR SPECTROMETER OPERATES
IDENTIFYING THE PRESENCE OF PARTICULAR GROUPS
INFRARED: APPLICATION
INFRARED: INTERPRETATION
INFRARED SPECTROSCOPY
INTERPRETING INFRARED SPECTRA
CARBON NITROGEN BONDS
IR10. MORE PRACTICE WITH IR SPECTRA
IR11. APPENDIX: IR TABLE OF ORGANIC COMPOUNDS
IR2. HYDROCARBON SPECTRA
IR3. SUBTLE POINTS OF IR SPECTROSCOPY
IR4. CARBON CARBON MULTIPLE BONDS
IR5. CARBON OXYGEN SINGLE BONDS
IR6. CARBON OXYGEN DOUBLE BONDS
IR8. MORE COMPLICATED IR SPECTRA
IR9. MISLEADING PEAKS
WHAT DOES AN IR SPECTRUM LOOK LIKE?
IR SPECTROSCOPY BACKGROUND
THE FINGERPRINT REGION
BACK MATTER

INDEX
RAMAN SPECTROSCOPY
RAMAN: APPLICATION
RAMAN: THEORY
RESONANT VS. NONRESONANT RAMAN SPECTROSCOPY

ELECTRONIC SPECTROSCOPY
Electron spectroscopy is an analytical technique to study the electronic structure and its dynamics in atoms and molecules. In general an
excitation source such as x-rays, electrons or synchrotron radiation will eject an electron from an inner-shell orbital of an atom.

CIRCULAR DICHROISM
ELECTRONIC SPECTROSCOPY: APPLICATION
ELECTRONIC SPECTROSCOPY - INTERPRETATION
ELECTRONIC SPECTROSCOPY BASICS
A DOUBLE BEAM ABSORPTION SPECTROMETER
BONDING THEORY FOR UV-VISIBLE ABSORPTION SPECTRA
ELECTROMAGNETIC RADIATION
THE BEER-LAMBERT LAW
USING UV-VISIBLE ABSORPTION SPECTROSCOPY
WHAT CAUSES MOLECULES TO ABSORB UV AND VISIBLE LIGHT
FLUORESCENCE AND PHOSPHORESCENCE
JABLONSKI DIAGRAM
JABLONSKI DIAGRAM: SHOCKWAVE
METAL TO LIGAND AND LIGAND TO METAL CHARGE TRANSFER BANDS
RADIATIVE DECAY

3 7/7/2021
CHEMILUMINESCENCE
FLUORESCENCE
PHOSPHORESCENCE
SELECTION RULES FOR ELECTRONIC SPECTRA OF TRANSITION METAL COMPLEXES
DERIVATION OF LAPORTE RULE
SPIN-ORBIT COUPLING
ATOMIC TERM SYMBOLS
MOLECULAR TERM SYMBOLS
THE RUSSELL SAUNDERS COUPLING SCHEME
TWO-PHOTON ABSORPTION

PHOTOELECTRON SPECTROSCOPY
Photoelectron spectroscopy involves the measurement of kinetic energy of photoelectrons to determine the bonding energy,intensity and
angular distributions of these electrons and use the information obtained to examine the electronic structure of molecules.

PHOTOELECTRON SPECTROSCOPY: APPLICATION


PHOTOELECTRON SPECTROSCOPY: THEORY

X-RAY SPECTROSCOPY
EXAFS: THEORY
X-RAYS
XANES: APPLICATION
XANES - THEORY

XANES - THEORY II
XAS - THEORY
BACK MATTER

INDEX

PHOTOACOUSTIC SPECTROSCOPY
Photoacoustic spectroscopy (PAS) uses acoustic waves produced from materials which are exposed to light to measure its concentration.
PAS is unique in that it combines heat measurements with optical microscopy. This makes PAS highly accurate and useful for sensitive
detectors. Interest sparked after Alexandar Graham Bell wrote about his findings when he discovered the acoustic effect in 1880.

MÖSSBAUER SPECTROSCOPY
Mössbauer spectroscopy is a versatile technique used to study nuclear structure with the absorption and re-emission of gamma rays. The
technique uses a combination of the Mössbauer effect and Doppler shifts to probe the hyperfine transitions between the excited and
ground states of the nucleus. Many isotopes exhibit Mössbauer characteristics but the most commonly studied isotope is Fe-57.

4 7/7/2021
CHAPTER OVERVIEW
FUNDAMENTALS OF SPECTROSCOPY

ELECTROMAGNETIC RADIATION
INTRODUCTION TO SPECTROSCOPY
LINESHAPE FUNCTIONS
Four different mathematical descriptions of the lineshape for an absorptive transition are discussed
below. These transitions may involve electronic, rotational, or vibrational (i.e. visible, microwave
or infrared radiation) eigenstates.

SELECTION RULES
A selection rule describes how the probability of transitioning from one level to another cannot be
zero. It has two sub-pieces: a gross selection rule and a specific selection rule. A gross selection
rule illustrates characteristic requirements for atoms or molecules to display a spectrum of a given kind, such as an IR spectroscopy or
a microwave spectroscopy.

SELECTION RULES AND TRANSITION MOMENT INTEGRAL


In chemistry and physics, selection rules define the transition probability from one eigenstate to another eigenstate. In this topic, we
are going to discuss the transition moment, which is the key to understanding the intrinsic transition probabilities. Selection rules have
been divided into the electronic selection rules, vibrational selection rules (including Franck-Condon principle and vibronic coupling),
and rotational selection rules.

THE POWER OF THE FOURIER TRANSFORM FOR SPECTROSCOPISTS


TIME-RESOLVED VS. FREQUENCY RESOLVED
BACK MATTER

INDEX

1 7/7/2021
Electromagnetic Radiation
As you read the print off this computer screen now, you are reading pages of fluctuating energy and magnetic fields. Light,
electricity, and magnetism are all different forms of electromagnetic radiation.

Contributors and Attributions


Nikita Patel (UCD), Kevin Vo (UCD), Mateo Hernandez (UCD)

6/25/2021 1 https://chem.libretexts.org/@go/page/1779
Introduction to Spectroscopy
Realizing that light may be considered to have both wave-like and particle-like characteristics, it is useful to consider that a
given frequency or wavelength of light is associated with a "light quanta" of energy we now call a photon. As noted in the
following equations, frequency and energy change proportionally, but wavelength has an inverse relationship to these
quantities.
c
ν = (1)
λ

and
ΔE = hν (2)

with
ν is the frequency of light
λ is the wavelength of ligt
c is the speed of light (3 × 10 m/sec )
8

ΔE is the transition energy (difference of energies between the initial and final states)

h is Planck's constant (h = 6.626069 × 10 J s)


−34

To "see" a molecule, we must use light having a wavelength smaller than the molecule itself (roughly 1 to 15 angstroms). Such
radiation is found in the X-ray region of the spectrum, and the field of X-ray crystallography yields remarkably detailed
pictures of molecular structures amenable to examination. The chief limiting factor here is the need for high quality crystals of
the compound being studied. The methods of X-ray crystallography are too complex to be described here; nevertheless, as
automatic instrumentation and data handling techniques improve, it will undoubtedly prove to be the procedure of choice for
structure determination. The spectroscopic techniques described below do not provide a three-dimensional picture of a
molecule, but instead yield information about certain characteristic features. A brief summary of this information follows:
Ultraviolet-Visible Spectroscopy: Absorption of this relatively high-energy light causes electronic excitation. The easily
accessible part of this region (wavelengths of 200 to 800 nm) shows absorption only if conjugated π electron systems are
present.
Infrared Spectroscopy: Absorption of this lower energy radiation causes vibrational and rotational excitation of groups of
atoms. within the molecule. Because of their characteristic absorptions, identification of functional groups is easily
accomplished.
Nuclear Magnetic Resonance (NMR) Spectroscopy: Absorption in the low-energy radio-frequency part of the spectrum
causes excitation of nuclear spin states. NMR spectrometers are tuned to certain nuclei (e.g. 1H, 13C, 19F & 31P). For a
given type of nucleus, high-resolution spectroscopy distinguishes and counts atoms in different locations in the molecule.

Contributors
William Reusch, Professor Emeritus (Michigan State U.), Virtual Textbook of Organic Chemistry

6/6/2021 1 https://chem.libretexts.org/@go/page/1781
Lineshape Functions
Four different mathematical descriptions of the lineshape for an absorptive transition are discussed below. These transitions
may involve electronic, rotational, or vibrational (i.e. visible, microwave or infrared radiation) eigenstates.

Contributors and Attributions


Stefan Franzen (North Carolina State University)

6/21/2021 1 https://chem.libretexts.org/@go/page/1782
Selection Rules
A selection rule describes how the probability of transitioning from one level to another cannot be zero. It has two sub-pieces:
a gross selection rule and a specific selection rule. A gross selection rule illustrates characteristic requirements for atoms or
molecules to display a spectrum of a given kind, such as an IR spectroscopy or a microwave spectroscopy. Once the atom or
molecules follow the gross selection rule, the specific selection rule must be applied to the atom or molecules to determine
whether a certain transition in quantum number may happen or not.
Selection rules specify the possible transitions among quantum levels due to absorption or emission of electromagnetic
radiation. Incident electromagnetic radiation presents an oscillating electric field E cos(ωt) that interacts with a transition 0

dipole. The dipole operator is μ = e ⋅ r where r is a vector pointing in a direction of space.


A dipole moment of a given state is


μz = ∫ ψ1 μz ψ1 dτ (1)

A transition dipole moment is a transient dipolar polarization created by an interaction of electromagnetic radiation with a
molecule


(μz )12 = ∫ ψ1 μz ψ2 dτ (2)

In an experiment we present an electric field along the z axis (in the laboratory frame) and we may consider specifically the
interaction between the transition dipole along the x, y, or z axis of the molecule with this radiation. If μ is zero then a z

transition is forbidden. The selection rule is a statement of when μ is non-zero. z

We can consider selection rules for electronic, rotational, and vibrational transitions.

Electronic transitions
We consider a hydrogen atom. In order to observe emission of radiation from two states mu must be non-zero. That is z


(μz )12 = ∫ ψ e ⋅ z ψ2 dτ ≠ 0 (3)
1

For example, is the transition from ψ 1s to ψ 2s allowed?


(μz )12 = ∫ ψ1s e ⋅ z ψ2s dτ (4)

Using the fact that z = r cosq in spherical polar coordinates we have


r
−r/a0 −r/a0 2
(μz )12 = e ∭ e r cos θ(2 − )e r sin θdrdθ dϕ (5)
a0

We can consider each of the three integrals separately.


∞ π 2π
−r/a0
r −r/a0 2
∫ e r(2 − )e r dr ∫ cos θ sin θ dθ ∫ dϕ (6)
0
a0 0 0

If any one of these is non-zero the transition is not allowed. We can see specifically that we should consider the q integral. We
make the substitution x = cos q, dx = − sin q dq and the integral becomes
−1 2 −1
x ∣
−∫ xdx = − ∣ =0 (7)
1
2 ∣
1

which is zero. The result is an even function evaluated over odd limits. In a similar fashion we can show that transitions along
the x or y axes are not allowed either. This presents a selection rule that transitions are forbidden for Δl = 0 . For electronic
transitions the selection rules turn out to be Δl = ±1 and Δm = 0 . These result from the integrals over spherical harmonics

6/21/2021 1 https://chem.libretexts.org/@go/page/1783
which are the same for rigid rotator wavefunctions. We will prove the selection rules for rotational transitions keeping in mind
that they are also valid for electronic transitions.

Rotational transitions
We can use the definition of the transition moment and the spherical harmonics to derive selection rules for a rigid rotator.
Once again we assume that radiation is along the z axis.
\[(\mu_z)_{J,M,{J}',{M}'}=\int_{0}^{2\pi } \int_{0}^{\pi }Y_{J'}^{M'}(\theta,\phi )\mu_zY_{J}^{M}
(\theta,\phi)\sin\theta\,d\phi,d\theta\\]
Notice that m must be non-zero in order for the transition moment to be non-zero. This proves that a molecule must have a
permanent dipole moment in order to have a rotational spectrum. The spherical harmonics can be written as
|M| iMϕ
M
Y (θ, ϕ) = N JM P (cos θ)e (8)
J J

where N is a normalization constant. Using the standard substitution of


JM x = cos q we can express the rotational transition
moment as
2π 1
′ ′
I(M−M )ϕ |M | |M|
(μz )J,M, J ′ , M ′ = μ N JM N ′
J M
′ ∫ e dϕ ∫ P ′ (x)P (x)dx (9)
J J
0 −1

The integral over f is zero unless M = M' so ΔM = 0 is part of the rigid rotator selection rule. Integration over ϕ for
M =M gives 2π so we have

1

|M | |M|
(μz )J,M, J ′ , M ′ = 2πμ N JM N J M
′ ′ ∫ P ′ (x)P (x)dx (10)
J J
−1

We can evaluate this integral using the identity


|M] |M| |M|
(2J + 1)x P (x) = (J − |M | + 1)P (x) + (J − |M |)P (x) (11)
J J+1 J−1

Substituting into the integral one obtains an integral which will vanish unless J ′
= J +1 or J ′
= J −1 .
1
|M |
′ (J − |M | + 1) |M|
(J − |M |) |M|
∫ P ′
(x)( P (x) + P (x))dx (12)
J J+1 J−1
−1 (2J + 1) (2J + 1)

This leads to the selection rule ΔJ = ±1 for absorptive rotational transitions. Keep in mind the physical interpretation of the
quantum numbers J and M as the total angular momentum and z-component of angular momentum, respectively. As stated
above in the section on electronic transitions, these selection rules also apply to the orbital angular momentum (Δl = ±1 ,
Δm = 0 ).

Vibrational transitions
The harmonic oscillator wavefunctions are
2
1/2 −α q /2
ψ v (q) =N v H v (α q)e (13)

where H v (a1/2q) is a Hermite polynomial and a = (km/á2)1/2.


The transition dipole moment for electromagnetic radiation polarized along the z axis is

2 2
1/2 −α q /2 1/2 −α q /2
(μz )v,v′ = ∫ N v N v′ H v′ (α q)e H μz (α q)e dq (14)
−∞

Note that we continue to use the general coordinate q although this can be z if the dipole moment of the molecule is aligned
along the z axis. The transition moment can be expanded about the equilibrium nuclear separation.
∂μ
μz (q) = μ0 +( )q+. . . . . (15)
∂q

6/21/2021 2 https://chem.libretexts.org/@go/page/1783
where m0 is the dipole moment at the equilibrium bond length and q is the displacement from that equilibrium state. From the
first two terms in the expansion we have for the first term

2 2
1/2 −α q /2 1/2 −α q /2
(μz )v,v′ = μ0 ∫ N v N v′ H v′ (α q)e Hv (α q)e dq (16)
−∞

This term is zero unless v = v’ and in that case there is no transition since the quantum number has not changed.

∂μ 1/2 −α q
2
/2 1/2 −α q
2
/2
(μz )v,v′ =( )∫ N vN v
′ H v
′ (α q)e Hv (α q)e dq (17)
∂q −∞

This integral can be evaluated using the Hermite polynomial identity known as a recursion relation
1
x Hv (x) = vHv−1 (x) + Hv+1 (x) (18)
2

where x = Öaq. If we now substitute the recursion relation into the integral we find
N vN v
′ ∂μ
(μz )v,v′ = −
− ( ) (19)
√α ∂q


1/2 −α q
2
/2 1/2
1 1/2
∫ H ′
v (α q)e (vHv−1 (α q) + Hv+1 (α q))dq (20)
−∞
2

which will be non-zero if v’ = v – 1 or v’ = v + 1. Thus, we see the origin of the vibrational transition selection rule that v = ±
1. We also see that vibrational transitions will only occur if the dipole moment changes as a function nuclear motion.

Contributors and Attributions


Stefan Franzen (North Carolina State University)

6/21/2021 3 https://chem.libretexts.org/@go/page/1783
Selection rules and transition moment integral
In chemistry and physics, selection rules define the transition probability from one eigenstate to another eigenstate. In this
topic, we are going to discuss the transition moment, which is the key to understanding the intrinsic transition probabilities.
Selection rules have been divided into the electronic selection rules, vibrational selection rules (including Franck-Condon
principle and vibronic coupling), and rotational selection rules.

Introduction
The transition probability is defined as the probability of particular spectroscopic transition to take place. When an atom or
molecule absorbs a photon, the probability of an atom or molecule to transit from one energy level to another depends on two
things: the nature of initial and final state wavefunctions and how strongly photons interact with an eigenstate. Transition
strengths are used to describe transition probability. Selection rules are utilized to determine whether a transition is allowed or
not. Electronic dipole transitions are by far the most important for the topics covered in this module.

Transition Moment
In an atom or molecule, an electromagnetic wave (for example, visible light) can induce an oscillating electric or magnetic
moment. If the frequency of the induced electric or magnetic moment is the same as the energy difference between one
eigenstate Ψ1 and another eigenstate Ψ2, the interaction between an atom or molecule and the electromagnetic field is resonant
(which means these two have the same frequency). Typically, the amplitude of this (electric or magnetic) moment is called the
transition moment. In quantum mechanics, the transition probability of one molecule from one eigenstate Ψ1 to another
eigenstate Ψ2 is given by |M⃗  |2, and
21
⃗ 
M 21 is called the transition dipole moment, or transition moment, from Ψ1 to Ψ2. In
mathematical form it can be written as

⃗ 
M 21 = ∫ Ψ2 μ⃗ Ψ1 dτ (1)

The Ψ1 and Ψ2 are two different eigenstates in one molecule, M⃗  is the electric dipole moment operator. If we have a system
21

with n molecules and each has charge Qn, and the dipole moment operator is can be written as

μ⃗ = ∑ Qn x⃗ n (2)

the x⃗  is the position vector operator.


n

Transition Moment Integral


Based on the Born-Oppenheimer approximation, the fast electronic motion can be separated from the much slower motion of
the nuclei. As a result, the total wavefunction can be separated into electronic, vibrational, and rotational parts:
Ψ(r, R) = ψe (r, Re )ψv (R)ψr (R)

The Born-Oppenheimer approximation assumes that the electronic wavefunction, ψ , is approximated in all electronic
e

coordinates at the equilibrium nuclear coordinates (Re). Since mass of electrons is much smaller than nuclear mass, the
rotational wavefunction, ψ , only depends on nuclear coordinates. The rotational wavefunction could provide important
r

information for rotational selection rules, but we will not consider the rotational wavefuntion any further for simplicity
because most of the spectra are not rotationally resolved. With the rotational part removed, the transition moment integral can
be expressed as
′ ′ ′′ ′′
 M = ∬ ψe (r, Re ) ⋅ ψv (R)(μe + μn )ψe (r, Re ) ⋅ ψv (R)drdR

where the prime and double prime represent the upper and lower states respectively. Both the nuclear and electronic parts
contribute to the dipole moment operator. The above equation can be integrated by two parts, with μ and μ respectively. A
n e

product of two integral is obtained:


′ ′′ ′ ′′ ′ ′′ ′ ′′
 M = ∫ ψe (r, Re ) ⋅ μe ⋅ ψe (r, Re )dr ∫ ψv (R) ⋅ ψv (R)dR + ∫ ψv (R) ⋅ μn ⋅ ψv (R)dR ∫ ψe (r, Re )ψe (r, Re )dr

Because different electronic wavefunctions must be orthogonal to each other, hence ′ ′′


∫ ψe (r, Re )ψe (r, Re )dr is zero, the
second part of the integral should be zero.

6/27/2021 1 https://chem.libretexts.org/@go/page/1784
The transition moment integral can be simplified as
′ ′′ ′ ′′
 M = ∫ ψe (r, Re ) ⋅ μe ⋅ ψe (r, Re )dr ∫ ψv (R) ⋅ ψv (R)dR

The above equation is of great importance because the first integral defines the electronic selection rules, while the second
integral is the basis of vibrational selection rules.

Electronic Selection Rules


In atoms
Atoms are described by the primary quantum number n, angular momentum quantum number L, spin quantum number S, and
total angular momentum quantum number J. Based on Russell-Saunders approximation of electron coupling, the atomic term
symbol can be represented as 2S+1LJ.
1. The total spin cannot change, ΔS=0;
2. The change in total orbital angular momentum can be ΔL=0, ±1, but L=0 ↔ L=0 transition is not allowed;
3. The change in the total angular momentum can be ΔJ=0, ±1, but J=0 ↔ J=0 transition is not allowed;
4. The initial and final wavefunctions must change in parity. Parity is related to the orbital angular momentum summation
over all elections Σli, which can be even or odd; only even ↔ odd transitions are allowed.

In molecules
The electronic-state configurations for molecules can be described by the primary quantum number n, the angular momentum
quantum number Λ, the spin quantum number S, which remains a good quantum number, the quantum number Σ (S, S-1, ..., -
S), and the projection of the total angular momentum quantum number onto the molecular symmetry axis Ω, which can be
derived as Ω=Λ+Σ. The term symbol for the electronic states can be represented as
2S+1 (+/−)
Λ (3)
Ω,(g/u)

Group theory makes great contributions to the prediction of the electronic selection rules for many molecules. An example is
used to illustrate the possibility of electronic transitions via group theory.
1. The total spin cannot change, ΔS=0; the rule ΔΣ=0 holds for multiplets;
If the spin-orbit coupling is not large, the electronic spin wavefunction can be separated from the electronic
wavefunctions. Since the electron spin is a magnetic effect, electronic dipole transitions will not alter the electron spin.
As a result, the spin multiplicity should not change during the electronic dipole transition.
2. The total orbital angular momentum change should be ΔΛ=0, ±1;
For heteronuclear diatomic molecules with  C ∞vsymmetry, a Σ+ ↔ Π transition is allowed according to ΔΛ selection
rule. In order to prove the allowance of this transition, the direct product of Σ ⊗ Π yields Π irreducible
+

representation from the direct product table. Based on the  C character table below, the operator in the x and y
∞v

direction have doubly degenerate Π symmetry. Therefore, the transition between Σ+ ↔ Π must be allowed since the
multiplication of any irreducible representation with itself will provide the totally symmetric representation. The
electronic transition moment integral can be nonzero. We can use the same kind of argument to illustrate that Σ ↔ Φ −

transition is forbidden.
C∞v character table
C∞v E 2C∞Φ ... ∞σv

Σ+ 1 1 ... 1 z (x2+y2), z2

Σ- 1 1 ... -1 Rz

Π 2 2 cos(Φ) ... 0 (x, y), (Rx, Ry) xz, yz

Δ 2 2 cos(2Φ) ... 0 (x2-y2), xy

Γ 2 2 cos(3Φ) ... 0

... ... ... ... ...

6/27/2021 2 https://chem.libretexts.org/@go/page/1784
3. Parity conditions are related to the symmetry of the molecular wavefunction reflecting against its symmetry axis. For
homonuclear molecules, the g ↔ u transition is allowed. For heteronuclear molecules, + ↔ + and - ↔ - transitions apply;
For hetero diatomic molecules with  C ∞v symmetry, we can use group theory to reveal that Σ ↔ Σ and Σ ↔ Σ
+ + − −

transitions are allowed, while Σ ↔ Σ transitions are forbidden. The direct product of either Σ ⊗ Σ or Σ ⊗ Σ
+ − + + − −

yields the same irreducible representation Σ+ based on the direct product table. The z component of the dipole operator has
Σ+ symmetry. The transition is allowed because the electronic transition moment integral can generate the totally
symmetric irreducible representation Σ+. We can use the same method to prove that Σ ↔ Σ transitions are forbidden.
+ −

Similarly, for a molecule with an inversion center, a subscript g or u is used to reveal the molecular symmetry with respect
to the inversion operation, i. The x, y, and z components of the transition dipole moment operator have u inversion
symmetry in molecule with inversion center (g and u are short for gerade and ungerade in German, meaning even and odd).
The electronic transition moment integral need to yield totally symmetric irreducible representation Ag for allowed
transitions. Therefore, only g ↔ u transition is allowed.

Vibrational Selection rules


1. Transitions with Δv=±1, ±2, ... are all allowed for anharmonic potential, but the intensity of the peaks become weaker as
Δv increases.
2. v=0 to v=1 transition is normally called the fundamental vibration, while those with larger Δv are called overtones.
3. Δv=0 transition is allowed between the lower and upper electronic states with energy E1 and E2 are involved, i.e. (E1, v''=n)
→ (E2, v'=n), where the double prime and prime indicate the lower and upper quantum state.

The geometry of vibrational wavefunctions plays an important role in vibrational selection rules. For diatomic molecules, the
vibrational wavefunction is symmetric with respect to all the electronic states. Therefore, the Franck-Condon integral is always
totally symmetric for diatomic molecules. The vibrational selection rule does not exist for diatomic molecules.
For polyatomic molecules, the nonlinear molecules possess 3N-6 normal vibrational modes, while linear molecules possess
3N-5 vibrational modes. Based on the harmonic oscillator model, the product of 3N-6 normal mode wavefunctions contribute
to the total vibrational wavefunction, i.e.

ψvib = ∏ ψ1 ψ2 ψ3 . . . ψ3N −6 (4)

3N −6

where each normal mode is represented by the wavefunction ψ . Comparing to the Franck-Condon factor for diatomic
i

molecules with single vibrational overlap integral, a product of 3N-6 (3N-5 for linear molecules) overlap integrals needs to be
evaluated. Based on the symmetry of each normal vibrational mode, polyatomic vibrational wavefunctions can be totally
symmetric or non-totally symmetric. If a normal mode is totally symmetric, the vibrational wavefunction is totally symmetric
with respect to all the vibrational quantum number v. If a normal mode is non-totally symmetric, the vibrational wavefunction
alternates between symmetric and non-symmetric wavefunctions as v alternates between even and odd number.
If a particular normal mode in both the upper and lower electronic state is totally symmetric, the vibrational wavefunction for
the upper and lower electronic state will be symmetric, resulting in the totally symmetric integrand in the Franck-Condon
integral. If the vibrational wavefunction of either the lower or upper electronic state is non-totally symmetric, the Franck-
Condon integrand will be non-totally symmetric.
We will use CO2 as an example to specify the vibrational selection rule. CO2 has four vibrational modes as a linear molecule.
The vibrational normal modes are illustrated in the figure below:

6/27/2021 3 https://chem.libretexts.org/@go/page/1784
The vibrational wavefunction for the totally symmetric C-O stretch, v1, is totally symmetric with respect to all the vibrational
quantum numbers. However, the vibrational wavefunctions for the doubly degenerate bending modes, v2, and the
antisymmetric C-O stretch, v3, are non-totally symmetric. Therefore, the vibrational wavefunctions are totally symmetric for
even vibrational quantum numbers (v=0, 2, 4...), while the wavefunctions remain non-totally symmetric for v odd (v=1, 3,
5...).
Therefore, any value of Δv1 is possible between the upper and lower electronic state for mode v1. On the other hand, modes v2
and v3 include non-totally symmetric vibrational wavefunctions, so the vibrational quantum number can only change evenly,
such as Δv=± 2, ± 4, etc..

Franck-Condon Principle
Franck-Condon principle was proposed by German physicist James Franck (1882-1964) and U.S. physicist Edward U. Condon
(1902-1974) in 1926. This principle states that when an electronic transition takes place, the time scale of this transition is so
fast compared to nucleus motion that we can consider the nucleus to be static, and the vibrational transition from one
vibrational state to another state is more likely to happen if these states have a large overlap. It successfully explains the reason
why certain peaks in a spectrum are strong while others are weak (or even not observed) in absorption spectroscopy.
′ ′′ ′ ′′
 M = ∫ ψe (r, Re ) ⋅ μe ⋅ ψe (r, Re )dr ∫ ψv (R) ⋅ ψv (R)dR

The second integral in the above equation is the vibrational overlap integral between one eigenstate and another eigenstate. In
addition, the square of this integral is called the Franck-Condon factor:

′ ′′ 2
 F ranck − C ondon F actor =∣ ∫ ψv ψv dR∣

It governs the vibrational transition contribution to the transition probability and shows that in order to have a large vibrational
contribution, the vibrational ground state and excited state must have a strong overlap.

6/27/2021 4 https://chem.libretexts.org/@go/page/1784
The above figure shows the Franck-Condon principle energy diagram, since electronic transition time scale is small compared
to nuclear motion, the vibrational transitions are favored when the vibrational transition have the smallest change of nuclear
coordinate, which is a vertical transition in the figure above. The electronic eigenstates favors the vibrational transition v'=0 in
the ground electronic state to v"=2 in the excited electronic state, while peak intensity of v'=0 to v"=0 transition is expected to
be low because the overlap between the v'=0 wavefunction and v''=0 wavefunction is very low.

The figure above is the photoelectron spectrum of the ionization of hydrogen molecule (H2), it is also a beautiful example to
formulate the Franck-Condon principle. It shows the appropriate energy curves and the vibrational energy levels, and with the
help of the Franck-Condon principle, the transition between ground vibrational state v0" and excited vibrational state v2' is
expected to be the most intense peak in the spectrum.

Vibronic Coupling
Why can some electronic-forbidden transitions be observed as weak bands in spectrum? It can be explained by the interaction
between the electronic and vibrational transitions. The word "vibronic" is the combination of the words "vibrational" and
"electronic." Because the energy required for one electronic state to another electronic state (electronic transition, usually in
the UV-Vis region) is larger than one vibrational state to another vibrational state (vibrational transition, usually in the IR
region), sometimes energy (a photon) can excite a molecule to an excited electronic and vibrational state. The bottom figure

6/27/2021 5 https://chem.libretexts.org/@go/page/1784
shows the pure electronic transition (no vibronic coupling) and the electronic transition couples with the vibrational transition
(vibronic coupling).

We now can go back to the original question: Why can some electronic-forbidden transitions be observed as weak bands in
spectrum? For example, the d-d transitions in the octahedral transition metal complexes are Laporte forbidden (same
symmetry, parity forbidden), but they can be observed in the spectrum and this phenomenon can be explained by vibronic
coupling.

Now we consider the Fe(OH2)62+ complex which has low spin d6 as ground state (see figure above). Let's examine the one-
electron excitation from t2g molecular orbital to eg molecular orbital. The octahedral complex has Oh symmetry and therefore
the ground state has A1g symmetry from the character table. The symmetry of the excited state, which is the direct product of
all singly occupied molecular orbitals (eg and t2g in this case):
Γt2g ⊗ Γeg = T1g + T2g

The transition moment integral for the electronic transition can be written as

⃗  ∗
M =∫ ψ μ⃗ ψdτ

where ψ is the electronic ground state and ψ' is the electronic excited state. The condition for the electronic transition to be
allowed is to make the transition moment integral nonzero. The μ^ is the transition moment operator, which is the symmetry of

the x^, y
^, z^ operators from the character table. For octahedral symmetry, these operators are degenerate and have T1u

symmetry. Therefore, we can see that both 1T1g←1A1g and 1T2g←1A1g are electronically forbidden by the direct product (do
not contain the totally symmetric A1g):
T1g ×  T1u ×  A1g = A1u + Eu + T1u + T2u

T2g ×  T1u ×  A1g = A2u + Eu + T1u + T2u

6/27/2021 6 https://chem.libretexts.org/@go/page/1784
For octahedral complex, there are 15 vibrational normal modes. From the Oh character table we can get these irreducible
representations:
Γvib = a1g + eg + 2 t1u + t2g + t2u

When we let the vibrational transition to couple with the electronic transition, the transition moment integral has the form:

⃗  ∗ ∗
M =∫ ψ ψ μ⃗ ψe ψv dτ
e′ v′

For the vibronic coupling to be allowed, the transition moment integral has to be nonzero. Use these vibrational symmetries of
the octahedral complex to couple with the electronic transition:
A1g × { T1g ×  T1u ×  A1g } = A1u + Eu + T1u + T2u

Eg × { T1g ×  T1u ×  A1g } = Eg × { A1u + Eu + T1u + T2u } = A1u + A2u + 2 Eu + 2 T1u + 2 T2u

T1u × { T1g ×  T1u ×  A1g } = T1u × { A1u + Eu + T1u + T2u } = A1g + …

T2g × { T1g ×  T1u ×  A1g } = T2g × { A1u + Eu + T1u + T2u } = A1u + A2u + 2 Eu + 3 T1u + 4 T2u

T2u × { T1g ×  T1u ×  A1g } = T2u × { A1u + Eu + T1u + T2u } = A1g + …

Since the t1u and t2u representation can generate the totally symmetric representation in the integrand. Therefore, we can see
that t1u and t2u can couple with the electronic transition to form the allowed vibronic transition. Therefore, the d-d transition
band for Fe(OH2)62+ complex can be observed through vibronic coupling.

Rotational Selection rules


1. Transitions with ΔJ=±1 are allowed;
Photons do not have any mass, but they have angular momentum. The conservation of angular momentum is the fundamental
criteria for spectroscopic transitions. As a result, the total angular momentum has to be conserved after a molecule absorbs or
emits a photon. The rotational selection rule relies on the fact that photon has one unit of quantized angular momentum.
During the photon emission and absorption process, the angular moment J cannot change by more than one unit.
Let's consider a single photon transition process for a diatomic molecule. The rotational selection rule requires that transitions
with ΔJ=±1 are allowed. Transitions with ΔJ=1 are defined as R branch transitions, while those with ΔJ=-1 are defined as P
branch transitions. Rotational transitions are conventional labeled as P or R with the rotational quantum number J of the lower
electronic state in the parentheses. For example, R(2) specifies the rotational transition from J=2 in the lower electronic state to
J=3 in the upper electronic state.
2. ΔJ=0 transitions are allowed when two different electronic or vibrational states are involved: (X'', J''=m) → (X', J'=m).
The Q branch transitions will only take place when there is a net orbital angular momentum in one of the electronic states.
Therefore, Q branch does not exist for   Σ ↔ Σ electronic transitions because Σ electronic state does not possess any net
1 1

orbital angular momentum. On the other hand, the Q branch will exist if one of the electronic states has angular momentum. In
this situation, the angular momentum of the photon will cancel out with the angular momentum of the electronic state, so the
transition will take place without any change in the rotational state.
The schematic of P, Q, and R branch transitions are shown below:

6/27/2021 7 https://chem.libretexts.org/@go/page/1784
With regard to closed-shell non-linear polyatomic molecules, the selection rules are more complicated than diatomic case. The
rotational quantum number J remains a good quantum number as the total angular momentum if we don't consider the nuclear
spin. Under the effect of single photon transition, the change of J is still limited to a maximum of ± 1 based on the
conservation of angular momentum. However, the possibility of Q branch is greatly enhanced irrelevant to the symmetry of the
lower and upper electronic states. The rotational quantum number K is introduced along the inertial axis. For polyatomic
molecules with symmetric top geometry, the transition moment is polarized along inertial axis. The selection rule becomes
ΔK=0.

References
1. Schwartz, S. E. J. Chem. Educ. 1973, 50, 608-610. doi:10.1021/ed050p608
2. Byers, L. R. J. Chem. Educ. 1975, 52, 790. doi:10.1021/ed052p790
3. Ahmed, F. J. Chem. Educ. 1987, 64, 427-428. doi:10.1021/ed064p427
4. Dunbrack, R. L. J. Chem. Educ. 1986, 63, 953-955. doi:10.1021/ed063p953
5. Standard, J. M., Clark, B. K. J. Chem. Educ. 1999, 76, 1363-1366. doi:10.1021/ed076p1363
6. Ellis, A. M., Fehér, M., & Wright, T. G. (2005). Transition probabilities. Electronic and photoelectron spectroscopy:
fundamentals and case studies(pp. 51-64). Cambridge: Cambridge University Press.

Problems
1. What are the differences between Born-Oppenheimer approximation and Franck-Condon principle?
2. In evaluating the transition moment integral, why there must be a totally symmetric irreducible representation to make the
transition allowed?
3. Because of the d-d transition, transition metal complexes are known for its various color. However, Mn2+ does not have
intense color (pale pink), why?
4. Assume a Pd2+ square planar complex (D4h symmetry), is the following electronic transition allowed? If not, can it gain
intensity by vibronic coupling?

5. What's the vibrational selection rule for polyatomic molecules?

6/27/2021 8 https://chem.libretexts.org/@go/page/1784
6. Under what condition can we see the Q branch transition?

Answers to Problems
1. The Born Oppenheimer approximation states that the fast electronic motion can be separated from the much slower motion
of the nuclei, so the total wavefunction can be separated into the electronic, vibrational and rotational part. The Franck-
Condon principle states during an electronic transition, the timescale for an electron to move from one orbital to another is
so short that the nuclear position does not change during the transition process. The Franck Condon factor is the square of
the overlap integral, which defines the vibrational contribution to the transition probability.
2. n the character table, only A1 or A1g are totally symmetric irreducible representations. When you are integrating an odd
function over space, the integrand will be 0. Only the integration of an even function over space will yield a non-zero
value, representing the possibility of the transition.
3. The ground state Mn2+ has a high spin d5 electronic configuration, which has one electron in each of the eg and t2g orbitals,
and according to the electronic transition selection rule, the total spin cannot change:

then we will find out it is impossible for a high spin d5 octahedral complex since it violates Pauli exclusion principle.
Therefore, the d-d transition for this is Laporte forbidden and also spin forbidden, resulting in a pale color.
4. The symmetry of the excited state will have the symmetry:
Γa ⊗ Γb = B1g (5)
1g 1g

In the point group of D4h, the dipole moment operator transforms as eu(x,y) and a2u(z). Then by evaluating the transition
moment integral for the electronic transition, we can easily find out that there is no totally symmetric irreducible
representation in the integrand.Therefore, the electronic transition is not allowed:
Eu Eu
B1g × ( ) ×  A1g = ( ) (6)
A2u B2u

Next step we examine the vibrational irreducible representation a D4h complex has:
Γvib = a1g + b1g + b2g + a2u + b2u + 2 eu (7)

we can see that eu and b2u can potentially serve as the promoting mode for vibronic transition, that is, generating the totally
symmetric A1g in the integrand. Therefore, this electronic transition for D4h complex is forbidden, however, the transition can
be allowed through vibronic coupling.
5. If the vibrational mode is totally symmetric, any change in the vibrational quantum number v is possible between the lower
electronic state and upper electronic state. If the vibrational mode is non-totally symmetric, the vibrational wavefunction for
the lower and upper electronic states are non-totally symmetric. The vibrational quantum number can only change evenly, such
as Δv=± 2, ± 4, etc..
6.The Q branch transitions will only take place when there is a net orbital angular momentum in one of the electronic states.

Contributors and Attributions


Chun-Yi Lin (University of California, Davis), Zhou Lu (University of California, Davis)

6/27/2021 9 https://chem.libretexts.org/@go/page/1784
The Power of the Fourier Transform for Spectroscopists
Fourier transform is a mathematical technique that can be used to transform a function from one real variable to another. It is a
unique powerful tool for spectroscopists because a variety of spectroscopic studies are dealing with electromagnetic waves
covering a wide range of frequency. In Fourier transform term  e   , when x represents frequency, the corresponding y is
−2πixy

time. This provides an alternate way to process signal in time domain instead of the conventional frequency domain. To realize
this idea, Fourier transform from time domain to frequency domain is the essential process that enable us to translate raw data
to readable spectra. Recent prosperity of Fourier transform in spectroscopy should also attribute to the development of
efficient Fast Fourier Transform algorithm.

Introduction
The nature of trigonometric function enables Fourier transform to convert a function from the domain of one variable to
another and reconstruct it later on. This is a robust mathematical tool to process data in different domains under different
circumstances. Taking this principal idea and applying it in spectroscopy showed many impressive results in the early stage,
which in other ways are very difficult to resolve. These benefits triggered a wide exploration of Fourier transform based
methodology in a variety of spectroscopic techniques. At the same time, Fourier transform spectroscopic instruments are
developed with great efforts by physicists and engineers. All these factors give rise to the wide use of Fourier transform
spectroscopy.
In the following topics, the relevant mathematical background, the implementation of Fourier transform in spectroscopy and a
brief overview of various Fourier transform Spectrometers will be addressed in sequence.

Fourier Series
The motivation of Fourier transform arises from Fourier series, which was proposed by French mathematician and physicist
Joseph Fourier when he tried to analyze the flow and the distribution of energy in solid bodies at the turn of the 19th century.
He claimed that the temperature distribution could be described as an infinite series of sines and cosines of the form shown in
equation (1):

a0 nπx nπx
f (x) = + ∑ (an cos + bn sin ) (1)
2 L L
n=1

It turns out that this combination of sines and cosines series can be used to express any periodical function. As n increases, the
series will approach to the original function more closely.
If we use Euler's Identity (equation 2), as well as the exponential representations of sine (equation 3} and cosine (equation 4)
in our Fourier Series, we will find a natural redefinition of our coefficients a and b into a single complex coefficient C_n
n n

(equation 5).

e = cos θ + i sin θ (2)

1 iθ −iθ
cos θ = (e +e ) (3)
2

1 iθ −iθ
sin θ = (e −e ) (4)
2i

1
Cn = (Cn − i bn ) (5)
2

Using a bit of clever mathematics (complex conjugation, properties of odd function, rearranging summation set) we can
represent our original Fourier series in terms of a complex exponential, shown as equation 6.


f (x) = ∑ Cn e (6a)

n=−∞

2πnx
θ = (6b)
L

7/4/2021 1 https://chem.libretexts.org/@go/page/1785
Writing the Fourier series in this exponential form helps to simplify many formulas and expressions involved in the
transformation.

Fourier Transform
Then we can consider an extreme case, when L in equation 1, the summation becomes an integral as shown in equation (4)
∞ ∞ ∞
iπx iπx iπxn
n n
f (x) = ∑ cn e L
=∫ cn e L
dn = ∫ cn e L
dn (7)

−∞ −∞ −∞

This naturally gives the Fourier transform pair of f(x) and F(y). The relationships are shown below :
+∞
−i2πyx
F (y) = ∫ f (x)e dx (8a)
−∞

+∞
i2πxy
f (x) = ∫ F (y)e dx (8b)
−∞

Another important formula widely used to benefit In other cases, it is used to simplify the integral in the Fourier transform
based on the symmetry of the function. But so far, all these are just about mathematics. Its story with spectroscopy should start
from the mathematical description of electromagnetic waves.

Mathematical description of electromagnetic waves


Maxwell–Faraday equation and Ampère's circuital law give us electromagnetic wave equations to describe the characteristics
of an electromagnetic wave.[1] Using the linearity of Maxwell's equations in a vacuum, the solutions of the equation can be
decomposed into a superposition of sinusoids as shown below[3]:

→ →
E(r, t) = E0 cos(2πf t − k ⋅ r + ϕ0 ) (9a)


→ →
B(r, t) = B0 cos(2πf t − k ⋅ r + ϕ0 ) (9b)

Where t is time, f is the frequency, k=(kx,ky,kz) is the wave vector and ϕ is the phase angle.
This indicates that electromagnetic wave can be written as the sum of trigonometric functions with specific frequencies.
Scientists already discovered the fact that frequency and time is a classic Fourier transform pair in Fourier transform
relationship. All the Fourier transform pairs are connected by the Fourier transform term e . Regarding this case, we can−i2πyx

use the term to transform between two variables in this pair, namely time and frequency. In this way, we can measure the
properties of the electromagnetic wave in both conventional frequency domain and somehow more robust time domain.

Applying Fourier Transform-Fourier Transform Spectroscopy


Fourier transform are widely involved in spectroscopy in all research areas that require high accuracy, sensitivity, and
resolution. All these spectroscopic techniques using Fourier transform are considered Fourier transform spectroscopy. By
definition, Fourier transform spectroscopy is a spectroscopic technique where interferograms are collected by measurements of
the coherence of an electromagnetic radiation source in the time-domain or space-domain, and translated into frequency
domain through Fourier transform.

Interferometer-What it is used for and how it works?


How to introduce a time-domain or space-domain variable in the spectrometer is the primary question that needed to be
addressed when we consider constructing a Fourier transform spectrometer. In the experimental set-up, a Michelson
interferometer is commonly used to solve this problem. Different from the classical Michelson interferometer with two fixed
mirrors (Figure 1.a), the interferometer used in Fourier transform spectrometer has a moving mirror at one arm (Figure 1.b).
a. b.

7/4/2021 2 https://chem.libretexts.org/@go/page/1785
Figure 1. Scheme for Michelson interferometer [components: coherent light source; half-silvered beam-splitting mirror; two
highly polished reflective mirrors; detector] (a) Stationary version [two fixed mirrors] (b) Movable version [One movable
mirror and one movable mirror]
As shown in Figure 1.b, when a parallel beam of coherent light hits a half-silvered mirror, it is divided into two beams of equal
intensities by partial reflection and transmission. After being reflected back, the two beams meet at the half-silvered mirror and
recombine to produce an interference pattern, which is later detected by the detector. Manipulating the difference between
these two paths of light is the core of Michelson interferometer. If these two paths differ by a whole number of wavelengths,
the resulting constructive interference will give a strong signal at the detector. If they differ by a whole number and a half of
wavelengths, destructive interference will cancel the intensity of the signal.

Measuring Interferograms
With a Fourier transform spectrometer equipped with an interferometer, we can easily vary the parameter in time domain or
spatial domain by changing the position of the movable mirror.
But how data are collected by a Fourier transform spectrometer? A quick comparison between a conventional spectrometer
and a Fourier transform Spectrometer may help to find the answer.
Conventional spectrometer:
Monochromator is commonly used. It can block off all other wavelengths except for a certain wavelength of interest. Then
measuring the intensity of a monochromic light with that particular wavelength becomes practical. To collect the full spectrum
over a wide wavelength range, monochromator needs to vary the wavelength setting every time.
Fourier transform spectrometer:
Rather than allowing only one wavelength to pass through the sample at a time, an interferometer can let through a beam with
the whole wavelength range at once, and measure the intensity of the total beam at that optical path difference. Then by
changing the position of the moving mirror, a different optical path difference is modified and the detector can measure
another intensity of the total beam as the second data point. If the beam is modified for each new data point by scanning the
moving mirror along the axis of the moving arm, a series of intensity versus each optical path length difference are collected.
So instead of obtaining a scan spectrum directly, raw data recorded by the detector in a Fourier transform spectrometer is less
intuitive to reveal the property of the sample. The raw data is actually the intensity of the interfering wave versus the optical
path difference (also called Interferogram). The spectrum of the sample is actually encoded into this interferogram.

Extracting the spectrum from raw data


Based on the previous discussion, it is predictable that, without further translation, the raw data collected on a Fourier
transform spectrometer will be quite difficult to read. A Fourier transform needs to be performed to decode interferogram and
extract actual spectrum I(v̄ ) from it. The following shows how to conduct a Fourier transform to decode:
¯
¯

The intensity collected by the detector is a function of the path length differences in the interferometer p and wavenumber v [3]: ¯
¯¯

¯¯ ¯¯ ¯¯
I (p, v̄) = I (v̄)[1 + cos(2π v̄p)] (9)

Thus, the total intensity measured at a certain optical path length difference (for each data point at a certain optical pathlength
difference p) is:

¯
¯¯ ¯
¯¯ ¯
¯¯ ¯
¯¯
I (p) = ∫ I (p, v) = I (v)[1 + cos(2π vp)] ⋅ dv (10)
0

7/4/2021 3 https://chem.libretexts.org/@go/page/1785
It shows that they have a cosine Fourier transform relationship. So by computing an inverse Fourier transform, we can resolve
the desired spectrum in terms of the measured raw data I(p) (10):

1
¯¯ ¯¯
I (v̄) =4∫ [I (p) − I (p = 0)] cos(2π v̄p) ⋅ dp (11)
0
2

An example to illustrate the raw data and the resolved spectrum is also shown in Figure 2.

Figure 2. Fourier transform between interferogram and actual spectrum[4]

The Fast Fourier Transform (FFT)


Fast Fourier Transform (FFT) is a very efficient algorithm to compute Fourier transform. It applies to Discrete Fourier
Transform (DFT) and its inverse transform. DFT is a method that decomposes a sequence of signals into a series of
components with different frequency or time intervals. This operation is useful in many fields, but in most cases computing it
directly from definition is too slow to be practical. Fast Fourier Transform algorithm can help to reduce DFT computation time
by several orders of magnitude without losing the accuracy of the result. This benefit becomes more significant when the
number of the components is very large. FFT is considered a huge improvement to make many DFT-based algorithms
practical. In Fourier transform spectrometer, signals are often collected by a series of optical or digital channels at the detector.
Then FFT is of great importance to quickly achieve the following signal processing and data extraction based on DFT method.
Combining all these steps together, we can take a look at how the data from the sample are processed. The diagram is shown in
Figure 3.

Figure 3. Data processing in FTIR


(Figure from ThermoNiolet at mmrc.caltech.edu/FTIR/FTIRintro.pdf)

Different operating modes in Fourier Transform Spectrometer


Continuous/Scanning FTS
Continuous Fourier transform spectroscopy refers to the scanning form of FTS, in which by step moving one mirror, the whole
range of optical path difference is measured. This is the most widely used mode in FTS, like most absorption spectra and
emission spectra obtained by FTS.
Pulsed FTS
In some Fourier transform spectrometers, depending on the feature of the involved spectroscopic technique and purpose of
measurement, a pulsed Fourier transform technique may be applied instead of the scanning mode.

7/4/2021 4 https://chem.libretexts.org/@go/page/1785
Pulsed FTS is different from conventional continuous FTS. It is not based on the transmittance technique, which is widely
used in the absorption spectra, like FTIR. Instead, in pulse FTS, the idea is that the sample is first exposed to an energizing
event, and this pulse induces a periodic response. The frequency of this response relative to the field strength is determined by
the properties of the sample. Using Fourier transform to resolve the frequency will tell the information about the targeted
analyte.
Pulse FTS is a relatively new improvement of FTS. Some examples are from pulse-Fourier Transform-Nuclear Magnetic
Resonnance (FT-NMR), pulse-Fourier Transform-Electron Paramagnetic Resonnance (FFT-EPR) and Fourier Transform-Mass
Spectrometry (FT-MS). Please refer to the following topics for more details about how they work.
Stationary FTS
In addition to the continuous/scanning mode of FTS, a number of stationary Fourier transform spectrometers are also available
to meet special needs.
The principle of the interferometer and the analysis of its output signal is similar to the typical scanning FTS. But the signal is
collected at certain optical path length differences rather than scanning over the whole range of the path difference.

An overview of various FTS techniques


Fourier transform spectroscopy can be applied to a variety of regions of spectroscopy and it continues to grow in application
and utilization including optical spectroscopy, infrared spectroscopy (IR), nuclear magnetic resonance, electron paramagnetic
resonance spectroscopy, mass spectrometry, and magnetic resonance spectroscopic imaging (MRSI). Among them, Fourier
Transform Infrared Spectroscopy (FTIR) has been most intensively developed, which uses scanning Fourier transform to
measure the mid-IR absorption spectra. .
Nuclear Magnetic Resonance (NMR) and Electron Spin Resonance Spectroscopy (EPR) are two magnetic techniques that use
pulse Fourier transform mode. A Radio Frequency Pulse (RF Pulse) in a strong ambient magnetic field background is used as
the energizing event. This RF Pulse directs the magnetic particles at an angle to the ambient strong magnetic field, causing
gyration of the particle. Then the resulting gyrating spin induces a periodic current in the detector coil. This periodic current is
recorded as the signal. Each gyrating spin has a characteristic frequency relative to the strength of the ambient magnetic field,
which is also governed by the properties of the sample.
Fourier Transform Mass Spectrometry (MS) is also operated at pulse Fourier transform mode. Different from NMR and EPR,
the injection of the charged sample into the strong electromagnetic field of a cyclotron acts as the energizing event in MS. The
injected charged particles travel in circles under the strong electromagnetic field. The circular pathway will thus induce a
current in a fixed coil at one point in their circle. Each traveling particle exhibits a characteristic cyclotron frequency relative
to the field strength, which is determined by the masses in the sample.

References
1. R.A. Serway, J.W. Jewett.Principles of physics: a calculus-based text (4th edition), 809
2. Maxwell; James Clerk, A Dynamical Theory of the Electromagnetic Field, Philosophical Transactions of the Royal Society
of London 155, 459-512 (1865).
3. Peter Atkins, Julio De Paula. 2006. Physical Chemistry, 8th ed. Oxford University Press: Oxford, UK.
4. Smith, B.C. Fourier Transform Infrared Spectroscopy. Boca Raton: CRC Press Inc,1996
5. Bernhard Jaun, Analytische Chemie IV-Structure determination by NMR, 1. Practical aspects of pulse Fourier transform
NMR spectroscopy, 1.1-1.21

Outside Links
Please visit the free source for a consice introductory liquid pulse FT-NMR textbook[5] :
www.analytik.ethz.ch/praktika...nmr/ft-nmr.pdf
Fourier Series en.Wikipedia.org/wiki/Fourier_series
Electromagnetic wave equation en.Wikipedia.org/wiki/Electro..._wave_equation
Fourier Transform en.Wikipedia.org/wiki/Fourier_transform
Fourier Transform Spectroscopy en.Wikipedia.org/wiki/Fourier...m_spectroscopy
Fast Fourier Transform en.Wikipedia.org/wiki/Fast_Fourier_transform
Discrete Fourier Transform en.Wikipedia.org/wiki/Discret...rier_transform

7/4/2021 5 https://chem.libretexts.org/@go/page/1785
Problems
1. Search literatures to find the advantages and the limitations of Fourier transform spectroscopic techniques.
2. Perform a Fourier transform to show how to extract spectrum, equation (11), from the raw data in equation (10).
3. Compare interferometer with monochromator (at least two aspects).
4. What are the important components to make Fourier transform spectrometer practical? What are they used for?
5. Based on the information introduced in this module, design any one of the Fourier transform spectrometers mentioned in
the context.

Solutions
1. Advantages
Firstly, Fourier transform spectrometers have a multiplex advantage (Fellgett advantage) over dispersive spectral
detection techniques for signal, but a multiplex disadvantage for noise; Moreover, measurement of a single spectrum is
faster(in the FTIR technique) because the information at all frequencies is collected simultaneously. This allows
multiple samples to be collected and averaged together also resulting in an improvement in sensitivity; In addition, FT
spectrometers are cheaper than conventional spectrometers because building of interferometers is easier than the
fabrication of a monochromator (in the FTIR technique). So most commercial IR spectrometers are built based on
FTIR techniques.
Limitations: practical frequency regions limited (FT UV-vis is not quite practical)
2. See the following figure for the solution:

3. Interferometer vs. Monochromator


Interometer: a. Collect signal in time or spatial domain; b. Measure all frequencies in the incident beam at one time; c.
Determined by the interferometer, raw data from FT spectrometer is an interogram, which needs to be Fourier Transform back
to get spectrum.
Monochromator: a. Collect signal in frequency domain; b. Scan each wavelength and measure the intensity for each single
wavelength at a time; c. Determined by the feasure of monochromator, spectrum can be directly collected from the
spectrometer;
4. Interferometer and Fast Fourier Transform Data Analyzer
Interferometer: to generate continuous optical path length difference and enable the idea to collect data in the time or spatial
domain;
Fast Fourier Transform Data Analyzer: to quickly transform the raw data (interferogram) to spectrum by using fast Fourier
transform algorithm;
5. FTIR instrumentation: Figure 4

7/4/2021 6 https://chem.libretexts.org/@go/page/1785
Figure 4. A FTIR Spectrometer Layout (Figure from ThermoNiolet at mmrc.caltech.edu/FTIR/FTIRintro.pdf)
FT-NMR instrumentation:
A modern high resolution liquid FT-NMR instrumentation is shown in Figure 5:

Figure 5. A schematic diagram of liquid FT-NMR[5]

Contributor
Jing Zhao, Chemistry, University of California, Davis

7/4/2021 7 https://chem.libretexts.org/@go/page/1785
Time-resolved vs. Frequency Resolved
Spectroscopic measurements are typically taken in one of two domains: frequency or time. These measurements are given the
terms frequency-resolved or time-resolved. Frequency-resolved measurements are the most familiar forms of spectroscopy.
UV/Visible, IR, Raman, and X-ray spectroscopy are typically done in the frequency domain. This type of spectroscopy
acquires data across a range of frequencies (or wavelengths). The data acquired is typically in the form of an light intensity
which can in turn be interpreted as absorbance, transmittance, reflectance, or photon scattering depending on the instrument
and technique being used.

Introduction
The less familiar time-resolved spectroscopy includes Ultrafast laser spectroscopy and florescence. In this form of
spectroscopy data is acquired over a range of time. This data can be at a single wavelength or at multiple wavelengths,
depending on the specific technique. Some spectroscopic techniques, such as Ultrafast laser spectroscopy, FT-NMR and FT-IR,
span both frequency and time domains. In the case of Ultrafast laser spectroscopy, useful data is acquired in both the time and
frequency domains. FT-NMR and FT-IR acquire data in the time domain. That data is then converted into a signal in the
frequency domain using a process called Fourier Transform. This is covered in more detail below.

Frequency Resolution
Frequency is defined as inverse time. The unit is typically given in inverse seconds(s-1) or hertz(hz). Frequency is used to
represent the number of cycles occurring in a given time period. These cycles could be any repetitive process including: the
periodic motion of a harmonic oscillator, the sinusoidal propagation of electromagnetic radiation, or the rotation of a rigid
rotator. The most relevant for spectroscopy is the propagation or electromagnetic radiation, or light. This is often represented
in many different forms that, though not technically frequency, are related to frequency, and therefore fall into the frequency
domain. These include representing frequency as wavelength (nm), wavenumber (cm-1), and photon energy (eV). These are all
connected by a few simple equations given below. (c=speed of light, E=energy, v=frequency, w=wavenumber, h=Planck's
constant,lambda=wavelength)
c
λ =
ν

E = hν

1
w =
λ

The frequency domain is the most familiar domain in spectroscopy. UV/visible, infrared, photoelectron, microwave, and X-ray
spectroscopy all have applications in the frequency domain. The results of these spectroscopic techniques are typically given
in some form of intensity versus wavelength. The most familiar is likely the steady state ultraviolet/visible absorption
spectrum(an example is shown below).
Frequency Resolved Visible Absorption Spectrum

The energy of absorbed light corresponds to the energy of transition between two eigen states of the system. In the case of
visible spectroscopy these states are electronic states.

4/24/2021 1 https://chem.libretexts.org/@go/page/1786
Time Resolution
Spectra can also be acquired in the time domain. Rather than acquiring spectra by averaging data over a relatively long time
range, data is acquired over discrete time intervals or, in some cases, continuously. This may be done over one or many
wavelengths. Time resolved spectroscopy observes the change in eigenstates with respect to time. In order for data from time-
resolved spectroscopy to be useful, the spectroscopy must be suited to the time scale of the process of interest. Below is a table
of the approximate time scales and spectral ranges of physical processes that may be interesting.

Process Time Applicable Spectral Range

Singlet Electronic Excited State Lifetime femto-nanoseconds Visible

Triplet Electronic Excited State Lifetime nanoseconds-minutes Visible

Molecular Vibration Excited State Lifetime pico-milliseconds Infrared

Nuclear Rotation pico-microseconds Radio

Molecular Reaction Kinetics eons-picoseconds Varies

If a spectroscopy with suitable spectral region and time resolution is available, time resolved spectroscopy can be used to study
kinetics, reactions, and lifetimes. Common applications of time-resolved spectroscopy include ultrafast laser spectroscopy and
time-resolved florescence.

Fourier Transform
The process of Fourier Transform is a mathematical process used to move from one set of coordinates to another. The most
spectroscopically relevant fourier transform is from the time domain to the frequency domain. In this case, a signal originally
measured in the time domain can be converted into a signal in the frequency domain. This is done via the mathematical
process shown below.

−2iπvt
F (v) = ∫ f (t)e dt
−∞


−2iπvt
f (t) = ∫ F (v)e dv
−∞

A mathematical relation known as Euler's Formula is an important identity when using Fourier Transform, particularly with
sine and cosine functions. This is shown below.
±ix
e = cos(x) ± isin(x)

A simple graphic representation of Fourier Transform is shown below.


Original Signal in Time Domain

Fourier Transformed: Signal in Frequency Domain

4/24/2021 2 https://chem.libretexts.org/@go/page/1786
Fourier Transformed Again: Signal in Time Domain (The Components of the Original Signal)

Sum of the Components(Equivalent to the Original Signal)

An example fourier transform with sine is given in the links section.

Common Application of Fourier Transform


Some fields of spectroscopy use measurements taken in the time domain to gain information about the frequency domain.
These spectroscopies include Nuclear Magnetic Resonance, Fourier Transform Ion Cyclotron Resonance Mass
Spectroscopy(FT-ICR MS) and Fourier Transform Infrared spectroscopy(FT-IR). Unlike the above graphic representation of
Fourier Transform, these systems yield transforms that are impossible to compute by hand. Computer algorithms must be used.
The initial signal typically forms a beat pattern. This can be in the form of an interferogram or a free induction decay(FID), in
the cases of FT-IR and NMR, respectively. Graphic representations of an interferogram and FID are shown below alongside IR
and NMR spectra.
FT-IR Spectra

4/24/2021 3 https://chem.libretexts.org/@go/page/1786
Interferogram(X-axis: Time|Y-Axis:Amplitude) Infrared Spectrum(X-axis: Wave Number (cm-1)|Y-axis: Percent Transmission)
FT-NMR Spectra

Free Induction Decay(X-axis: Time|Y-axis: Amplitude) NMR Spectrum(X-axis: Chemical Shift (ppm)|Y-axis: Amplitude)

Spectroscopy in Both Time and Frequency Domains


Some spectroscopies yield data in both time and frequency domains. The most prominent of these techniques is time-resolved
laser spectroscopy. By measuring complete spectra at discrete time intervals, spectral evolution with respect to time can be
monitored. This technique is unique in it's ability to collect data in multiple domains with femtosecond time resolution. This
allows electronic states to be monitored both as they evolve over time and statically at any particular time along the time scale
of the instrument. Below is an example of a typical ultrafast spectrum, notice both wavelength and time domains. This three-
dimensional spectrum can be deconstructed to yield either time or frequency dependent results in two-dimensions. This is also
shown below.
Time Resolved Visible Absorbance Spectrum

(X-axis:Wavelength(nm)|Y-axis:Time(ps)|Z-axis:(ABS)[Shown as color])
The time resolved spectrum shown above is plotted as a contour plot showing the Z-axis as a color gradient from red(low
signal) to blue(high signal). This contour plot contains a multitude of information, but is by itself not terribly useful. It is most
easily analyzed by taking crossections at a single wavelength or time. Each is shown below.
Frequency Domain Signal (A crossection of the Time Resolved Spectrum constant time)

4/24/2021 4 https://chem.libretexts.org/@go/page/1786
The above spectrum is a crossection of the complete time resolved spectrum. This particular crossection is the frequency
domain signal at 56.75 picoseconds after the sample has been excited by a laser pulse.
Time Domain Signal (A crossection of the Time Resolved Spectrum at a single Wavelength)

The above spectrum is a crossection of the complete time resolved spectrum. This crossection is the time resolved signal at
618 nanometers. This crossection contains kinetic data on the eigen state that absorbs at 618nm.

References
1. Skoog, Douglas A. Principles of Instrumental Analysis. 5th ed. New York:Thompson Learning Incorporated,1998.
2. Atkins, Peter. Physical Chemistry. 7th ed. New York: W.H. Freeman, 2002.

Outside Links
For a more in-depth/mathematical look at fourier transform: en.Wikipedia.org/wiki/Fourier_transform
Fourier transform of a sine function: http://mathworld.wolfram.com/FourierTransformSine.html
For a more in depth look at Euler's Formula: http://en.Wikipedia.org/wiki/Euler%27s_formula
Use of Fourier Transform in FT-ICR Mass Spectroscopy: www.youtube.com/watch?v=7EHngA4S3Ws

4/24/2021 5 https://chem.libretexts.org/@go/page/1786
SECTION OVERVIEW
MAGNETIC RESONANCE SPECTROSCOPIES
The interaction of an externally applied magnetic field and spins (either nuclear or electron).

ELECTRON PARAMAGNETIC RESONANCE


Electron Paramagnetic Resonance (EPR) is a remarkably useful form of spectroscopy used to study
molecules or atoms with an unpaired electron. It is less widely used than NMR because stable
molecules often do not have unpaired electrons (i.e., paramagnetic). However, EPR can be used
analytically to observe labeled species in situ either biologically or in chemical reactions.

ENDOR: THEORY
EPR: APPLICATION
EPR: INTRODUCTION
EPR: PARALLEL MODE OPERATION
EPR: THEORY
EPR - INTERPRETATION
HYPERFINE SPLITTING

NUCLEAR MAGNETIC RESONANCE


Nuclear Magnetic Resonance (NMR) Spectroscopy uses the electromagnetic radiation of radio waves to probe the local electronic
interactions of a nucleus. NMR is a non-destructive technique and has found uses in fields of medicine, chemistry, and environmental
science.

NMR: BACKGROUND PHYSICS AND MATHEMATICS


NMR: EXPERIMENTAL

2D NMR
2D NMR: INDIRECT DETECTION
2D NMR BACKGROUND
2D NMR BASICS
2D NMR EXPERIMENTS
2D NMR INTRODUCTION
HETERONUCLEAR CORRELATIONS
HOMONUCLEAR CORRELATIONS
CARBON-13 NMR

INTERPRETING C-13 NMR SPECTRA


DIFFUSION ORDERED SPECTROSCOPY (DOSY)
MAGNETIC RESONANCE IMAGING
NMR - INTERPRETATION
INTEGRATION IN NMR
PASCAL’S TRIANGLE
NMR HARDWARE
NMR SPECTROSCOPY IN LAB: COMPLICATIONS
PULSE SEQUENCES
PHASE CYCLING AND COHERENCE TRANSFER PATHWAYS
SOLID STATE EXPERIMENTS
CAR-PURCELL-MEIBOOM-GILL (CPMG) ECHO TRAIN ACQUISITION
MAGIC ANGLE HOPPING (MAH)
SOLID STATE NMR EXPERIMENTAL SETUP
NMR: STRUCTURAL ASSIGNMENT
(N+1) RULE
BACKGROUND TO C-13 NMR
DETERMINE STRUCTURE WITH COMBINED SPECTRA

1 7/7/2021
HIGH RESOLUTION PROTON NMR SPECTRA
INTEGRATION IN PROTON NMR
INTERPRETING C-13 NMR SPECTRA
INTRODUCTION TO PROTON NMR
LOW RESOLUTION PROTON NMR SPECTRA
MORE ABOUT ELECTRONICS
MULTIPLICITY IN PROTON NMR
NMR11. MORE ABOUT MULTIPLICITY
NMR14. MORE PRACTICE WITH NMR SPECTROSCOPY
NMR2. CARBON-13 NMR
NMR3. SYMMETRY IN NMR
NMR4. 13C NMR AND GEOMETRY
NMR5. 13C NMR AND ELECTRONICS
NMR8. CHEMICAL SHIFT IN 1H NMR
NMR APPENDIX. USEFUL CHARTS FOR NMR IDENTIFICATION
NMR - THEORY
BLOCH EQUATIONS
LARMOR PRECESSION
NMR INTERACTIONS
CHEMICAL SHIFT (SHIELDING)
DIPOLAR COUPLING
J-COUPLING (SCALAR)
PASCAL’S TRIANGLE CONSTRUCTION
QUADRUPOLAR COUPLING
QUANTUM MECHANIC TREATMENT
RELAXATION
NMR: KINETICS
NUCLEAR OVERHAUSER EFFECT
SOLOMON EQUATIONS
SPIN-SPIN RELAXATION
SPIN LATTICE RELAXATION
ROTATIONS AND IRREDUCIBLE TENSOR OPERATORS
NMR: INTRODUCTION
BACK MATTER
INDEX

2 7/7/2021
CHAPTER OVERVIEW
ELECTRON PARAMAGNETIC RESONANCE
Electron Paramagnetic Resonance (EPR) is a remarkably useful form of spectroscopy used to study
molecules or atoms with an unpaired electron. It is less widely used than NMR because stable
molecules often do not have unpaired electrons (i.e., paramagnetic). However, EPR can be used
analytically to observe labeled species in situ either biologically or in chemical reactions.

ENDOR: THEORY
Electron-nuclear double resonance spectroscopy (ENDOR) is a powerful advanced EPR technique
that probes the environment surrounding paramagnetic centers. It is of great use to further examine
paramagnetic samples which give complicated spectra via the standard EPR method due to
electronic-nuclear interactions manifested as the hyperfine interaction.

EPR: APPLICATION
EPR: INTRODUCTION
Though less used than Nuclear Magnetic Resonance (NMR), Electron Paramagnetic Resonance (EPR) is a remarkably useful form of
spectroscopy used to study molecules or atoms with an unpaired electron. It is less widely used than NMR because stable molecules
often do not have unpaired electrons. However, EPR can be used analytically to observe labeled species in situ either biologically or
in a chemical reaction.

EPR: PARALLEL MODE OPERATION


This module presents the theory that describes how EPR transitions can be induced in integer high spin systems by the application of
a modulating magnetic field parallel to the bond axis (z-axis), as well as some of the applications of this technique to various
molecular systems.

EPR: THEORY
Electron Paramagnetic Resonance (EPR), also called Electron Spin Resonance (ESR), is a branch of magnetic resonance spectroscopy
which utilizes microwave radiation to probe species with unpaired electrons, such as radicals, radical cations, and triplets in the
presence of an externally applied static magnetic field.

EPR - INTERPRETATION
Electron paramagnetic resonance spectroscopy (EPR), also called electron spin resonance (ESR), is a technique used to study
chemical species with unpaired electrons. EPR spectroscopy plays an important role in the understanding of organic and inorganic
radicals, transition metal complexes, and some biomolecules.

HYPERFINE SPLITTING
This splitting occurs due to hyperfine coupling (the EPR analogy to NMR’s J coupling) and further splits the fine structure (occurring
from spin-orbit interaction and relativistic effects) of the spectra of atoms with unpaired electrons. Although hyperfine splitting
applies to multiple spectroscopy techniques such as NMR, this splitting is essential and most relevant in the utilization of EPR.

1 7/7/2021
ENDOR: Theory
Electron-nuclear double resonance spectroscopy (ENDOR) is a powerful advanced EPR technique that probes the environment surrounding paramagnetic centers. It is of great use to further
examine paramagnetic samples which give complicated spectra via the standard EPR method due to electronic-nuclear interactions manifested as the hyperfine interaction.

Introduction
The primary interest of this advanced EPR method lies in elucidation of the hyperfine interactions or coupling (hfc) of an unpaired electron with surrounding nuclei. In a standard CW-EPR
experiment, one can often detect these hyperfine splittings, however accurate calculation of their values is generally a challenge. This is due to complexity of the environment of the spin
system, which can result in splitting of the EPR transition due to coupling of the unpaired electron with several neighboring nuclei, resulting in overlapping signals that cannot always be
distinguished. Inhomogeneous broadening of the EPR lineshapes due to a distribution in some property of the system, such as the orientation of the spins in a powder or frozen solution, also
broadens each peak, thereby severely limiting assignment of hfc parameters and identification of these nuclei. ENDOR is able to greatly simplify these CW-EPR spectra by reducing the
number of peaks, displaying only 2N peaks as opposed to 2N peaks for N-sets of equivalent nuclei, and allowing for selectivity of which couplings are observed through optimization of the
ENDOR parameters. For a comprehensive discussion of EPR theory and relevant information to comprehensive of the content of this module please refer to the EPR: Theory module.

The Hyperfine Interaction


The permanent magnetic dipole of nuclei such as hydrogen, with I = ½, can interact with that of an unpaired electron if they are located closely in space. This hyperfine interaction is
independent of the strength of the static magnetic field that is applied in a magnetic resonance experiment and leads to a splitting of the electronic spin states by an energy difference of half of
the hyperfine coupling, A . The energy splitting can be described by the equation:

Ehf c = A mI  mS (ENDOR.1)

where A is the hyperfine coupling constant (hfc). This hfc constant can contain both isotropic and anisotropic contributions, the latter of which is discussed in the dipolar coupling section of
the module, but depends on both the electronic and nuclear g values, the electronic and nuclear magnetic moments, and the distance between them. These magnetic moments with the B0 field
along the z-axis are:
^ ^
μez = γe Sz ℏ = −gβe Sz (ENDOR.2)

^ ^
μnz = γn Iz ℏ = +gn βn Iz (ENDOR.3)

for the electron and nucleus respectively. The isotropic hyperfine constant, that in the absence of a dipolar coupling contribution, as well as the Hamiltonain describing isotropic hfc are given
by the equations below.
2μ0 2
Aiso = gβe βn |Ψ(0)| (ENDOR.4)
3

2μ0 2
^ ^ ^
H = gβe βn |Ψ(0)| Sz Iz (ENDOR.5)
3

The hfc constant is generally expressed in terms of units of frequency as:


Aiso
(ENDOR.6)
h

and can also be expressed in magnetic field units with the equation:
Aiso
aiso = (ENDOR.7)
ge βe

It is generally this parameter that is sought to be determined in via ENDOR experiments.

Fundamentals of the technique


The key to the ENDOR technique lies in the use of both microwave, MW, and radiofrequency, RF, (B1) applied fields, resulting in a hybrid EPR and NMR experiment. Like all EPR
experiments, the sample is placed in a homogenous magnetic field, B0, and secondary fields are applied to induce transitions between electronic spin states, however ENDOR also induces
nuclear spin transitions to extract hfc parameters. The important requirement of the technique is the saturation of a specific EPR transition, resulting in an almost equal population in each of the
two spin states, for example ms = ± ½. Then an RF frequency signal is applied to induce and also partly saturate NMR transitions, which results in a desaturation of the corresponding EPR
transition. The selection rules for ENDOR are the same as that for NMR, ΔM = 0,  ΔM = ±1 . The saturation requirement for ENDOR can be shown below by the following two
S I

equations:
2 2 2 2
γe B1 T1e T2e ≥ 1,  γn B1 T1n T2n ≥ 1 (ENDOR.8)

which show that the product of the gyromagnetic ratios, the saturating MW and RF field intensities, and the characteristic relaxation times of the system must be unity or larger.

From here we will consider the simple case of the interaction between one electron, S = ½, and one proton, I = ½, for which the energy levels are shown below. The first splitting arises from
the electron zeeman splitting discussed below, which removes the degeneracy of the electron ms = ± ½ spin states. The next splitting arises from the hyperfine coupling of the electron and
nuclear spins, split by half of the hfc magnitude. The final splittings show the ENDOR/NMR transitions which comprise the spectrum. The wide arrows indicate the EPR transitions which are
saturated.

Population Polarization
The diagram below shows the changes in populations that occur during an ENDOR experiment, as it is these changes that lead to the ENDOR effect and that are detected. Prior to the
application of MW frequency photons, or a preparation pulse, the system exists with the majority of the population in the ms = - ½ spin manifold, as determined by the Boltzmann population
distribution and show in (a) of the figure below. With the application of a MW field, or preparation MW pulse, the population is moved from the ms = - ½ to ms = + ½ manifold as expected by
the inversion sequence, seen in (b) of the figure. Finally, after the mixing period and induction of on-resonance NMR transitions, population transfer occurs between the upper mI levels as seen
in (c.) of the figure below. This is then detected as a change in the EPR signal with the CW technique, or with an ESE, electron spin echo, sequence using a pulsed ENDOR strategy. If the RF π
pulse is off resonance, then this final population polarization does not occur.

7/7/2021 ENDOR.1 https://chem.libretexts.org/@go/page/1790


Continuous Wave (CW) ENDOR
The first ENDOR technique invented was continuous wave ENDOR, in which the radio frequency source is modulated while the B0 applied field is held constant unlike a standard EPR
experiment. The first derivative spectra are thus still obtained via this techniques. The EPR transitions seen above are saturatured, and an applied RF field desaturates these transitions, which
are observed in the CW ENDOR spectrum as changes in the EPR signal intensity as a function of the frequency of the RF field applied. This is the most simple ENDOR experiment to discuss,
however it has become a less commonly used method for acquisition of ENDOR spectra in recent years.

Pulsed ENDOR
Pulsed techniques offering several advantages over continuous wave ENDOR. CW ENDOR is the most sensitive of the techniques, however pulsed ENDOR provides higher resolution as well
as allowing observation of weakly coupled nuclei. It also offers insight into the relaxation times of the system, is less susceptible to artifacts, and does not require balancing of relaxation times
with RF power like the CW technique. Therefore, pulsed-ENDOR has been the most common technique in the past several decades, using pulse sequences predominantly developed by Mims
and Davies. For an ENDOR experiment, either Mims or Davies, the homogenous applied magnetic field B0 is held constant, as opposed to the normal scanning of this field in a standard CW-
EPR experiment. The field is selected where the EPR transition occurs by performing an ESE field sweep prior to beginning the experiment, thus fixing both the B0 and MW-B1 fields. This
leaves the RF spectrum to be swept during the mixing period, much like in a standard NMR experiment. This is done by repeating the pulse sequence millions of times and using a different RF
frequency during the mixing period for each sequence, which is done by randomly sampling each of the RF frequencies in the range. Electron spin echo, or ESE, is a technique used for signal
detection in most pulsed advanced EPR methods

Davies ENDOR
In Davies ENDOR, a preparation π, 180o, pulse is used in order to invert the magnetization of the spins in the applied static B0 field. This essentially creates a hole in the EPR spectrum, whose
width and depth depend on the length of the pulse applied, with long pulse producing narrow holes. During the mixing period a π RF pulse is applied and if the RF frequency is resonant with
an NMR transition, magnetization will be transferred to the other ms spin manifold, otherwise no mixing will occur to fill in the hole that the inversion pulse creates. During the detection
period, the z-component of the magnetization is measured using a two pulse echo sequence, and one detects essentially the EPR signal that is restored during the mixing period. This technique
is most suited to nuclei with large hfc values.

Mims ENDOR
The Mims ENDOR technique is based on a stimulated electron spin echo (ESE) sequence, using a two , 90o, preparation pulse sequence to invert the electron spin population, and a final
π

2
π

pulse after the mixing period to stimulate the ESE for signal detection. Between the preparation pulses and the final pulse, a radio frequency ppulse is used to invert the nuclear spin population,
resulting in polarization transfer between the nuclear and electronic transitions in the so called “mixing period.”

This results in the actual ENDOR transitions, changes in the intensity of the EPR transition, that are detected by use of the ESE MW-pulse sequence. The echo intensity is subsequently
measured as a function of the RF frequency to give the characteristic ENDOR spectrum shown below.

The hfc can be determined experimentally:


−1
| νn 1 ∓ νn 2 | = h |A|

where the plus sign is for the low field limit and the minus sign for the high field limit. In the low field limit, the nuclear Larmor frequency, or frequency of the precession of the nuclear spin in
the presence of a magnetic field, is less than half of the hyperfine coupling, A. In the high field limit, the nuclear Larmor frequency is greater than half of the hyperfine coupling term, resulting
in different assignments of the hfc parameter A according to the above equation.
The two techniques are very similar to each other in their implementation, however they are relatively complementary in their results and usefulness. The Mims ENDOR technique is most
suited to weakly coupled nuclei, with small hyperfine coupling constants, A, generally less than 2 MHz and as low as 0.1 MHz. This makes the Davies ENDOR pulse sequence most useful for
those nuclei with relatively large hyperfine couplings, therefore making the assignments of strongly and weakly coupled nuclei greatly simplified, as well as the corresponding spectra
associated with each of the pulse sequences.

Experimental Parameters
The parameters to be specified in any given ENDOR experiment include the static B0 field, the microwave frequency, the RF power and range, the pulse lengths, the delay time (τ ), and the
repetition rate. The delay time, or time between data collection at each RF frequency, is determined by the timescale on which equilibrium magnetization is restored. This is determined by a
couple relaxation parameters, most importantly the spin lattice relaxation time (T1) and the transverse relaxation time (T2). T1 determines how quickly the magnitization realigns with the z-axis
(B0) field, and T2 determines how quickly the magnetization in the x-y plane (transverse) disappears.
In a Davies ENDOR experiment, the microwave frequency bandwidth is the most important parameter to be set. The preparation inversion pulse burns a hole in the EPR spectrum, i.e. saturates
an EPR transition, which is then filled in through population transfer by the RF inversion pulse. It is this parameter that needs to be optimized in order to get the best spectral intensity and
resolution. In Mims ENDOR, the τ value is the most important parameter to set, and the microwave frequency bandwidth is less important, which is the opposite case for Davies. However, a
hole is still burned in the EPR spectrum, now with the use of two pulses in the preparation period, but it is modulated according to the following equation.
π

7/7/2021 ENDOR.2 https://chem.libretexts.org/@go/page/1790


(B0 − B)
1 + cos 2π τ (ENDOR.9)

Application of the RF inversion pulse during the mixing period shifts the magnetization away from B0 and changes this magnetization pattern. Mims ENDOR intensity then depends on the
differences in the magnetization pattern following the preparation pulse, and those following the mixing period. Therefore Mims intensity is larger for larger differences in sinusoidal
magnetization.

Spin Hamiltonian
Although the spin Hamiltonian applies to all EPR techniques, it is imperative to highlight the term of most importance for ENDOR spectroscopy which is the hyperfine coupling term
describing the interaction of an unpaired electron with nuclei in its vicinity. The Hamiltonian for a system consisting of two spins, an electron with S = ½ and proton with I = ½ is shown below.
^ =H
H ^ ^ ^ ^
EZ + H N Z + H H F + H Q (ENDOR.10)

T T
^ =μ B T ^ T^ ^ ^ ^ ^
H B g S − μn gn B I +S AI + I QI (ENDOR.11)

The terms shown are the electron Zeeman term, nuclear Zeeman term, hyperfine coupling term, and the nuclear quadrapole interaction term (which does not apply to nuclei with I < 1). Both
the electron and nuclear Zeeman terms arise from the splitting of the spin states which are degenerate in the absence of an external magnetic field. The splitting of the spin up/down energy
levels are linearly dependent on the strength of the applied B0 field according to these terms.

The third term describing the hyperfine interaction of the electron spin with nuclear magnetic moment is the most important term of the ENDOR technique. The hyperfine coupling magnitude
and the identification of the nuclei from which the hfc arises is generally what ENDOR is used to elucidate, especially for the cases mentioned in the intro such as couplings to several nuclei or
inhomogeneous broadening of a CW-EPR spectrum. The hyperfine coupling term A has two components arising from both an isotropic term, a , and in the case of a dipolar coupling
iso

interaction, an anisotropic term, a . The final term is the nuclear quadrapole interaction that arises for nuclei with I > 1, and this term is inherently anisotropic.
dip

\hat{\mathcal{H}} = \mu_B \bf{B}}^T{\bf{g} \hat{\bf{S}} - \mu_n g_n \bf{B}}^T\hat{\bf{I}+ \hat{\bf{S}}^T \bf{A}}\hat{\bf{I}}+\hat{\bf{I}}^T \bf{Q} \hat{\bf{I}}+\hat{\bf{S}}^T \bf{D}\hat{\bf{S}}+\hat{\b

The full spin Hamiltonian is shown above which has terms that do not arise from the one electron, S = ½, and one proton, I = ½ being described, which includes a zero field splitting term as
well as an electron-electron interaction term, both of which arise from the presence of more than one electron spin.

Anisotropy and Dipolar Coupling


The nuclear quadrupole interaction is intrinsically an anisotropic parameter of a system with I > ½, but the most important parameters of interest to the ENDOR technique that may contain
anisotropy are hyperfine coupling, and less so the electron g value. The electronic g value anisotropy arises from the effect of orbital angular momentum on electron spin energy levels, which
can provide information on the orbitals containing unpaired electrons. Even more important to ENDOR spectroscopy is hyperfine anisotropy, which is manifested in the dipolar coupling
component of the hyperfine interaction. Hfc has two components, the isotropic contribution Aiso, and the anisotropic dipolar coupling contribution, denoted by Adip or T. Aiso is solely a result
of unpaired electron spin in s-type orbitals, as they are the only type with an electron density at the nucleus that is finite.
2 2
Aiso = μ0 μe μn | Ψ(0) | (ENDOR.12)
3

or
2μ0
Aiso = (ge βe gn βn )⟨ρx ⟩ (ENDOR.13)
3

This unpaired spin density at the nucleus can arise from contact with sp-type ligand orbitals as well as by the unpaired spin density in other orbitals causing polarization of core s electrons, and
is classified as direct or local contact. The anisotropic component, arising from dipolar coupling, can be local or nonlocal, and is only seen in frozen solutions or powders by EPR. Indirect
dipolar coupling is a through space interaction between an electron point-dipole and nuclear-point dipole, such as that between a paramagnetic metal and a hydrogen nucleus bound to the metal
through an oxygen atom. The value of the dipolar coupling can be determined by the equation
2
−μ0 3 cos θ−1
Adip = (ge βe gn βn ) (ENDOR.14)
3
4π R

Where R is the distance from the electron point-dipole to nuclear point-dipole and theta is the angle between the bond connecting the two particles. Direct dipolar coupling would arise from the
metal being directly bound to the hydrogen atom, not through a donor atom. These anisotropic effects of the hyperfine coupling can make the ENDOR spectrum distorted from the above
sketches showing two symmetric peaks separated by a clean baseline.

The above X-band spectra simulated using the EasySpin toolbox for MATLAB, developed by Dr. Stefan Stoll, demonstrates this fact. An anisotropic hyperfine coupling tensor was used to
simulate the above left Mims ENDOR spectrum of a single proton and single electron, and an isotropic g tensor was used for each to highlight the effect of an anisotropic hyperfine tensor. The
spectrum no longer contains two symmetric peaks like the idealized spectra shown in the Mims ENDOR spectrum on the right, that was simulated with an isotropic hyperfine term, due to the

7/7/2021 ENDOR.3 https://chem.libretexts.org/@go/page/1790


anisotropy of the hyperfine term from the dipolar coupling contribution discussed above. This anisotropy can complicate the interpretation of ENDOR spectra, however it can also provide a
plethora of information on the symmetry and environment of a paramagnetic center.

References
1. Weil, J. A. & Bolton, J. R. (2007). Electron Paramagnetic Resonance Spectroscopy: Elementary Theory and Applications, Second Edition. Wiley-Interscience.
2. Schweiger, A. & Jeschke, G. (2001). Principles of Pulse Electron Paramagnetic Resonance. Oxford University Press.
3. Hoff, A. J. (1989). Advanced EPR: Applications in Biology and Biochemistry. Elsevier Science Publication Company Inc.
4. Mims, W. B. (1965). "Pulsed ENDOR Experiments". Proceedings of the Royal Society of London. Series A, Mathematical and Physical Sciences. Vol. 283, No. 1395, pp. 452-457

Problems
1. For a system with a single unpaired electron interacting with 4 identical nuclei all with I = 1/2 how many lines would be observed in the CW-EPR spectrum and the ENDOR spectrum?
2. What effect does an increase in the static B0 magnetic field have in a one electron one nucleus system (S = 1/2, I = 1/2) on the separation of the two peaks and the nuclear Larmor
frequency?
3. Why is it that an ENDOR experiment is so much more sensitive than an ordinary NMR experiment?
4. For the Hydrogen atom, with S = 1/2 and I = 1/2, which electronic orbitals will give rise to a hyperfine interaction term in the spin Hamiltonian?

Solutions
1. The number of CW-EPR peaks for the one electron (S = 1/2) and 4 nuclei (I = 1/2) system is 24 = 16 peaks observed in this spectrum. ENDOR reduces the number of peaks in half, hence
2*4 = 8 peaks are seen in this spectrum.
2. An increase in the static applied magnetic field has no effect on the separation of the 2 ENDOR peaks, as the value of the hyperfine coupling remains constant regardless of the strength of
the applied field. However, since the nuclear Larmor frequency (as well as electron Larmor frequency) is linearly dependent on the strength of the applied field, the spectrum would be
shifted to a higher frequency as the strength of the field is increased as g β B.
n n

3. This is mainly due to differences in population distribution in both types of experiments. ENDOR is more sensitive due to larger population distributions between spin up and down becuase
the transitions are inherently larger energy, the rate of absoprtion of radiation is higher in EPR/ENDOR, and in an ENDOR experiment the electron magnetic field adds with the external
field leading to a large net field at the nucleus and thus a larger population difference between spin states.
4. For all atoms, in order for isotropic hyperfine splitting to occur there must be a non-zero probability that the electron may reside at the nucleus. It is apparent that only s-type orbitals give
rise to a non-zero probability at the nucleus, as p, d, f, ..., orbitals all contain a node at the nucleus (for | Ψ(0) | ). Therefore, the 1s1 ground state electronic configuration of Hydrogen will
2

give rise to hyperfine splitting, as well as all excited states with the electron in an s orbital, e.g. 2s1, 3s1, 4s1, etc..

Contributors and Attributions


Weston Michl

7/7/2021 ENDOR.4 https://chem.libretexts.org/@go/page/1790


EPR: Application
This article is about the experimental application of Electron Paramagnetic Resonance spectroscopy (EPR). For theoretical
background on EPR, see alternate article EPR: Theory. The most basic application of Electron Paramagnetic Resonance
spectroscopy requires the use of a microwave radiation source, an electromagnet, a resonator, and a detector.

Contributors and Attributions


Alan Wilder

6/15/2021 EPR.1 https://chem.libretexts.org/@go/page/1791


EPR: Introduction
Though less used than Nuclear Magnetic Resonance (NMR), Electron Paramagnetic Resonance (EPR) is a remarkably useful
form of spectroscopy used to study molecules or atoms with an unpaired electron. It is less widely used than NMR because
stable molecules often do not have unpaired electrons. However, EPR can be used analytically to observe labeled species in
situ either biologically or in a chemical reaction.

Introduction
Electron Paramagnetic Resonance (EPR), also known as Electron Spin Resonance (ESR). The sample is held in a very
strong magnetic field, while electromagnetic (EM) radiation is applied monochromatically (Figure 1).

Figure 1(3)-monochromatic electromagnetic beam


This portion of EPR is analogous to simple spectroscopy, where absorbance by the sample of a single or range of wavelengths
of EM radiation is monitored by the end user ie absorbance. The unpaired electrons can either occupy +1/2 or -1/2 ms value
(Figure 2). From here either the magnetic field "B0" is varied or the incident light is varied. Today most researchers adjust the
EM radiation in the microwave region, the theory is the find the exact point where the electrons can jump from the less
energetic ms=-1/2 to ms=+1/2. More electrons occupy the lower ms value (see Boltzmann Distribution).

Figure 2: Resonance of a free electron.


Overall, there is an absorption of energy. This absorbance value, when paired with the associated wavelength can be used in
the equation to generate a graph of showing how absorption relates to frequency or magnetic field.
ΔE = hν = ge βB B0 (EPR.1)

-24 -1
where ge equals to 2.0023193 for a free electron; βB is the Bohr magneton and is equal to 9.2740 * 10 J T ; and B0
indicates the external magnetic field.

Theory
Like NMR, EPR can be used to observe the geometry of a molecule through its magnetic moment and the difference in
electron and nucleus mass. EPR has mainly been used for the detection and study of free radical species, either in testing or
anylytical experimentation. "Spin labeling" species of chemicals can be a powerfull technique for both quantification and
investigation of otherwise invisible factors.

7/4/2021 EPR.1 https://chem.libretexts.org/@go/page/1793


The EPR spectrum of a free electron, there will be only one line (one peak) observed. But for the EPR spetrum of hydrogen,
there will be two lines (2 peaks) observed due to the fact that there is interaction between the nucleus and the unpaired
electron. This is also called hyperfine splitting. The distance between two lines (two peaks) are called hyperfine splitting
constant (A).
By using (2NI+1), we can calculate the components or number of hyperfine lines of a multiplet of a EPR transtion, where N
indicates number of spin, I indicates number of equivalent nuclei. For example, for nitroxide radicals, the nuclear spin of 14N
is 1, N=1, I=1, we have 2 x 1 + 1 = 3, which means that for a spin 1 nucleus splits the EPR transition into a triplet.
To absorb microwave, there must be unpaired electrons in the system. no EPR signal will be observed if the system contains
only paired electrons since there will be no resonant absorption of microwave energy. Molecules such as NO, NO2, O2 do have
unpaired electrons in groud states. EPR can be also performed on proteins with paramagnetic ions such as Mn2+, Fe3+ and
Cu2+. Additionally, molecules containing stable nitroxide radicals such as 2,2,6,6-tetramethyl-1-piperidinyloxyl (TEMPO,
Figure 3) and di-tert-butyl nitroxide radical.

Figure 3-The nitroxide radical TEMPO


Examples of EPR spectra:

Figure 4 -Stimulated EPR spectrum of CH3 radical

Figure 5 - Stimulated EPR spectrum of methoxymethl ( H2C(OCH3) )radical

References
1. G. Maksina, Yu. S. Arkhangel'skaya and B. A. Dainyak. Russian State Medical University, Moscow. Translated from
Meditsinskaya Tekhnika, No. 5, pp. 32–34, September–October, 1995. external link:

www.springerlink.com/content/n73312406t6941l4/fulltext.pdf
2. E.F. Block. The Role in Coherent Resonance in Human Fair: Part One-Electromagnetic and Gravity. December 2010.
http://journalinformationalmedicine.org/cr1.htm
3. Thomas Engel, Gary Drobny and Philip Reid. Physical Chemistry for the Life Science.s 2008 Pearson Education, Inc.
Upper Saddle River, NJ 07458. pp. 514-516
4. Cortel, Adolf. "Demonstrations on Paramagnetism with an Electronic Balance." J. Chem. Educ. 1998 75 61.

7/4/2021 EPR.2 https://chem.libretexts.org/@go/page/1793


5. Geselbracht, Margaret J.; Cappellari, Ann M.; Ellis, Arthur B.; Rzeznik, Maria A.; Johnson, Brian J. "Rare Earth Iron
Garnets: Their Synthesis and Magnetic Properties." J. Chem. Educ. 1994, 71, 696.
6. Shimada, Hiroshi; Yasuoka, Takashi; Mitsuzawa, Shunmei. "Observation of paramagnetic property of oxygen by simple
method: A simple experiment for college chemistry and physics courses (TD)." J. Chem. Educ. 1990, 67, 63.

7/4/2021 EPR.3 https://chem.libretexts.org/@go/page/1793


EPR: Parallel Mode Operation
This module presents the theory that describes how EPR transitions can be induced in integer high spin systems by the
application of a modulating magnetic field parallel to the bond axis (z-axis), as well as some of the applications of this technique
to various molecular systems.

Introduction
Standard mode Electron Paramagnetic Resonance (EPR), in which the modulating magnetic field is perpendicular to the applied
field, is capable of detecting transitions between eigenstates in systems with fractional spins (i.e.; S=1/2, 3/2, ...), so called
'Kramer Systems'. This provides a sensitive experimental technique for detecting the electronic environment of unpaired electrons
in various molecular systems. There are, however, systems with an integer spin value (ie; S=2), called 'Non-Kramer Systems'. In
order to probe the electronic environment of such systems a method in which the modulating field is parallel to the applied field
is employed. This method is called Parallel Mode EPR.

Spin Hamiltonian
To describe the theory of parallel mode EPR we must first define the system that we wish to characterize. If we have a ligand
field that has axial symmetry, the collection of 2S+1 states of the system will be split, in the case of integral spin systems, into S
doublets and a singlet. If we consider a system with S=2, the five states will be split into two doublets and a singlet. Application
of a magnetic field gives the following spin Hamiltonian for the system.
1
^ 2 2 2
H = D[ Sz − S(S + 1)] + E(Sx − Sy ) + βB ⋅ g ⋅ S (EPR.1)
3

1
2 2 2
= D[ Sz − S(S + 1)] + E(Sx − Sy ) + g∥ βBSz cos θ + g⊥ βBSz sin θ (EPR.2)
3

where the D is the axial field splitting term and E is the rhombic splitting term, these terms are a measure of the energy gap
between the various states when no magnetic field is applied. These values are a reflection of the symmetry of the system. B is
the magnetic field vector, β is the Bohr magneton, Sz is the z projection of the spin, and θ is the angle of the applied magnetic
field with respect to the symmetry axis of the system in the zx-plane.

Eigenfunctions and Eigenstates


If we have the condition that
gβB ≪ D (EPR.3)

and we neglect the rhombic field splitting term in the spin Hamiltonian, then the states, and their respective energy values, are
given by the following, to first order in perturbation theory.
∣±2⟩ ⇒ E±2 = 2D ± 2 g∥ βB cos θ (EPR.4)

∣±1⟩ ⇒ E±1 = −D ± g∥ βB cos θ (EPR.5)

∣0⟩ → E0 = −2D (EPR.6)

Analysis of the energy matrix of these states shows that there are no non-zero elements for the terms that represent transitions in
the doublet states. Therefore these states have no allowed transitions. We will employ a mathematical trick to describe these states
in a way that allows us to ascertain the EPR transitions that are allowed for the given system. We will adopt a new set of basis
functions
∣ 2 ⟩, ∣ 2 ⟩, ∣ 1 ⟩, and ∣ 1 ⟩
s a s a

which are symmetric and antisymmetric linear combinations, respectively, of the form

s
1
∣2 ⟩ = – (∣ +2⟩+ ∣ −2⟩) (EPR.7)
√2

a
1
∣2 ⟩ = – (∣ +2⟩− ∣ −2⟩) (EPR.8)
√2

6/21/2021 EPR.1 https://chem.libretexts.org/@go/page/1796


Since these basis functions are linear combinations of basis functions that are eigenfunctions of the spin Hamiltonian, it follows
that these linear combinations are also eigenfunctions of the spin Hamiltonian. When this change of basis is adopted, and we
account for the rhombic field splitting term in the spin Hamiltonian we obtain the following energies, correct to second order in
perturbation theory.
2 2
(g⊥ βB sin θ) 1 Δ2
2 1/2
E±2 = 2D + + Δ2 ± [(2 g∥ βB cos θ) +( ) ] (EPR.9)
E2 − E1 2 2

2 1/2
2 2 2
(g⊥ βB sin θ) 3(g⊥ βB sin θ) 3(g⊥ βB sin θ)
2
E±1 = −D + + ± [(g∥ βB cos θ) + (3E + ) ] (EPR.10)
E1 − E2 2(E1 − E0 ) 2(E1 − E0 )

2 2
12E 3(g⊥ βB sin θ)
E0 = −2D + + (EPR.11)
E0 − E2 E0 − E1

If we have no rhombic splitting term (E=0), there are no transitions in the ∣±2⟩ doublet that are allowed. However, for the ∣±1⟩
doublet, the states are actually a linear combination of the ∣1⟩, ∣0⟩, and ∣−1⟩ states and, due to this admixture, for some value of θ
there is a weakly allowed transition. This is due to the second order Zeeman effects. This transition becomes forbidden in the case
where θ = 0 , ie; when the magnetic field is oriented along the z-axis.
If E is not equal to zero then the ∣±2⟩ states are also linear combinations. If θ = 0 , we have energy levels for the ∣±2⟩ states that
are given by the following
1 2 2 1/2
E±2 = ± [(4 g∥ βB cos θ) +Δ ] (EPR.12)
2
2

2
12E
where Δ 2 = . The states of this doublet can then be written in the form
E2 − Eo


∣+2 ⟩ = cos α ∣ +⟩ + sin α ∣ −⟩ (EPR.13)


∣−2 ⟩ = cos α ∣ +⟩ − sin α ∣ −⟩ (EPR.14)

These new states, along with their respective energies and transitions, are shown schematically in Figure (1).

Transitions and the Resonance Condition


The states of the doublet described above have a matrix element that couples them together that is given by
′ ′
⟨2 + ∣ 4 g∥ β Sz ∣ − ⟩ = 2 g∥ β sin 2α (EPR.15)

This shows that there are allowed transitions in the ∣±2⟩ manifold as long as the modulating magnetic field is oriented with the z
axis. The resonance condition for these transitions becomes
2 2 1/2
ΔE = hν = [(4 g∥ βB cos θ) +Δ ] (EPR.16)

and the probability of a transition occurring is given by

6/21/2021 EPR.2 https://chem.libretexts.org/@go/page/1796


2
2 2
Δ
2
∣ μz ∣ = 4 g β (EPR.17)
∥ 2
(hν )

We have shown that, by taking linear combinations of the ∣±1⟩ and ∣±2⟩ states that show no transitions, we can construct a
description of the system that accounts for the transitions observed in the parallel mode EPR spectra of integer spin systems.

Applications
The parallel mode electron paramagnetic resonance technique, in which the modulating magnetic field is parallel to the applied
field, allows for the detection of transitions between eigenstates for systems with integer spin. This technique has been applied to
a variety of systems to ascertain the nature of the spin states, some of which are described below.

Biophysical Applications
Both heme and non-heme iron proteins have been shown to posses high spin ferrous ions. The parallel mode EPR technique has
been applied to several of these systems to verify their integer spin state. These proteins include the mononuclear ferrous sites of
myoglobin and transferrin, as well as the polynuclear ferrous sites of methane monoxygenase, ferredoxin II, and aconitase. It has
also been used to study the nature of the Mn(III)-Salen system and it's catalysis of the epoxidation of cis-beta-methylstyrene.

Inorganic Applications
Advances have been made in the field of synthetic inorganic chemistry such that there are now examples in the literature of
synthetic high spin non-heme iron(IV) systems. This systems are predicted to have a high degree of reactivity and are being
probed for their catalytic properties. The parallel mode EPR technique has been applied to these systems to extract the true nature
of their spin state and to measure the g values of their absorptions. An example of a parallel mode EPR spectrum of a high spin
(S=2) Fe(IV)-oxo compound, measured by Hendrich and coworkers at Pennsylvania State University, is shown in figure 2.

Figure 2: Parallel mode EPR spectrum of a S=2, Fe(IV)-oxo species.


Parallel mode EPR has also been used to describe the spin state of other high spin inorganic complexes, including an S=6
chromium system synthesized by Piligko and coworkers at the University of Manchester.

Outside Links
1. CalEPR facility website, www.brittepr.ucdavis.edu.

References
1. Abragam, A. and Bleaney, B.; Electron Paramagnetic Resonance of Transition Ions; Dover Publishing; 1986.
2. Weil, J. A. & Bolton, J. R. (2007). Electron Paramagnetic Resonance Spectroscopy: Elementary Theory and Applications,
Second Edition. Wiley-Interscience.
3. Hendrich, M. et al.; Biophysical Journal; (1989), 56, 489-500.
4. Campbell, et al.; J. Am. Chem. Soc.; 2001; 123; 5710-5719.
5. Piligko, et al.; Physical Review B; 2004; 69; 134424.
6. Hendrich, M. et al.; J.A.C.S.; 2010; 132, 1-3.

Contributors and Attributions


Damon Robles (UCD)

6/21/2021 EPR.3 https://chem.libretexts.org/@go/page/1796


EPR: Theory
Electron Paramagnetic Resonance (EPR), also called Electron Spin Resonance (ESR), is a branch of magnetic resonance
spectroscopy which utilizes microwave radiation to probe species with unpaired electrons, such as radicals, radical cations,
and triplets in the presence of an externally applied static magnetic field. In many ways, the physical properties for the basic
EPR theory and methods are analogous to Nuclear Magnetic Resonance (NMR). The most obvious difference is that the direct
probing of electron spin properties in EPR is opposed to nuclear spins in NMR. Although limited to substances with unpaired
electron spins, EPR spectroscopy has a variety of applications, from studying the kinetics and mechanisms of highly reactive
radical intermediates to obtaining information about the interactions between paramagnetic metal clusters in biological
enzymes. EPR can even be used to study the materials with conducting electrons in the semiconductor industry.

Historical Development of EPR


In 1896, the line splitting in optical spectra in a static magnetic field was first found by a Dutch physicist Zeeman. In 1920s,
Stern and Gerlach sent a beam of silver atoms through an inhomogeneous magnetic field and the beam splits into two distinct
parts, indicating the intrinsic angular momentum of electrons and atoms. Then Uhlenbeck and Goudsmit proposed that the
electrons have an angular momentum. In 1938, Isidor Rabi measured the magnetic resonance absorption of lithium chloride
molecules, which means he could measure different resonances to get more detailed information about molecular structure.
After World War II, microwave instrumentation’s widespread availability sped up the development of electron paramagnetic
resonance (EPR). The first observation of a magnetic resonance signal was detected by a Soviet physicist Zavoisky in several
salts, including hydrous copper chloride, copper sulfate and manganese sulfate in 1944. Later the Oxford group proposed the
basic theory of magnetic resonance. Contributed by many researchers, such as Cummerow & Halliday and Bagguley &
Griffiths, EPR was extensively studied. Between 1960 and 1980, continuous wave (CW) EPR was developed and pulsed EPR
was mainly studied in Bell laboratories. EPR was usually applied for organic free radicals. In the 1980s, the first commercial
pulsed EPR spectrometer appeared in the market and was then extensively used for biological, medical field, active oxygen
and so on. Nowadays, EPR has become a versatile and standard research tool.

Comparison between EPR and NMR


EPR is fundamentally similar to the more widely familiar method of NMR spectroscopy, with several important
distinctions. While both spectroscopies deal with the interaction of electromagnetic radiation with magnetic moments of
particles, there are many differences between the two spectroscopies:
1. EPR focuses on the interactions between an external magnetic field and the unpaired electrons of whatever system it is
localized to, as opposed to the nuclei of individual atoms.
2. The electromagnetic radiation used in NMR typically is confined to the radio frequency range between 300 and 1000
MHz, whereas EPR is typically performed using microwaves in the 3 - 400 GHz range.
3. In EPR, the frequency is typically held constant, while the magnetic field strength is varied. This is the reverse of how
NMR experiments are typically performed, where the magnetic field is held constant while the radio frequency is
varied.
4. Due to the short relaxation times of electron spins in comparison to nuclei, EPR experiments must often be performed
at very low temperatures, often below 10 K, and sometimes as low as 2 K. This typically requires the use of liquid
helium as a coolant.
5. EPR spectroscopy is inherently roughly 1,000 times more sensitive than NMR spectroscopy due to the higher
frequency of electromagnetic radiation used in EPR in comparison to NMR.
It should be noted that advanced pulsed EPR methods are used to directly investigate specific couplings between
paramagnetic spin systems and specific magnetic nuclei. The most widely application is Electron Nuclear Double
Resonance (ENDOR). In this method of EPR spectroscopy, both microwave and radio frequencies are used to perturb the
spins of electrons and nuclei simultaneously in order to determine very specific couplings that are not attainable through
traditional continuous wave methods.

Origin of the EPR Signal

6/15/2021 EPR.1 https://chem.libretexts.org/@go/page/1794


An electron is a negatively charged particle with certain mass, it mainly has two kinds of movements. The first one is spinning
around the nucleus, which brings orbital magnetic moment. The other is "spinning" around its own axis, which brings spin
magnetic moment. Magnetic moment of the molecule is primarily contributed by unpaired electron's spin magnetic moment.
−−−−−−− h
MS = √ S(S + 1) (EPR.1)

MS is the total spin angular moment,


S is the spin quantum number and
h is Planck’s constant.
In the z direction, the component of the total spin angular moment can only assume two values:
h
MS = mS ⋅ (EPR.2)
Z

The term ms have (2S + 1) different values: +S, (S − 1), (S − 2),.....-S. For single unpaired electron, only two possible values
for ms are +1/2 and −1/2.
The magnetic moment, μe is directly proportional to the spin angular momentum and one may therefore write
μe = −ge μB MS (EPR.3)

The appearance of negative sign due to the fact that the magnetic mom entum of electron is collinear, but antiparallel to the
spin itself. The term (geμB) is the magnetogyric ratio. The Bohr magneton, μB, is the magnetic moment for one unit of
quantum mechanical angular momentum:
eh
μB = (EPR.4)
4πme

where e is the electron charge, me is the electron mass, the factor ge is known as the free electron g-factor with a value of 2.002
319 304 386 (one of the most accurately known physical constant). This magnetic moment interacts with the applied magnetic
field. The interaction between the magnetic moment (μ) and the field (B) is described by

E = −μ ⋅ B (EPR.5)

For single unpaired electron, there will be two possible energy states, this effect is called Zeeman splitting.
1
E = gμB B (EPR.6)
1
2
+
2

1
E =− gμB B (EPR.7)
1
2

2

In the absence of external magnetic field,

E+1/2 = E−1/2 = 0 (EPR.8)

However, in the presence of external magnetic field (Figure 1), the difference between the two energy states can be written as
ΔE = hv = gμB B (EPR.9)

Figure 1: Energy levels for an electron spin (MS = ±1/2) in an applied magnetic field B.

6/15/2021 EPR.2 https://chem.libretexts.org/@go/page/1794


With the intensity of the applied magnetic field increasing, the energy difference between the energy levels widens until it
matches with the microwave radiation, and results in absorption of photons. This is the fundamental basis for EPR
spectroscopy. EPR spectrometers typically vary the magnetic field and hold the microwave frequency. EPR spectrometers are
available in several frequency ranges, and X band is currently the most commonly used.
Table 1: Different Microwave bands for EPR Spectroscopy
Microwave band Frequency/GHz Wavelength/cm B (electron)/Tesla

S 3.0 10.0 0.107

X 9.5 3.15 0.339


K 23 1.30 0.82
Q 35 0.86 1.25
W 95 0.315 3.3

Energy Level Structure and the g-factor


EPR is often used to investigate systems in which electrons have both orbital and spin angular momentum, which necessitates
the use of a scaling factor to account for the coupling between the two momenta. This factor is the g-factor, and it is roughly
equivalent in utility how chemical shift is used in NMR. The g factor is associated with the quantum number J, the total
angular momentum, where J = L + S .
J(J + 1)(gL + gs ) + (L(L + 1) − S(S + 1))(gL − gs )
gJ = (EPR.10)
2J(J + 1)

Here, g is the orbital g value and gs is the spin g value. For most spin systems with angular and spin magnetic momenta, it
L

can be approximated that gL is exactly 1 and gs is exactly 2. T his equation reduces to what is called the Landé formula:

3 L(L + 1) − S(S + 1))


gJ = − (EPR.11)
2 2J(J + 1)

And the resultant electronic magnetic dipole is:


μJ = −gJ μB J (EPR.12)

In practice, these approximations do not always hold true, as there are many systems in which J-coupling does occur,
especially in transition metal clusters where the unpaired spin is highly delocalized over several nuclei. But for the purposes of
a elementary examination of EPR theory it is useful for the understanding of how the g factor is derived. In general this is
simply referred to as the g-factor or the Landé g-factor.
The g-factor for a free electron with zero angular momentum still has a small quantum mechanical corrective g value, with
g=2.0023193. In addition to considering the total magnetic dipole moment of a paramagnetic species, the g-value takes into
account the local environment of the spin system. The existence of local magnetic fields produced by other paramagnetic
species, electric quadrupoles, magnetic nuclei, ligand fields (especially in the case of transition metals) all can change the
effective magnetic field that the electron experiences such that

Bef f = B0 + Blocal (EPR.13)

These local fields can either:


1. be induced by the applied field, and hence have magnitude dependence on B o r are 0

2. permanent and independent of B o ther than in orientation.


0

In the case of the first type, it is easiest to consider the effective field experienced by the electron as a function of the applied
field, thus we can write:
Bef f = B0 (1 − σ) (EPR.14)

where σ is the shielding factor that results in decreasing or increasing the effective field. The g-factor must then be replaced by
a variable g factor geff such that:

6/15/2021 EPR.3 https://chem.libretexts.org/@go/page/1794


g
Bef f = B0 ⋅ ( ) (EPR.15)
gef f

Many organic radicals and radical ions have unpaired electrons with L near zero, and the total angular momentum quantum
number J becomes approximately S. As result, the g-values of these species are typically close to 2. In stark contrast, unpaired
spins in transition metal ions or complexes typically have larger values of L and S, and their g values diverge from 2
accordingly.
After all of this, the energy levels that correspond to the spins in an applied magnetic field can now be written as:

Ems = ms ge μB B0 (EPR.16)

And thus the energy difference associated with a transition is given as:

ΔEms = Δms ge μB B (EPR.17)

Typically, EPR is performed perpendicular mode, where the magnetic field component of the microwave radiation is oriented
perpendicular to the magnetic field created by the magnet. Here, the selection rule for allowed EPR transitions is Δm = ±1 , s

so the energy of the transition is simply:


(17)
ΔEms = ge μB B (EPR.18)

There is a method called Parallel Mode EPR in which the microwaves are applied parallel to the magnetic field, changing the
selection rule to Δm = ±2 . This is more fully explained in the Parallel Mode EPR: Theory module.
s

Sensitivity
At the thermal equilibrium and external applied magnetic field, the spin population is split between the two Zeeman levels
(Figure 1) according to the Maxwell–Boltzmann law. Absorption can occur as long as the number of particles in the lower
state is greater than the number of particles in the upper state. At equilibrium, the ratio predicted by the Boltzmann
distribution:
−ΔE gμB
Nupper −
kB T kB T
=e =e (EPR.19)
Nlower

with kB is the Boltzmann constant.


At regular temperatures and magnetic fields, the exponent is very small and the exponential can be accurately approximated by
the expansion,
–x
e ≈ 1– x (EPR.20)

Thus
Nupper gμB
=1− (EPR.21)
Nlower kB T

At 298 K in a field of about 3000 G the distribution shows that Nupper /Nlower=0.9986, which means the difference between is
Nupperand Nlower is very small. The populations of the two Zeeman levels are nearly the same, but the slight excess in the lower
level gives rise to a net absorption.
gμB N gμB
Nlower − Nupper = Nlower [1 − (1 − )] = (EPR.22)
kB T 2 kB T

This expression tells us that EPR sensitivity (net absorption) increases as temperature decreases and magnetic field strength
increases, and magnetic field is proportional to microwave frequency. Theoretically speaking, the sensitivity of spectrometer
with K-band or Q-band or W-band shoulder be greater than spectrometer with X-band. However, since the K-, Q- or W-band
waveguides are smaller, samples are necessarily smaller, thus canceling the advantage of a more favorable Boltzmann factor.

Spin Operators and Hamiltonians

6/15/2021 EPR.4 https://chem.libretexts.org/@go/page/1794


Any system which has discrete energy levels and is described by defined quantum numbers can be represented by an
eigenvalue equation, such that if we define an operator (Λ
^
) that is appropriate to the property being observed, the
eigenfunction equation is:
^
Λψ =  λk ψ (EPR.23)
k k

Here λk is an eigenvalue of a state “k” for which the eigenfunction is ψk. EPR is most concerned with the quantization of spin
angular momentum, therefore, the operator must be defined is a spin operator that operated on a function that describes a spin
state. In the case of a system with a total electron spin of S = ½, the two states are described by the quantum numbers Ms =
+1/2 and Ms = -1/2, which measure the components Ms of angular momentum along the z-direction of the magnetic field. In
most systems, it is convenient to treat the direction of the magnetic field as the z-direction, and thus the spin operator is
denoted Ŝz, where Ŝ is the angular momentum operator. So, omitting the k index, the z-component of the angular momentum
operator can be written as:
^
Sz ϕe = Ms ϕe (EPR.24)

where ms is the eigenvalue of the operator Sz, and ϕ e(Ms) is the corresponding eigenfunction. Adopting the α- notation for
spin states, where α(e) = ϕ e(Ms=+1/2) and β(e) = ϕ e(Ms=-1/2), this expression can be written:
1
^
Sz α (e) = + α (e) (EPR.25)
2

1
^
Sz β (e) = − β (e) (EPR.26)
2

In a similar fashion, the eigenfunctions for the nuclear spin operator for a nucleus with spin = ½ can be written:
1
^ α (n) = +
I α (n) (EPR.27)
z
2

1
^ β (n) = −
I β (n) (EPR.28)
z
2

Written in the convenient Dirac notation, these expressions become:


1∣
^ ∣
Sz ∣ α (e)⟩ = + ∣ α (e)⟩ (EPR.29)
2∣

1∣
^ ∣
Sz ∣ β (e)⟩ = + ∣ β (e)⟩ (EPR.30)
2∣

and
1∣
^
Iz ∣
∣ α (n)⟩ = + ∣ α (n)⟩ (EPR.31)
2∣

1∣
^∣
Iz ∣ β (n)⟩ = + ∣ β (n)⟩ (EPR.32)
2∣

Using the time-independent Schrödinger equation, we can define the energies associated with the systems described by these
equations as such:
^
He ∣
∣ ϕek ⟩ = Eek | ϕek ⟩ (EPR.33)

^
Hn ∣
∣ ϕnk ⟩ = Enk | ϕnk ⟩ (EPR.34)

So that
^
He ∣ ∣
∣ α(e)⟩ = Eα(e) ∣ α(e)⟩ (EPR.35)

^
He ∣ ∣
∣ β(e)⟩ = Eα(e) ∣ β(e)⟩ (EPR.36)

6/15/2021 EPR.5 https://chem.libretexts.org/@go/page/1794


^
Hn ∣ ∣
∣ α(n)⟩ = Eα(n) ∣ α(n)⟩ (EPR.37)

^
Hn ∣ ∣
∣ β(n)⟩ = Eα(n) ∣ β(n)⟩ (EPR.38)

Here Ĥ is the Hamiltonian operator and represents the operator for the total energy, and commutes with both I and S operators.

Electron/Nuclear Zeeman Interactions using Operators


Using the Hamiltonians derived in the last section, we can develop hamiltonians for the perturbed case in which an external
magnetic field is introduced. For the simple case of the hydrogen atom with S=1/2 and I=1/2, interaction with a strong
magnetic field oriented along the z-direction will be considered. Using the operator form, the Hamiltonian takes the form:
^
 H = −Bμ
^ (EPR.39)
z

Here, the electron magnetic moment operator μez is proportional to the electron spin operator. Likewise, the nuclear magnetic
moment operator μnzis proportional to the nuclear spin operator Iz. Therefore,
^ ^
 μ
^ =  γ S z h =   − gμ Sz (EPR.40)
ez e B

^ ^ ^
 μ nz
=  γn I z h =   + gn μn I z (EPR.41)

Now the electron and nuclear spin Hamiltonians can be defined as:
^ ^
 He =  ge μ Sz (EPR.42)
B

^ ^
 Hn =   − gn μn Iz (EPR.43)

We now have a quantum mechanical framework for the energies of electronic and nuclear spin states that will be further useful
in developing a description of the interactions between the magnetic moments of the two classes of particles.

Nuclear Hyperfine Structure


According to the figure 1, we should observe one spectra line in a paramagnetic molecule, but in reality, we usually observe
more than one split line. The reason for that is hyperfine interactions, which results from interaction of the magnetic moment
of the unpaired electron and the magnetic nuclei. The hyperfine patterns are highly valuable when it comes to determine the
spatial structure of paramagnetic species and identify the paramagnetic species. As a result, nuclear spins act as probes which
are sensitive to the magnetitude and direction of the field due to the unpaired electron.
In general, there are two kinds of hyperfine interactions between unpaired electron and the nucleus. The first is the interaction
of two dipoles. We refer it as the anisotropic or dipolar hyperfine interaction, which is the interaction between electron spin
magnetic moment and the nuclei magnetic moment, and it depends on the shape of electronic orbital and the average distance
of electron and nucleus. This interaction can help us to determine the possible position of a paramagnetic species in a solid
lattice.
The second interaction is known as the Fermi contact interaction, and only takes the electrons in s orbital into consideration,
since p, d and f orbitals have nodal planes passing through the nucleus. We refer to this type of interaction as isotropic, which
depends on the presence of a finite unpaired electron spin density at the position of the nucleus, not on the orientation of the
paramagnetic species in the magnetic field.
8 2
A =− π ⟨μn ⋅ μe ⟩ ⋅ |ψ (0)| (EPR.44)
3

A is the isotropic hyperfine coupling constant and is related to the unpaired spin density, μn is the nuclear magnetic moment,
μe is the electron magnetic moment and Ψ(0) is the electron wavefunction at the nucleus. The Fermi contact interaction
happens in s orbital when electron density is not zero. Thus nuclear hyperfine spectra not only includes the interaction of
nuclei and their positions in the molecule but also the extent to which part or all of the molecule is free to reorientate itself
according to the direction of the applied magnetic field.

Isotropic Hyperfine Interaction


In the case of one unpaired electron, the spin hamiltonian can be written as below for the isotropc part of nuclear hyperfine
interaction.

6/15/2021 EPR.6 https://chem.libretexts.org/@go/page/1794


H = HEZ − HN Z − HH F S (EPR.45)

EZ means electron Zeeman, NZ means nuclear Zeeman and HFS represents hyperfine interaction. The equation can also be
written as
^
H = gμB H SZ − gN μN ⋅ BIZ + h ⋅ S ⋅ aI (EPR.46)

The term aS*I is introduced by Fermi contact interaction. I is the nucleus spin, H is the external field. Since μBis much larger
than μN, the equation can take the form as:
^
H = gμB H SZ + h ⋅ S ⋅ aI (EPR.47)

When one unpaired electron interacts with one nucleus, the number of EPR lines is 2I+1. When one unpaired electron interacts
with N equivalent nuclei, the number of EPR lines is 2NI+1. When one electron interacts with nonequivalent nucleis (N1,
N2.....), the numer of EPR lines is
k

∏(2 Ni Ii + 1) (EPR.48)

i=1

In the case of DPPH, I=1 and two nitrogen nucleui are equivalent. 2NI+1=5, we can get five lines: 1:2:3:2:1.

Figure 2: DPPH
The table below shows the relative intensities of the lines according to unpaired electrons interacting with multiple equivalent
nuclei.
Number of Equivalent Nuclei Relative Intensities

1 1:1

2 1:2:1
3 1:3:3:1
4 1:4:6:4:1
5 1:5:10:10:5:1
6 1:6:15:20:15:6:1

We can observe that increasing number of nucleuses leads to the complexity of the spectrum, and spectral density depends on
the number of nuclei as equation shown below:
k
∏i=1 2 Ni Ii + 1
SpectraldensityEP R = (EPR.49)
k
∑ 2 | ai | Ni Ii
i=1

a is the isotropic hyperfine coupling constant.

The g Anisotropy
From the below equation, we can calculate g in this way:

ΔE = hv = gμB B (EPR.50)

If the energy gap is not zero, g factor can be remembered as:

1 ν [GH Z]
g ≈ (EPR.51)
14 B [T ]

6/15/2021 EPR.7 https://chem.libretexts.org/@go/page/1794


The g factor is not necessarily isotropic and needs to be treated as a tensor g. For a free electron, g factor is close to 2. If
electrons are in the atom, g factor is no longer 2, spin orbit coupling will shift g factor from 2. If the atom are placed at an
electrostatic field of other atoms, the orbital energy level will also shift, and the g factor becomes anisotropic. The anisotropies
lead to line broadening in isotropic ESR spectra. The Electron-Zeeman interaction depends on the absolute orientation of the
molecule with respect to the external magnetic field. Anisotropic is very important for free electrons in non-symmetric orbitals
(p,d).
In a more complex spin system, Hamiltonian is required to interpret as below:
^ ⃗  ^ ⃗  ⃗ 
H s = μB B ⋅ g ⋅ S + ∑ I i ⋅ Ai ⋅ S (EPR.52)

g and Ai are 3*3 matrices representing the anisotropic Zeeman and nuclear hyperfine interactions, thus it is more accurate to
describe g-factor as a tensor like:
−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−
2 2 2 2 2 2 2 2
g = √ gx ⋅ si n α ⋅ cos β + gy ⋅ si n α ⋅ cos β + gz cos α (EPR.53)

Alpha and beta is the angle between magnetic field with respect to principle axis of g tensor. If gx=gy, it can be expressed as:
−−−−−−−−−−−−−−−−
2 2 2 2
g = √ gx ⋅ si n α + gz cos α (EPR.54)

Thus, we can identify the g tensor by measuring the angular dependence in the above equation.

Spin Relaxation Mechanisms


The excess population of lower state over upper state for a single spin system is very small as we can calculate from the
following example. With the temperature of 298K in a magnetic field of 3000G, Nupper /Nlower=0.9986, which means the
populations of the two energy levels are almost equal, yet the slight excess in the lower level leads to energy absorption. In
order to maintain a population excess in the lower level, the electrons from the upper level give up the hν energy to return to
the lower level to satisfy the Maxwell–Boltzmann law. The process of this energy releasing is called spin relaxation process, of
which there are two types, known as spin–lattice relaxation and spin–spin relaxation.

Spin-lattice relaxation
this implies interaction between the species with unpaired electrons, known as "spin system" and the surrounding molecules,
known as "lattice". The energy is dissipated within the lattice as vibrational, rotational or translational energy. The spin lattice
relaxation is characterized by a relaxation time T1e, which is the time for the spin system to lose 1/eth of its excess energy.
Rapid dissipation of energy (short T1e) is essential if the population difference of the spin states is to be maintained. Slow spin-
lattice relaxation, which is of frequent occurence in systems containing free radicals, especially at low temperatures, can cause
saturation of the spin system. This means that the population difference of the upper and lower spin states approaches zero,
and EPR signal ceases.

Spin-spin relaxation
Spin-spin relaxation or Cross relaxation, by which energy exchange happens between electrons in a higher energy spin state
and nearby electrons or magnetic nuclei in a lower energy state, without transfering to the lattice. The spin–spin relaxation can
be characterized by spin-spin relaxation time T2e.
When both spin–spin and spin–lattice relaxations contribute to the EPR signal, the resonance line width (ΔB) can be written as
1 1
ΔB ∝ + (EPR.55)
T1e T2e

From the equation, we can tell that when T1e > T2e, ΔB depends primarily on spin–spin interactions. Decreasing the spin-spin
distance, which is the spin concentration, T1e will become very short, approximately below roughly 10−7 sec, thus the spin-
lattice relaxation will have a larger influence on the linewidth than spin-spin ralaxation. In some cases, the EPR lines are
broadened beyond detection. When a spin system is weakly coupled to the lattice, the system tends to have a long T1e and
electrons do not have time to return to the ground state, as a result the population difference of the two levels tends to approach
zero and the intensity of the EPR signal decreases. This effect, known as saturation, can be avoided by exposing the sample to
low intensity microwave radiation. Systems with shorter T1e are more difficult to saturate.

6/15/2021 EPR.8 https://chem.libretexts.org/@go/page/1794


Parting Thoughts
Although EPR is limited to investigation of compounds and materials with unpaired electrons, it is undoubtedly the most
direct and useful spectroscopic method for probing the properties of these specific systems. Another advantage is that sample
preparation is simple and EPR does not cause destruction or activation in the sample. By probing the fundamental splitting of
energy levels of spins with regard to their orientation in an external magnetic field, interactions between paramagnetic spin
systems and their local environments can be detected. EPR spectra is highly sensitive to the local electonic structure, oxidation
state and the proximity of magnetic nuclei to the system in question.

References
1. Abragam, A. and Bleaney, B. Electron Paramagnetic Resonance of Transition Ions; Dover Publishing; 1986.
2. Poole, C.P. Electron Spin Resonance; Interscience Publishers; 1967.
3. Schweiger, A. & Jeschke, G. Principles of Pulse Electron Paramagnetic Resonance. Oxford University Press; 2001.
4. Weil, J. A. & Bolton, J. R. Electron Paramagnetic Resonance Spectroscopy: Elementary Theory and Applications, Second
Edition. Wiley-Interscience; 2007.
5. Gale, Robert J ed. Spectroelectrochemistry: theory and practice. Plenum Press; 1988.
6. Ayscough, Peter B. Electron Spin Resonance in Chemistry. Butler & Tanner Ltd; 1967.

Problems
1. Why are microwaves necessary to study the electron spin resonances?
2. What is the g-factor? What does it mean if the value is 2?

Answer
1. EPR is targeted to unpaired electrons. Single unpaired electron behaves like a small magnetic bar when placed in a large
maganetic field. It will orient itself parallel to the large magnetic field. At a particular magnetic field inensity, the
microwave irradiation will induce unpaired electrons to orient against the large magnetic field. This effect will cause
Zeeman splitting, when the energy difference between the lower and the higher energy level matches the microwave
frequency. There will be absorption of energy, thus producing an EPR resonance, and detected by the spectrometer.
2. As we have discussed before, g-factor can be calculated from measuring the magnetic field and frequency. g-factor is an
effective Zeeman factor. For a free electron, isotropic ge = 2.0023193043617. It is predicted by quantum electrodynamics.
This free electron has spin angular momentum but no orbital angular momentum. For electrons in an atom, g factor will
shift from ge due to spin orbit coupling. thus g factor is the characteristic of different electronic structures.

Contributors and Attributions


Parminder Kaur, Paul Oyala and Yu Guo.

6/15/2021 EPR.9 https://chem.libretexts.org/@go/page/1794


EPR - Interpretation
Electron paramagnetic resonance spectroscopy (EPR), also called electron spin resonance (ESR), is a technique used to study
chemical species with unpaired electrons. EPR spectroscopy plays an important role in the understanding of organic and
inorganic radicals, transition metal complexes, and some biomolecules. For theoretical background on EPR, please refer to
EPR:Theory.

Introduction
Like most spectroscopic techniques, EPR spectrometers measure the absorption of electromagnetic radiation. A simple
absorption spectra will appear similar to the one on the top of Figure 1. However, a phase-sensitive detector is used in EPR
spectrometers which converts the normal absorption signal to its first derivative. Then the absorption signal is presented as its
first derivative in the spectrum, which is similar to the one on the bottom of Figure 1. Thus, the magnetic field is on the x-axis
of EPR spectrum; dχ″/dB, the derivative of the imaginary part of the molecular magnetic susceptibility with respect to the
external static magnetic field in arbitrary units is on the y-axis. In the EPR spectrum, where the spectrum passes through zero
corresponds to the absorption peak of absorption spectrum. People can use this to determine the center of the signal. On the x-
axis, sometimes people use the unit “gauss” (G), instead of tesla (T). One tesla is equal to 10000 gauss.

Figure 1. Comparison of absorption spectrum and EPR spectrum. Image used with permisison (Public Domain).

Proportionality factor (g-factor)


As a result of the Zeeman Effect, the state energy difference of an electron with s=1/2 in magnetic field is
ΔE = gβB (1)

where β is the constant, Bohr magneton. Since the energy absorbed by the electron should be exactly the same with the state
energy difference ΔE, ΔE=hv ( h is Planck’s constant), the Equation 1 can be expressed as
hν = gβB (2)

People can control the microwave frequency v and the magnetic field B. The other factor, g, is a constant of proportionality,
whose value is the property of the electron in a certain environment. After plugging in the values of h and β in Equation 2, g
value can be given through Equation 3:
g = 71.4484v(in GHz)/B (in mT) (3)

A free electron in vacuum has a g value ge= 2.00232. For instance, at the magnetic field of 331.85 mT, a free electron absorbs
the microwave with an X-band frequency of 9.300 GHz. However, when the electron is in a certain environment, for example,
a transition metal-ion complex, the second magnetic field produced by the nuclei, ΔB, will also influence the electron. At this
kind of circumstance, Equation 2 becomes
hν = gβ(Be + ΔB) (4)

6/21/2021 1 https://chem.libretexts.org/@go/page/1792
since we only know the spectrometer value of B, the Equation 4 is written as:
hν = (ge + Δg)βB] (5)

From the relationship shown above, we know that there are infinite pairs of v and B that fit this relationship. The magnetic
field for resonance is not a unique “fingerprint” for the identification of a compound because spectra can be acquired at
different microwave frequencies. Then what is the fingerprint of a molecule? It is Δg. This value contains the chemical
information that lies in the interaction between the electron and the electronic structure of the molecule, one can simply take
the value of g = ge+ Δg as a fingerprint of the molecule. For organic radicals, the g value is very close to ge with values
ranging from 1.99-2.01. For example, the g value for •CH3 is 2.0026. For transition metal complexes, the g value varies a lot
because of the spin-orbit coupling and zero-field splitting. Usually it ranges from 1.4-3.0, depending on the geometry of the
complex. For instance, the g value of Cu(acac)2 is 2.13. To determine the g value, we use the center of the signal. By using
Equation 3, we can calculate the g factor of the absorption in the spectrum. The value of g factor is not only related to the
electronic environment, but also related to anisotropy. About this part, please refer to EPR:Theory, Parallel Mode EPR: Theory
and ENDOR:Theory. An example from UC Davis is shown below[1] (Britt group, Published in J.A.C.S. ):

Figure 2. EPR spectra of some proteins (WT mitoNEET, H87C mitoNEET, and ferredoxin). Frequency is 30.89 GHz.1

Hyperfine Interactions
Another very important factor in EPR is hyperfine interactions. Besides the applied magnetic field B0, the compound contains
the unpaired electrons are sensitive to their local “micro” environment. Additional information can be obtained from the so-
called hyperfine interaction. The nuclei of the atoms in a molecule or complex usually have their own fine magnetic moments.
Such magnetic moments occurrence can produce a local magnetic field intense enough to affect the electron. Such interaction
between the electron and the nuclei produced local magnetic field is called the hyperfine interaction. Then the energy level of
the electron can be expressed as:
E = gmBB0MS + aMsmI (6)
In which a is the hyperfine coupling constant, mI is the nuclear spin quantum number. Hyperfine interactions can be used to
provide a wealth of information about the sample such as the number and identity of atoms in a molecule or compound, as
well as their distance from the unpaired electron.
Table 1. Bio transition metal nuclear spins and EPR hyperfine patterns[3]

6/21/2021 2 https://chem.libretexts.org/@go/page/1792
The rules for determining which nuclei will interact are the same as for NMR. For isotopes which have even atomic and even
mass numbers, the ground state nuclear spin quantum number, I, is zero, and these isotopes have no EPR (or NMR) spectra.
For isotopes with odd atomic numbers and even mass numbers, the values of I are integers. For example the spin of 2H is 1.
For isotopes with odd mass numbers, the values of I are fractions. For example the spin of 1H is 1/2 and the spin of 23Na is 7/2.
Here are more examples from biological systems:
Table 2. Bio ligand atom nuclear spins and their EPR hyperfine patterns[3]

The number of lines from the hyperfine interaction can be determined by the formula: 2NI + 1. N is the number of equivalent
nuclei and I is the spin. For example, an unpaired electron on a V4+ experiences I=7/2 from the vanadium nucleus. We can see
8 lines from the EPR spectrum. When coupling to a single nucleus, each line has the same intensity. When coupling to more
than one nucleus, the relative intensity of each line is determined by the number of interacting nuclei. For the most common
I=1/2 nuclei, the intensity of each line follows Pascal's triangle, which is shown below:

Figure 3. Pascal's triangle


For example, for •CH3, the radical’s signal is split to 2NI+1= 2*3*1/2+1=4 lines, the ratio of each line’s intensity is 1:3:3:1.
The spectrum looks like this:

6/21/2021 3 https://chem.libretexts.org/@go/page/1792
Figure 4. Simulated EPR spectrum of the •CH3 radical. en.Wikipedia.org/wiki/File:EPR_methyl.png
If an electron couples to several sets of nuclei, first we apply the coupling rule to the nearest nuclei, then we split each of those
lines by the coupling them to the next nearest nuclei, and so on. For the methoxymethyl radical, H2C(OCH3), there are
(2*2*1/2+1)*(2*3*1/2+1)=12 lines in the spectrum, the spectrum looks like this:

Figure 5. Simulated EPR spectrum of the H2C(OCH3) radical. http://en.Wikipedia.org/wiki/File:EP...hoxymethyl.png


For I=1, the relative intensities follow this triangle:

Figure 5. Relative Intensities of each line when I=1


The EPR spectra have very different line shapes and characteristics depending on many factors, such as the interactions in the
spin Hamiltonian, physical phase of samples, dynamic properties of molecules. To gain the information on structure and
dynamics from experimental data, spectral simulations are heavily relied. People use simulation to study the dependencies of
spectral features on the magnetic parameters, to predict the information we may get from experiments, or to extract accurate
parameter from experimental spectra.

EasySpin Simulations
Many methods were developed to simulate the EPR spectra. Dr. Stefan Stoll wrote EasySpin, a computational EPR
package for spectral simulation. EasySpin is based on Matlab, which is a numerical computing environment and fourth-
generation programming language. EasySpin is a powerful tool in EPR spectral simulation. It can simulate spectra under
many different conditions. Some functions are shown below:
Spectral simulations and fitting functions:
garlic: cw EPR (isotropic and fast motion)

6/21/2021 4 https://chem.libretexts.org/@go/page/1792
chili: cw EPR (slow motion)
pepper: cw EPR (solid state)
salt: ENDOR (solid state)
saffron: pulse EPR/ENDOR (solid state)
esfit: least-squares fitting
To learn more, please visit EasySpin: http://www.easyspin.org/.

References
1. Dicus, M.M; Conlan,A.; Nechushtai,R.; Jennings,P.A.;Paddock,M.L.; Britt,R.D.; Stoll S. J. AM. CHEM. SOC. 2010, 132,
2037–2049
2. Hagen,W.R. 2009. Biomolecular EPR Spectroscopy. Boca Raton: CRC Press.
3. Hagen,W.R. Dalton Trans., 2006, 4415–4434
4. Stoll,S., Schweiger,A. Journal of Magnetic Resonance 178 (2006) 42–55

Problems
1. If there is one unpaired electron in Cu2+ (I=3/2) and the copper ion is coordinated by one nitrogen atom (I=1) and one OH-
(I=1/2), how many lines can be expected in the EPR spectrum? ( (2*1*3/2+1)*(2*1*1+1)*(2*1*1/2+1)=24 )
2. For a radical, the magnetic field is 3810 G, the frequency of the microwave is 9600 MHz. What is the value of its g-factor?
−3
(71.4484)(9600 x 10 )
g = = 1.800 (6)
−1
3810 x 10

Contributors and Attributions


Pei Zhao (University of California, Davis)

6/21/2021 5 https://chem.libretexts.org/@go/page/1792
Hyperfine Splitting
This splitting occurs due to hyperfine coupling (the EPR analogy to NMR’s J coupling) and further splits the fine structure
(occurring from spin-orbit interaction and relativistic effects) of the spectra of atoms with unpaired electrons. Although
hyperfine splitting applies to multiple spectroscopy techniques such as NMR, this splitting is essential and most relevant in the
utilization of electron paramagnetic resonance (EPR) spectroscopy.

Introduction
Hyperfine Splitting is utilized in EPR spectroscopy to provide information about a molecule, most often radicals. The number
and identity of nuclei can be determined, as well as the distance of a nucleus from the unpaired electron in the molecule.
Hyperfine coupling is caused by the interaction between the magnetic moments arising from the spins of both the nucleus and
electrons in atoms. As shown in Figure 1, in a single electron system the electron with its own magnetic moment moves within
the magnetic dipole field of the nucleus.

Figure 1. B is magnetic field, μ is dipole moment, ‘N’ refers to the nucleus, ‘e’ refers to the electron
This spin interaction in turn causes splitting of the fine structure of spectral lines into smaller components called hyperfine
structure. Hyperfine structure is approximately 1000 times smaller than fine structure. Figure 2 shows a comparison of fine
structure with hyperfine structure splitting for hydrogen, though this is not to scale.

Figure 2. Splitting diagram of hydrogen


The total angular momentum of the atom is represented by F with regards to hyperfine structure. This is found simply through
the relation F=J+I where I is the ground state quantum number and J refers to the energy levels of the system.

Results of Nuclear-Electron Interactions


These hyperfine interactions between dipoles are especially relevant in EPR. The spectra of EPR are derived from a change in
the spin state of an electron. Without the additional energy levels arising from the interaction of the nuclear and electron
magnetic moments, only one line would be observed for single electron spin systems. This process is known as hyperfine
splitting (hyperfine coupling) and may be thought of as a Zeeman effect occurring due to the magnetic dipole moment of the
nucleus inducing a magnetic field.

7/3/2021 1 https://chem.libretexts.org/@go/page/1795
The coupling patterns due to hyperfine splitting are identical to that of NMR. The number of peaks resulting from hyperfine
splitting of radicals may be predicted by the following equations where Mi is the number of equivalent nuclei:
# of peaks = Mi I + 1 for atoms having one equivalent nuclei
# of peaks = (2 M1 I1 + 1)(2 M2 I2 + 1). . . .for atoms with multiple equivalent nuclei
For example, in the case of a methyl radical 4 lines would be observed in the EPR spectra. A methyl radical has 3 equivalent
protons interacting with the unpaired electron, each with I=1/2 as their nuclear state yielding 4 peaks.

Figure 3. Approximate peaks resulting from hyperfine splitting between two unequivalent protons
The relative intensities of certain radicals can also be predicted. When I = 1/2 as in the case for 1H, 19F, and 31P, then the
intensity of the lines produced follow Pascal's triangle. Using the methyl radical example, the 4 peaks would have relative
intensities of 1:3:3:1. The following figures2 show the different splitting that results from interaction between equivalent
versus nonequivalent protons.

Figure 4. Approximate peaks resulting from hyperfine splitting between two equivalent protons
It is important to note that the spacing between peaks is 'a', the hyperfine coupling constant. This constant is equivalent for
both protons in the equivalent system but unequal for the unequivalent protons.

The Hyperfine Coupling Constant


The hyperfine coupling constant (a ) is directly related to the distance between peaks in a spectrum and its magnitude
indicates the extent of delocalization of the unpaired electron over the molecule. This constant may also be calculated. The
following equation shows the total energy related to electron transitions in EPR.

ΔE = ge μe Ms B + ∑ gN μN MI (1 − σi ) + ∑ ai Ms MI (1)
i i i i

i i

The first two terms correspond to the Zeeman energy of the electron and the nucleus of the system, respectively. The third term
is the hyperfine coupling between the electron and nucleus where a is the hyperfine coupling constant. Figure 5 shows
i

splitting between energy levels and their dependence on magnetic field strength. In this figure, there are two resonances where
frequency equals energy level splitting at magnetic field strengths of B and B .
1 2

7/3/2021 2 https://chem.libretexts.org/@go/page/1795
Figure 5: Splitting between energy levels and their dependence on magnetic field strength
These parameters are essential in the derivation of the hyperfine coupling constant. By manipulating the total energy equation
(those interested in the entire derivation, refer to the first outside link), the following two relations may be derived.
hν − a/2
B1 = (2)
gμe

hν + a/2
B2 = (3)
gμe

From this, the hyperfine coupling constant (a ) may be derived where g is the g-factor.

ΔB = B2 − B1 (4)

hν + a/2 hν − a/2
= − (5)
gμe gμe

so solving for hyperfine coupling constant results in the following relationship:

a = gμe ΔB (6)

Isotropic and Anisotropic Interactions


Electron-nuclei interactions have several mechanisms, the most prevalent being Fermi contact interaction and dipole
interaction. Dipole interactions occur between the magnetic moments of the nucleus and electron as an electron moves around
a nucleus. However, as an electron approaches a nucleus, it has a magnetic moment associated with it. As this magnetic
moment moves very close to the nucleus, the magnetic field associated with that nucleus is no longer entirely dipolar. The
resulting interaction of these magnetic moments while the electron and nucleus are in contact is radically different from the
dipolar interaction of the electron when it is outside the nucleus. This non-dipolar interaction of a nucleus and electron spin in
contact is the Fermi contact interaction. A comparison of this is shown in Figure 6. The sum of these interactions is the overall
hyperfine coupling of the system.

Figure 6: Different electron-nuclei interactions resulting in hyperfine coupling


Fermi contact interactions predominate with isotropic interactions, meaning sample orientation to the magnetic field does not
affect the interaction. Due to the fact that this interaction only occurs when the electron is inside the nucleus, only electrons in
the s orbital exhibit this kind of interaction. All other orbitals (p,d,f) contain a node at the nucleus and can never have an
electron at that node. The hyperfine coupling constant in isotropic interactions is denoted 'a'.

7/3/2021 3 https://chem.libretexts.org/@go/page/1795
Dipole interactions predominate with anisotropic interactions, meaning sample orientation does change the interaction. These
interactions depend on the distance between the electron and nuclei as well as the orbital shape. The typical scheme is shown
in Figure 7.

Figure 7: Interaction between two diploes with radius 'r'


Dipole interactions can allow for positioning paramagnetic species in solid lattices. The hyperfine coupling constant in
isotropic interactions is denoted 'B'.

Superhyperfine Splitting
Further splitting may occur by the unpaired electron if the electron is subject to the influence of multiple sets of
equivalent nuclei. This splitting is on the order of 2nI+1 and is known as superhyperfine splitting. As hyperfine
structure splits fine structure into smaller components, superhyperfine structure further splits hyperfine structure. As a
result, these interactions are extremely small but are useful as they can be used as direct evidence for covalency. The more
covalent character a molecule exhibits, the more apparent its hyperfine splitting.
For example, in a CH2OH radical, an EPR spectrum would show a triplet of doublets. The triplet would arise from the
three protons, but superhyperfine splitting would cause these to split futher into doublets. This is due to the unpaired
electron moving to the different nuclei but spending a different length of time on each equivalent proton. In the methanol
radical example, the electron lingers the most on the CH2 protons but does move occasionally to the OH proton.

References
1. Bunce, N. "Introduction to the interpretation of electron spin resonance spectra of organic radicals." J. Chem. Educ., 1987,
64 (11), p 907
2. Griffiths, D. "Hyperfine Splitting in the ground state of Hydrogen." Am. J. Phys., 1982, 50 (8), p 698
3. Gasiorowicz, Stephen. Quantum Physics. New York: Wiley, 1974.2

Problems
1. What is the number of peaks of a benzene radical in EPR due to hyperfine coupling and what are their relative intensities?
2. What is the number of peaks for CH2(OCH3), a methoxymethyl radical in EPR due to hyperfine coupling?
3. Why are s orbitals only considered in Fermi contact interactions?
Answers:
1. 7 lines, 1:6:15:20:15:6:1 (benzene has 6 equivalent protons so the number of peaks is M+1 = (6+1) = 7 peaks, intensities
come from pascal's triangle)
2. 12 lines (has two different equivalent nuclei, one with two protons and one with 3 so (M1+1)(M2+1) = (2+1)(3+1) = 12
peaks)
3. p, d, and f orbitals have nodes at the nucleus and do not exhibit Fermi contact interactions

Contributors and Attributions


Stephanie Gray

7/3/2021 4 https://chem.libretexts.org/@go/page/1795
CHAPTER OVERVIEW
NUCLEAR MAGNETIC RESONANCE
Nuclear Magnetic Resonance (NMR) Spectroscopy uses the electromagnetic radiation of radio
waves to probe the local electronic interactions of a nucleus. NMR is a non-destructive technique
and has found uses in fields of medicine, chemistry, and environmental science.

NMR: BACKGROUND PHYSICS AND MATHEMATICS


NMR: EXPERIMENTAL

2D NMR
Nuclear Magnetic Resonance in liquid medium or liquid NMR is a technique used for the
structural analysis of a number of chemical molecules. It determines the structural parameters of
organic compounds. The most often studied nuclei are 1H,13C, and 31P.

2D NMR: INDIRECT DETECTION


2D NMR BACKGROUND
2D NMR BASICS
2D NMR EXPERIMENTS
2D NMR INTRODUCTION
HETERONUCLEAR CORRELATIONS
HOMONUCLEAR CORRELATIONS
CARBON-13 NMR
This page describes what a C-13 NMR spectrum is and how it tells you useful things about the carbon atoms in organic molecules.

INTERPRETING C-13 NMR SPECTRA


This page takes an introductory look at how you can get useful information from a C-13 NMR spectrum.

DIFFUSION ORDERED SPECTROSCOPY (DOSY)


Diffusion Ordered SpectroscopY (DOSY) utilizes magnetic field gradients to investigate diffusion processes occurring in solid and
liquid samples.

MAGNETIC RESONANCE IMAGING


Magnetic resonance imaging is a widely used noninvasive medical imaging technique to visualize the inner part of human body. It
applied the basic principles of nuclear magnetic resonance (NMR) spectroscopy, which provides both chemical and physical
information of molecules.

NMR - INTERPRETATION
NMR interpretation plays a pivotal role in molecular identifications. As interpreting NMR spectra, the structure of an unknown
compound, as well as known structures, can be assigned by several factors such as chemical shift, spin multiplicity, coupling
constants, and integration.

INTEGRATION IN NMR
PASCAL’S TRIANGLE
The Pascal’s triangle is a graphical device used to predict the ratio of heights of lines in a split NMR peak.

NMR HARDWARE
NMR SPECTROSCOPY IN LAB: COMPLICATIONS
PULSE SEQUENCES
A pulse sequence is a succinct visual representation of the pulses and delays used in a certain NMR experiment. Depending on the
experiment there may be hundreds of pulses! This page is dedicated to understanding what a pulse sequence is and how to understand
the pictorial representations, as well as terms commonly used to describe parts of pulse sequences.

PHASE CYCLING AND COHERENCE TRANSFER PATHWAYS


In NMR simply throwing a string of pulses with random phases is a recipe for disaster. Phase cycling is a way to choose the proper
coherence pathway in order to acquire the NMR signal which has the interactions you are trying to acquire. Additionally, phase
cycling eliminates artifacts that can appear in the NMR spectrum which are not real NMR signals.

1 7/7/2021
SOLID STATE EXPERIMENTS
The solid state NMR community has developed hundreds of techniques to investigate different parts of the NMR Hamiltonian. This
section gives a brief, by no means comprehensive of a few techniques that are helpful in probing select interactions.

CAR-PURCELL-MEIBOOM-GILL (CPMG) ECHO TRAIN ACQUISITION


CPMG is a NMR technique commonly employed to enhance the S/N of an NMR spectrum

MAGIC ANGLE HOPPING (MAH)


SOLID STATE NMR EXPERIMENTAL SETUP
NMR: STRUCTURAL ASSIGNMENT
Assignment of structures is a central problem which NMR is well suit to address. Explains how both 13C NMR spectra and low and
high resolution proton NMR spectra can be used to help to work out the structures of organic compounds.

(N+1) RULE
BACKGROUND TO C-13 NMR
DETERMINE STRUCTURE WITH COMBINED SPECTRA
HIGH RESOLUTION PROTON NMR SPECTRA
This page describes how you interpret simple high resolution nuclear magnetic resonance (NMR) spectra. It assumes that you have
already read the background page on NMR so that you understand what an NMR spectrum looks like and the use of the term
"chemical shift". It also assumes that you know how to interpret simple low resolution spectra.

INTEGRATION IN PROTON NMR


INTERPRETING C-13 NMR SPECTRA
INTRODUCTION TO PROTON NMR
LOW RESOLUTION PROTON NMR SPECTRA
MORE ABOUT ELECTRONICS
MULTIPLICITY IN PROTON NMR
NMR11. MORE ABOUT MULTIPLICITY
NMR14. MORE PRACTICE WITH NMR SPECTROSCOPY
NMR2. CARBON-13 NMR
NMR3. SYMMETRY IN NMR
NMR4. 13C NMR AND GEOMETRY
NMR5. 13C NMR AND ELECTRONICS
NMR8. CHEMICAL SHIFT IN 1H NMR
NMR APPENDIX. USEFUL CHARTS FOR NMR IDENTIFICATION
NMR - THEORY
Nuclear magnetic resonance has been play an important role in the fields of physical techniques available to the chemist for more than
25 years. It is becoming a more and more useful method to probe the structure of molecules. The primary object of this module is to
understand the fundamental concepts of NMR. It is assumed that the reader already understands the quantum numbers associated with
electrons.

BLOCH EQUATIONS
LARMOR PRECESSION
When placed in a magnetic field, charged particles will precess about the magnetic field. In NMR, the charged nucleus, will then
exhibit precessional motion at a characterisitc frequency known as the Larmor Frequency. The Larmor fequency is specific to each
nucleus. The Larmor fequency is measured during the NMR experiment, as it is dependent on the magnetic field that the nucleus
experineces.

NMR INTERACTIONS

CHEMICAL SHIFT (SHIELDING)


DIPOLAR COUPLING
J-COUPLING (SCALAR)
This page is dedicated to understanding the phenomena associated with J-couplings

PASCAL’S TRIANGLE CONSTRUCTION


The Pascal’s triangle is a graphical device used to predict the ratio of heights of lines in a split NMR peak.

QUADRUPOLAR COUPLING
QUANTUM MECHANIC TREATMENT

2 7/7/2021
RELAXATION
Relaxation in NMR is a fundamental concept which describes the coherence loss of the magnetization in the x-y plane and the
recovery of relaxation along the z-axis. There are many factors that contribute to the relaxation processes. This page will examine the
different types of relaxation at a basic level.

NMR: KINETICS
NUCLEAR OVERHAUSER EFFECT
SOLOMON EQUATIONS
SPIN-SPIN RELAXATION
SPIN LATTICE RELAXATION
ROTATIONS AND IRREDUCIBLE TENSOR OPERATORS
Irreducible Tensor Operators are extremely valuable to help reduce complex mathematical problems commonly found in NMR. This
page will be devoted to deriving the tensors and showing how to use the tensors in calculations.

NMR: INTRODUCTION
Nuclear Magnetic Resonance (NMR) is a nuceli (Nuclear) specific spectroscopy that has far reaching applications throughout the
physical sciences and industry. NMR uses a large magnet (Magnetic) to probe the intrinsic spin properties of atomic nuclei. Like all
spectroscopies, NMR uses a component of electromagnetic radiation (radio frequency waves) to promote transitions between nuclear
energy levels (Resonance). Most chemists use NMR for structure determination of small molecules.

BACK MATTER
INDEX

3 7/7/2021
NMR: Background Physics and Mathematics
Page Under Construction!

Introduction
This section will be devoted to understanding basic math and physics that are commonly encountered in NMR theory. These
explanations are going to be very simple and concise and the reader is encouraged to look at the references and other wiki
pages for fuller discussions.

Spin Angular Momentum

When talking about the spin angular momentum, nucleus can be considered as a mass point moving on a circular path. While
the momentum of a mass point moving along the straight path can be defined as
p ⃗ = m v ⃗  (NMR.1)

(where p and v are vectors), angular velocity is used to describe the motion of nucleus.
⃗ 
L = r ⃗ × p ⃗  (NMR.2)

(where L is angular momentum and r is the radius of the circle path)


Since L and p are perpendicular to each other, so
o
L = rp sin 90 = rp (NMR.3)

p = mv (NMR.4)

L = rmv (NMR.5)

The direction of L is determined by right-hand rule, so it is perpendicular to the circle plane.

Spherical Coordinates
For more advanced concepts in NMR, a spherical basis set is easier to use. Lets first consider a vector in Cartesian space
which can be described by

A = Ax ex + Ay ey Az ez (NMR.6)

where Ax,y,z is the projection of A onto the x y and z axes, and ex,y,z are the basis vectors.
The length of the vector is
−−−−−−−−−−−
2 2 2
|A| = √ Ax + Ay + Az (NMR.7)

Now switching to a spherical basis set A becomes

6/25/2021 NMR.1 https://chem.libretexts.org/@go/page/1798


+1 0 −1
A =A e+1 + A e0 + A e−1 = A+1 e + +1 + A0 e + 0 + A−1 e + −1 (NMR.8)

where the relation of the spherical basis set to the Cartesian basis set is

∓1
1
e±1 = −e =∓ – (ex ± i ey ) (NMR.9)
√2

0
e0 = e = ez (NMR.10)

∓1
1
A±1 = −A =∓ – (Ax ± i Ay ) (NMR.11)
√2

0
A0 = A = Az (NMR.12)

p p
Typically, Ap and A are called the contravariant and covariant components, while ep and e are the contravariant and covariant
basis vectors, respectively.

Diagonalization of a Matrix
Euler Angles
Euler Angles are a set of 3 angles that transform reference frames. These are commonly employed in NMR to switch between
reference frames. For example one set of Euler angles takes you from the laboratory frame to the rotating frame. A second set
of Euler angle can take you from the laboratory from the the CSA tensor frame. Below I've shown the ranges and how the
rotations work. More information may be found in Rotations section.
insert figure
γ and α range the full 2π radians while β ranges π radians

Unitary Evolution
According the 4th postulate of quantum mechanics "The evolution of a closed system is unitary (reversible). The evolution
given by the time-dependent Schrodinger equation
d|ψ >
i barh = H |ψ > (NMR.13)
dt

where H is the Hamiltonian of the system and h


¯
is the reduced Planck constant."
We can then express the evolution of a state using a unitary operator, known as the propagator
^
ψ(x, t) = U (t)ψ(x, 0) (NMR.14)

where the adjoint of the propagator is equal to the unity operator. We can show that this is equivalent to the time dependent
shcrodinger equation.
^
U ψ(x, 0)
¯ ^
ih = H U ψ(x, 0) (NMR.15)
dt

Since this equation must hold for any wave function then it also msut hold for ψ(x, 0) as well giving
^
dU
¯ ^
ih = HU (NMR.16)
dt

The only way this equation is solved is if the Hamiltonian is time independent and
−iH t

^ (t) = e
U
¯
h (NMR.17)

Fourier Transform
Mathmatical formula that takes you between frequency space and time space

Cross Product of 2 Matricies


Consider 2 2x2 matricies of the form

6/25/2021 NMR.2 https://chem.libretexts.org/@go/page/1798


a11 a12
A =[ ] (NMR.18)
a21 a22

b11 b12
B =[ ] (NMR.19)
b21 b22

the cross product is then defines as

a11 b11 a11 b12 a12 b11 a12 b12


⎡ ⎤

⎢ a11 b21 a11 b22 a12 b21 a12 b22 ⎥


A×B = ⎢ ⎥ (NMR.20)
⎢a ⎥
21 b11 a21 b12 a22 b11 a22 b 12

⎣ ⎦
a21 b21 a21 b22 a22 b21 a22 b22

Lorentz Forces
Raising and Lowering Operators
Reference Frames
First Order Differential equations

Laplace Equation
References
Outside Links
Problems
Contributors and Attributions
Name #1 here (if anonymous, you can avoid this) with university affiliation

6/25/2021 NMR.3 https://chem.libretexts.org/@go/page/1798


NMR: Experimental

Topic hierarchy

2D NMR
Nuclear Magnetic Resonance in liquid medium or liquid NMR is a technique used for the structural analysis of a number
of chemical molecules. It determines the structural parameters of organic compounds. The most often studied nuclei are
1H,13C, and 31P.

2D NMR: Indirect Detection


2D NMR Background
2D NMR Basics
2D NMR Experiments
2D NMR Introduction

Carbon-13 NMR
This page describes what a C-13 NMR spectrum is and how it tells you useful things about the carbon atoms in organic
molecules.

Interpreting C-13 NMR Spectra

Diffusion Ordered Spectroscopy (DOSY)


Diffusion Ordered SpectroscopY (DOSY) utilizes magnetic field gradients to investigate diffusion processes occurring in
solid and liquid samples.

Magnetic Resonance Imaging


Magnetic resonance imaging is a widely used noninvasive medical imaging technique to visualize the inner part of human
body. It applied the basic principles of nuclear magnetic resonance (NMR) spectroscopy, which provides both chemical
and physical information of molecules.

NMR - Interpretation
NMR interpretation plays a pivotal role in molecular identifications. As interpreting NMR spectra, the structure of an
unknown compound, as well as known structures, can be assigned by several factors such as chemical shift, spin
multiplicity, coupling constants, and integration.

Integration in NMR

7/7/2021 NMR.1 https://chem.libretexts.org/@go/page/1799


Pascal’s Triangle

NMR Hardware

NMR Spectroscopy in Lab: Complications

Pulse Sequences
A pulse sequence is a succinct visual representation of the pulses and delays used in a certain NMR experiment.
Depending on the experiment there may be hundreds of pulses! This page is dedicated to understanding what a pulse
sequence is and how to understand the pictorial representations, as well as terms commonly used to describe parts of
pulse sequences.

Phase Cycling and Coherence Transfer Pathways

Solid State Experiments


The solid state NMR community has developed hundreds of techniques to investigate different parts of the NMR
Hamiltonian. This section gives a brief, by no means comprehensive of a few techniques that are helpful in probing select
interactions.

Car-Purcell-Meiboom-Gill (CPMG) Echo Train Acquisition


Magic Angle Hopping (MAH)

Solid State NMR Experimental Setup

7/7/2021 NMR.2 https://chem.libretexts.org/@go/page/1799


2D NMR

6/21/2021 1 https://chem.libretexts.org/@go/page/1800
2D NMR: Indirect Detection

7/7/2021 2D NMR.1 https://chem.libretexts.org/@go/page/1801


2D NMR Background

6/29/2021 1 https://chem.libretexts.org/@go/page/1802
2D NMR Basics

6/18/2021 1 https://chem.libretexts.org/@go/page/1803
2D NMR Experiments

7/7/2021 1 https://chem.libretexts.org/@go/page/1804
2D NMR Introduction
Some general principles and techniques used in two-dimensional NMR are discussed. Applications covered are mostly
concerned with protein NMR, but additional 2D techniques and applications can be found in the references section.

Introduction
A two dimensional variation of NMR was first proposed by Jean Jeener in 1971; since then, scientists such as Richard Ernst
have applied the concept to develop the many techniques of 2D NMR. Although traditional, one-dimensional NMR is
sufficient to observe distinct peaks for the various funtional groups of small molecules, for larger, more complex molecules,
many overlapping resonances can make interpretation of an NMR spectrum difficult. Two-dimensional NMR, however, allows
one to circumvent this challenge by adding additional experimental variables and thus introducing a second dimension to the
resulting spectrum, providing data that is easier to interpret and often more informative.

Basics of 2D NMR
Experimental Set-up
In traditional 1D Fourier transform NMR, a sample under a magnetic field is hit with a series of RF pulses, as seen in the pulse
sequence below, and the Fourier transform of the outgoing signal results in a 1D spectra as a function chemical shift.

Figure 1: Pulse sequence of a typical FT-NMR experiment


A 2D NMR experiment, however, adds an additional dimension to the spectra by varying the length of time (τ )) the system is
allowed to evolve following the first pulse. The result is an outgoing signal f (τ , t2), which, when Fourier transformed, gives a
2D spectrum of F (ω1, ω2).
The use of two-dimensional NMR allows the researcher to better resolve signals which would normally overlap in 1D NMR.
Depending on the size of your molecule, different variations or combinations of 2D and multidimensional NMR experiments
are utilized.

The Spin Hamiltonian


The spin of a given nuclei during any NMR experiment is governmed by the spin Hamiltonian. If long-range spin interactions
are ignored, the spin Hamilitonian for a one-spin system is given the equation
0
^ ^ ^
H =H + H RF (1)

The magnetic field along the z-axis, shielding, and J-coupling with nearby nuclei are all constant and are accounted for in H0.
HRF is the induced magnetic field resulting from an RF pulse. For a system where two spins are coupled, the H0 is
^ 0 ^ 0 ^ ^ ^
H = ω I 1z + ω I 2z + 2π J12 I 1 I 2 (2)
1 2

Where ω is the Larmor frequency, I is the net magnetization vector of the given nucleus or nuclei, and J is the observed J
coupling between nuclei. ω is directly related to the chemical shift (δ ) by the equation
0 0 iso
ω = −γj B (1 + δ ) (3)
j j

Where γ is the gyromagnetic ratio of the given isotope. If nuclei 1 and 2 are of the same element and isotope, the system is
referred to as homonuclear. If they are different, it is a heteronuclear spin system.

Correlation Spectroscopy (COSY)

6/9/2021 1 https://chem.libretexts.org/@go/page/1805
The most basic form of 2D NMR is the 2D COSY (pulse sequence shown below) experiment, a homonuclear experiment with
a pulse sequence similar to the procedure dicussed above. It consists of a 90o RF pusle followed by an evolution time and an
additional 90o pulse. The resulting oscillating magnetization (symbolized by decaying the sinusoidal curve) is then acquired
during t2.

Figure 2: Pulse sequence of a 2D COSY experiment


The analysis of the acquired spectrum is discussed below, making it useful for determining the coupling between nuclei that
are connected through one to three bond lengths. However, in macromolecules such as proteins, coupling through bonds alone
is not sufficient to obtain substantial structural information. For this reason, the Nuclear Overhauser Effect (NOE) is often
used in protein NMR to obtain information on the distance between nuclei through space rather than through bonds.

Nuclear Overhauser Effect (NOE)


Thus far, only the coupling of nuclei through bonds has been considered. In bond coupling, the magnetization of nuclei affect
those closely bound to them through the electrons that make up those bonds; however, coupling directly between nuclei that
are in close spatial proximity to each other also occurs. This is called the Nuclear Overhauser Effect, and it arises when the
spin relaxation of nuclei A is felt by nearby nuclei B, stimulating a corresponding change in magnetization in B. In a typical
NMR spectrum, the interference of electrons makes this coupling undetectable. However, a sample can be decoupled to
“neutralize” the bond coupling through electrons, allowing the space coupling of the NOE to be detected. This is called
NOESY (NOE correlation spectroscopy) and is another type of homonuclear NMR.
The pulse sequence for a NOESY NMR experiment is depicted below.

Figure 3: Pulse sequence of NOESY experiment


o
Like COSY, the first step is a 90 pulse followed by a variable evolution time. Unlike COSY, however, pulse two actually
consists of two 90 degree pulses separated by a short delay. The first pulse converts the bulk magnetization from the transverse
plane to the z-plane, eliminating the effect of electron-aided bond coupling. Then, during the τ m, there is cross relaxation
between spatially adjacent nuclei. Finally, the last 90 degree pulse converts the space coupling of nuclei into an observable
transverse magnetization, which can be detected during t2.

2D NMR Spectra
As discussed earlier, by performing multiple one dimensional experiments at varying lengths (τ ) of the evolution period and
performing a Fourier transformation on the signal which converts f (τ , t2) to F(ω1, ω2), a two dimensional spectrum can be
formed into a 3D contour map.
A more useful representation of 2D data is called a correlation map. The correlation map of the steroid progesterone is shown
below.

6/9/2021 2 https://chem.libretexts.org/@go/page/1805
Figure 4: 2D COSY spectrum of progesterone
In this representation, the x- and y-axes correspond to the frequencies resulting from the Fourier transforms, and the intensity
of shade at each frequency coordinate indicates the peak intensity. Two types of peaks are observed in a homonuclear
correlation map—diagonal peaks and cross peaks. Diagonal peaks are found along the diagonal of the map where the x- and y-
axes have equal frequency values and simply correspond to the absorptions from a one-dimensional NMR experiment.
Because heteronuclear NMR does not involve the same isotope, diagonal peaks are not observed. Cross peaks, on the other
hand, give information on the coupling of two nuclei and are seen in both homo- and heteronuclear spectra.

Applications in Protein NMR


As previously mentioned, the major advantage of 2D NMR over 1D NMR is the ability to distinguish between the overlapping
signals that exist in larger molecules. Heteronuclear two-dimensional NMR is especially important in biological chemistry in
the elucidation of the three-dimensional structure of proteins.

Heteronuclear Single Quantum Coherence (HSQC)


A protein is make-up of a series of amino acid monomers. Although there are 19 different amino acids each with a distinct side
chain, the protein backbone is an invariable pattern of NH-C-CO as shown in Figure 5.

Figure 5: Structure of a amino acid chain with NMR active nuclei


When synthesized under the right conditions, a heavy atom protein can be produced which constains NMR active nuclei;
however, a 15N nucleus has a very low gyromagnetic ratio. According to the Hamiltonian operators discussed above, it will

6/9/2021 3 https://chem.libretexts.org/@go/page/1805
give a very weak signal in traditional 2D NMR. Fortunately, the nucleus can be detected indirectly by transferring polarization
through a 1H nucleus. This method is used in HSQC NMR.
In protein NMR, each HSQC experiment has three steps:
1. An INEPT (insensitive nuclei enhanced by polarization transfer) transfers the polarization of a 1H nuclei to the neighboring
15
N (see figure below)
2. The polarization is transferred back to the 1H nuclei
3. Signal from the 1H nuclei is recorded
The pulse sequence for a typical HSQC experiment is detailed below.

Figure 6: Pulse sequence of a HSQC experiment involving 1H and 15N nuclei


A 90o1H RF pulse creates a transverse polarization in 1H nuclei. Following the pulse, the nuclei are allowed to evolve for a
1/(4J) time period, which is the longitudinal relaxation time. Next, a 180o 1H and 15N pulse are used at the same time. During
the subsequent relaxation time, the 1H nuclei develop a polarization that is antiphase to 15N. Finally, a 90o 1H and 15N pulse,
again simultaneous, enacts the INEPT transfer of antiphase magnetization from the 1H nucleus to the 15N nucleus. Following
the INEPT transfer, the 15N nuclei are allowed to evolve during τ before a reverse INEPT transfer moves the 15N polarization
back to 1H and a 15N decoupled signal is recorded.
An example HSCQ spectrum from ubiquitin is shown below.

Figure 7: 1H15N HSQC spectrum of ubiquitin


Notice the greater clarity of spectra of the HSQC vs the COSY experiment. This is a strong advantage of heteronuclear NMR.
In this diagram, each peak corresponds to a cross peak, showing coupling between sets of 1H and 15N nuclei. Each peak
represents the 15N—1H of a unique amino acid along the backbone of the amino acid.
The 2D HSQC experiment which relates 1H and 15N is just the start on the long, complicated road of protein structural
characterization using 2D and multidimensional NMR techniques. The next experiment is typically a 3D NMR technique in
which coupling with 13C is also including in the spectra—which will give information on which NH peaks are associated with
which type of amino acid residue—followed by a NOE experiment—which gives spatial distances between nuclei. These
multidimensional techniques are outside the scope of this module; however, the principles which were applied in the

6/9/2021 4 https://chem.libretexts.org/@go/page/1805
construction and analysis of the 2D spectra can be carried over to 3D and 4D NMR. The only difference, naturally, is the
additional dimensions. By the time multiple experiments have been carried out, information regarding proximity of nuclei, the
types of amino acid associated with each nuclei, and secondary structure has been amassed, and by carefully piecing all the
data together, a 3D structure of a given protein can be constructed.

Other Applications of 2D NMR


2D NMR has many more applications beyond protein NMR, including characterization of pharmaceuticals, temperature
dependence of carbohydrate conformations, and metabolomics, to just name a few. For more information on these applications
and the 2D NMR techniques that are used in them, please see the “Further Reading” section.

References
1. Aue, W., E. Bartholdi, R.R. Ernst, Two‐dimensional spectroscopy. Application to nuclear magnetic resonance. The Journal
of Chemical Physics, 1976. 64: p. 2229.
2. Gomathi, L., Elucidation of secondary structures of peptides using high resolution NMR. Current Science, 1996. 71(7): p.
553.
3. Levitt, M.H., Spin Dynamics: Basics of Nuclear Magnetic Resonance. 2008: Wiley
4. Jacobsen, N.A., NMR Spectroscopy Explained: Simplified Theory, Applications and Examples for Organic Chemistry and
Structural Biology. 2007: Wiley. 668.
5. Ames, J.B., Hamasaki, N., Molchanova, T., Structure and calcium-binding studies of a recoverin mutant (E85Q) in an
allosteric intermediate state. Biochemistry, 2002. 41(18): p. 5776. DOI: 10.1021/bi012153k
6. Clore, G.G., Angela M., Determining structures of large proteins and protein complexes by NMR. Biological Magnetic
Resonance, 1998. 16.

Outside Links
http://en.Wikipedia.org/wiki/Two-dim...e_spectroscopy
http://triton.iqfr.csic.es/guide/eNM...nv/hsqc2d.html
www.bioc.aecom.yu.edu/labs/gi...rator_2012.pdf
www.chem.queensu.ca/facilitie.../hmqc.htm#hsqc

Problems
1. What is the effect of the 90° pulse on the bulk magnetization of a sample?
2. If a 90° pulse for a give sample is 4 fs long, how long is a 180° pulse on the same sample?
3. The magnetic effect of which type of particle must be removed from an NMR experiment in order to observe an NOE?
Solutions: 1) a 90o pulse moves the bulk magnetization into the transverse plane 2) 8 fs 3) electrons

Further Reading
1. Ludwig, C., Viant, Mark R., Two-dimensional J-resolved NMR spectroscopy: review of a key methodology in the
metabolomics toobox. Phytochemical Analysis, 2010. 21(1): p. 22-32. DOI: 10.1002/pca.1186
2. Brown, S.P., Applications of high-resolution 1H solid-state NMR. Solida State Nuclear Magnetic Resonance, 2012. 41: p.
1-27. DOI: 10.1016/j.ssnmr.2011.11.006
3. Shrot, Y.F., Lucio, Ghost-peak suppression in ultrafast two-dimensional NMR. Journal of Magnetic Resonance, 2003.
164(2): p. 351-357. DOI: 10.1016/S1090-7807(03)00177-0

Contributors and Attributions


MM UC- Davis

6/9/2021 5 https://chem.libretexts.org/@go/page/1805
Heteronuclear Correlations

7/4/2021 1 https://chem.libretexts.org/@go/page/1806
Homonuclear Correlations

6/21/2021 1 https://chem.libretexts.org/@go/page/1807
Carbon-13 NMR
This page describes what a C-13 NMR spectrum is and how it tells you useful things about the carbon atoms in organic
molecules.

Carbon-13 Nuclei as Little Magnets


About 1% of all carbon atoms are the C-13 isotope; the rest (apart from tiny amounts of the radioactive C-14) is C-12. C-13
NMR relies on the magnetic properties of the C-13 nuclei. Carbon-13 nuclei fall into a class known as "spin ½" nuclei for
reasons which do not really need to concern us at the introductory level this page is aimed at (UK A level and its equivalents).
The effect of this is that a C-13 nucleus can behave as a little magnet. C-12 nuclei do not have this property.
If you have a compass needle, it normally lines up with the Earth's magnetic field with the north-seeking end pointing north.
Provided it is not sealed in some sort of container, you could twist the needle around with your fingers so that it pointed south -
lining it up opposed to the Earth's magnetic field. It is very unstable opposed to the Earth's field, and as soon as you let it go
again, it will flip back to its more stable state.

Because a C-13 nucleus behaves like a little magnet, it means that it can also be aligned with an external magnetic field or
opposed to it. Again, the alignment where it is opposed to the field is less stable (at a higher energy). It is possible to make it
flip from the more stable alignment to the less stable one by supplying exactly the right amount of energy.

The energy needed to make this flip depends on the strength of the external magnetic field used, but is usually in the range of
energies found in radio waves - at frequencies of about 25 - 100 MHz. If you have also looked at proton-NMR, the frequency
is about a quarter of that used to flip a hydrogen nucleus for a given magnetic field strength.
It's possible to detect this interaction between the radio waves of just the right frequency and the carbon-13 nucleus as it flips
from one orientation to the other as a peak on a graph. This flipping of the carbon-13 nucleus from one magnetic alignment to
the other by the radio waves is known as the resonance condition.

The Importance of the Carbon's Environment


What we've said so far would apply to an isolated carbon-13 nucleus, but real carbon atoms in real bonds have other things
around them - especially electrons. The effect of the electrons is to cut down the size of the external magnetic field felt by the
carbon-13 nucleus.

Suppose you were using a radio frequency of 25 MHz, and you adjusted the size of the magnetic field so that an isolated
carbon-13 atom was in the resonance condition. If you replaced the isolated carbon with the more realistic case of it being

Jim Clark 6/23/2021 1 https://chem.libretexts.org/@go/page/1808


surrounded by bonding electrons, it wouldn't be feeling the full effect of the external field any more and so would stop
resonating (flipping from one magnetic alignment to the other). The resonance condition depends on having exactly the right
combination of external magnetic field and radio frequency.
How would you bring it back into the resonance condition again? You would have to increase the external magnetic field
slightly to compensate for the shielding effect of the electrons. Now suppose that you attached the carbon to something more
electronegative. The electrons in the bond would be further away from the carbon nucleus, and so would have less of a
lowering effect on the magnetic field around the carbon nucleus.

Electronegativity is a measure of the ability of an atom to attract a bonding pair of electrons. If you are not happy about
electronegativity, you could follow this link at some point in the future, but it probably is not worth doing it now!

The external magnetic field needed to bring the carbon into resonance will be smaller if it is attached to a more electronegative
element, because the C-13 nucleus feels more of the field. Even small differences in the electronegativities of the attached
atoms will make a difference to the magnetic field needed to achieve resonance.

Summary
For a given radio frequency (say, 25 MHz) each carbon-13 atom will need a slightly different magnetic field applied to it to
bring it into the resonance condition depending on what exactly it is attached to - in other words the magnetic field needed is a
useful guide to the carbon atom's environment in the molecule.

The C-13 NMR Spectrum for Ethanol


This is a simple example of a C-13 NMR spectrum. do not worry about the scale for now - we'll look at that in a minute.

The NMR spectra on this page have been produced from graphs taken from the Spectral Data Base System for Organic
Compounds (SDBS) at the National Institute of Materials and Chemical Research in Japan.
It is possible that small errors may have been introduced during the process of converting them for use on this site, but
these won't affect the argument in any way.

There are two peaks because there are two different environments for the carbons. The carbon in the CH3 group is attached to
3 hydrogens and a carbon. The carbon in the CH2 group is attached to 2 hydrogens, a carbon and an oxygen. The two lines are
in different places in the NMR spectrum because they need different external magnetic fields to bring them in to resonance at a
particular radio frequency.

The C-13 NMR Spectrum for a more complicated compound


This is the C-13 NMR spectrum for 1-methylethyl propanoate (also known as isopropyl propanoate or isopropyl propionate).

Jim Clark 6/23/2021 2 https://chem.libretexts.org/@go/page/1808


This time there are 5 lines in the spectrum. That means that there must be 5 different environments for the carbon atoms in the
compound. Is that reasonable from the structure?

Well - if you count the carbon atoms, there are 6 of them. So why only 5 lines? In this case, two of the carbons are in exactly
the same environment. They are attached to exactly the same things. Look at the two CH3 groups on the right-hand side of the
molecule.
You might reasonably ask why the carbon in the CH3 on the left is not also in the same environment. Just like the ones on the
right, the carbon is attached to 3 hydrogens and another carbon. But the similarity is not exact - you have to chase the
similarity along the rest of the molecule as well to be sure.
The carbon in the left-hand CH3 group is attached to a carbon atom which in turn is attached to a carbon with two oxygens on
it - and so on down the molecule. That's not exactly the same environment as the carbons in the right-hand CH3 groups. They
are attached to a carbon which is attached to a single oxygen - and so on down the molecule. We'll look at this spectrum again
in detail on the next page - and look at some more similar examples as well. This all gets easier the more examples you look
at.
For now, all you need to realize is that each line in a C-13 NMR spectrum recognizes a carbon atom in one particular
environment in the compound. If two (or more) carbon atoms in a compound have exactly the same environment, they will be
represented by a single line.

If you are fairly wide-awake, you might wonder why all this works, since only about 1% of carbon atoms are C-13. These
are the only ones picked up by this form of NMR. If you had a single molecule of ethanol, then the chances are only about
1 in 50 of there being one C-13 atom in it, and only about 1 in 10,000 of both being C-13.
But you have got to remember that you will be working with a sample containing huge numbers of molecules. The
instrument can pick up the magnetic effect of the C-13 nuclei in the carbon of the CH3 group and the carbon of the CH2
group even if they are in separate molecules. There's no need for them to be in the same one.

The Need for a Standard: TMS


Before we can explain what the horizontal scale means, we need to explain the fact that it has a zero point - at the right-hand
end of the scale. The zero is where you would find a peak due to the carbon-13 atoms in tetramethylsilane (TMS). Everything
else is compared with this.

You will find that some NMR spectra show the peak due to TMS (at zero), and others leave it out. Essentially, if you have to
analyse a spectrum which has a peak at zero, you can ignore it because that's the TMS peak. TMS is chosen as the standard for
several reasons. The most important are:
It has 4 carbon atoms all of which are in exactly the same environment. They are joined to exactly the same things in
exactly the same way. That produces a single peak, but it's also a strong peak (because there are lots of carbon atoms all

Jim Clark 6/23/2021 3 https://chem.libretexts.org/@go/page/1808


doing the same thing).
The electrons in the C-Si bonds are closer to the carbons in this compound than in almost any other one. That means that
these carbon nuclei are the most shielded from the external magnetic field, and so you would have to increase the magnetic
field by the greatest amount to bring the carbons back into resonance.
The net effect of this is that TMS produces a peak on the spectrum at the extreme right-hand side. Almost everything else
produces peaks to the left of it.

The Chemical Shift


The horizontal scale is shown as δ (ppm). δ is called the chemical shift and is measured in parts per million - ppm. A peak at a
chemical shift of, say, 60 means that the carbon atoms which caused that peak need a magnetic field 60 millionths less than the
field needed by TMS to produce resonance. A peak at a chemical shift of 60 is said to be downfield of TMS. The further to the
left a peak is, the more downfield it is.

If you are familiar with proton-NMR, you will notice that the chemical shifts for C-13 NMR are much bigger than for
proton-NMR. In C-13 NMR, they range up to about 200 ppm. In proton-NMR they only go up to about 12 ppm. You do
not need to worry about the reasons for this at this level.

Solvents for NMR Spectroscopy


NMR spectra are usually measured using solutions of the substance being investigated. A commonly used solvent is CDCl3.
This is a trichloromethane (chloroform) molecule in which the hydrogen has been replaced by its isotope, deuterium. CDCl3 is
also commonly used as the solvent in proton-NMR because it does not have any ordinary hydrogen nuclei (protons) which
would give a line in a proton-NMR spectrum. It does, of course, have a carbon atom - so why doesn't it give a potentially
confusing line in a C-13 NMR spectrum? In fact it does give a line, but the line has an easily recognisable chemical shift and
so can be removed from the final spectrum. All of the spectra from the SDBS have this line removed to avoid any confusion.

Contributors and Attributions


Jim Clark (Chemguide.co.uk)

Jim Clark 6/23/2021 4 https://chem.libretexts.org/@go/page/1808


Interpreting C-13 NMR Spectra
This page takes an introductory look at how you can get useful information from a C-13 NMR spectrum.

Introduction
Taking a close look at three 13C NMR spectra below. The 13C NMR spectrum for ethanol

The NMR spectra on this page have been produced from graphs taken from the Spectral Data Base System for Organic
Compounds (SDBS) at the National Institute of Materials and Chemical Research in Japan.

Remember that each peak identifies a carbon atom in a different environment within the molecule. In this case there are two
peaks because there are two different environments for the carbons. The carbon in the CH3 group is attached to 3 hydrogens
and a carbon. The carbon in the CH2group is attached to 2 hydrogens, a carbon and an oxygen. So which peak is which?
You might remember from the introductory page that the external magnetic field experienced by the carbon nuclei is affected
by the electronegativity of the atoms attached to them. The effect of this is that the chemical shift of the carbon increases if
you attach an atom like oxygen to it. That means that the peak at about 60 (the larger chemical shift) is due to the CH2 group
because it has a more electronegative atom attached.

In principle, you should be able to work out the fact that the carbon attached to the oxygen will have the larger chemical
shift. In practice, you always work from tables of chemical shift values for different groups (see below).
What if you needed to work it out? The electronegative oxygen pulls electrons away from the carbon nucleus leaving it
more exposed to any external magnetic field. That means that you will need a smaller external magnetic field to bring the
nucleus into the resonance condition than if it was attached to less electronegative things. The smaller the magnetic field
needed, the higher the chemical shift.

A table of typical chemical shifts in C-13 NMR spectra


carbon environment chemical shift (ppm)

C=O (in ketones) 205 - 220


C=O (in aldehydes) 190 - 200
C=O (in acids and esters) 170 - 185
C in aromatic rings 125 - 150
C=C (in alkenes) 115 - 140
RCH2OH 50 - 65
RCH2Cl 40 - 45
RCH2NH2 37 - 45
R3CH 25 - 35
CH3CO- 20 - 30
R2CH2 16 - 25

Jim Clark 6/21/2021 1 https://chem.libretexts.org/@go/page/1809


RCH3 10 - 15

In the table, the "R" groups will not necessarily be simple alkyl groups. In each case there will be a carbon atom attached to
the one shown in red, but there may well be other things substituted into the "R" group.
If a substituent is very close to the carbon in question, and very electronegative, that might affect the values given in the table
slightly. For example, ethanol has a peak at about 60 because of the CH2OH group. No problem! It also has a peak due to the
RCH3 group. The "R" group this time is CH2OH. The electron pulling effect of the oxygen atom increases the chemical shift
slightly from the one shown in the table to a value of about 18. A simplification of the table:
carbon environment chemical shift (ppm)

C-C 0 - 50
C-O 50 - 100
C=C 100 - 150
C=O 150 - 200

This may, of course, change and other syllabuses might want something similar. The only way to find out is to check your
syllabus, and recent question papers to see whether you are given tables of chemical shifts or not.

Example 1 : 3-buten-2-one
The 13C NMR spectrum for but-3-en-2-one. This is also known as 3-buten-2-one (among many other things!)

Here is the structure for the compound:

You can pick out all the peaks in this compound using the simplified table above.
The peak at just under 200 ppm is due to a carbon-oxygen double bond. The two peaks at 137 ppm and 129 ppm are
due to the carbons at either end of the carbon-carbon double bond. And the peak at 26 is the methyl group which, of
course, is joined to the rest of the molecule by a carbon-carbon single bond. If you want to use the more accurate
table, you have to put a bit more thought into it - and, in particular, worry about the values which do not always
exactly match those in the table!
The carbon-oxygen double bond in the peak for the ketone group has a slightly lower value than the table suggests for
a ketone. There is an interaction between the carbon-oxygen and carbon-carbon double bonds in the molecule which
affects the value slightly. This isn't something which we need to look at in detail for the purposes of this topic.
You must be prepared to find small discrepancies of this sort in more complicated molecules - but do not worry about
this for exam purposes at this level. Your examiners should give you shift values which exactly match the compound
you are given.
The two peaks for the carbons in the carbon-carbon double bond are exactly where they would be expected to be.
Notice that they aren't in exactly the same environment, and so do not have the same shift values. The one closer to
the carbon-oxygen double bond has the larger value.
And the methyl group on the end has exactly the sort of value you would expect for one attached to C=O. The table
gives a range of 20 - 30, and that's where it is.

Jim Clark 6/21/2021 2 https://chem.libretexts.org/@go/page/1809


One final important thing to notice. There are four carbons in the molecule and four peaks because they are all in different
environments. But they aren't all the same height. In C-13 NMR, you cannot draw any simple conclusions from the
heights of the various peaks.

Example 2 : C-13 NMR spectrum for 1-methylethyl propanoate


1-methylethyl propanoate is also known as isopropyl propanoate or isopropyl propionate.

Here is the structure for 1-methylethyl propanoate:

Two simple peaks


There are two very simple peaks in the spectrum which could be identified easily from the second table above.
The peak at 174 is due to a carbon in a carbon-oxygen double bond. (Looking at the more detailed table, this peak is
due to the carbon in a carbon-oxygen double bond in an acid or ester.)
The peak at 67 is due to a different carbon singly bonded to an oxygen. Those two peaks are therefore due to:

If you look back at the more detailed table of chemical shifts, you will find that a carbon singly bonded to an oxygen has
a range of 50 - 65. 67 is, of course, a little bit higher than that.
As before, you must expect these small differences. No table can account for all the fine differences in environment of a
carbon in a molecule. Different tables will quote slightly different ranges. At this level, you can just ignore that problem!
Before we go on to look at the other peaks, notice the heights of these two peaks we've been talking about. They are both
due to a single carbon atom in the molecule, and yet they have different heights. Again, you can't read any reliable
information directly from peak heights in these spectra.
The three right-hand peaks
From the simplified table, all you can say is that these are due to carbons attached to other carbon atoms by single bonds.
But because there are three peaks, the carbons must be in three different environments.
The easiest peak to sort out is the one at 28. If you look back at the table, that could well be a carbon attached to a carbon-
oxygen double bond. The table quotes the group as CH CO , but replacing one of the hydrogens by a simple CH3 group
3

will not make much difference to the shift value.


The right-hand peak is also fairly easy. This is the left-hand methyl group in the molecule. It is attached to an admittedly
complicated R group (the rest of the molecule). It is the bottom value given in the detailed table.

Jim Clark 6/21/2021 3 https://chem.libretexts.org/@go/page/1809


The tall peak at 22 must be due to the two methyl groups at the right-hand end of the molecule - because that's all that's
left. These combine to give a single peak because they are both in exactly the same environment.
If you are looking at the detailed table, you need to think very carefully which of the environments you should be looking
at. Without thinking, it is tempting to go for the R2CH2 with peaks in the 16 - 25 region. But you would be wrong! The
carbons we are interested in are the ones in the methyl group, not in the R groups. These carbons are again in the
environment: RCH3. The R is the rest of the molecule. The table says that these should have peaks in the range 10 - 15,
but our peak is a bit higher. This is because of the presence of the nearby oxygen atom. Its electronegativity is pulling
electrons away from the methyl groups - and, as we've seen above, this tends to increase the chemical shift slightly.

Working out Structures from C-13 NMR Spectra


So far, we have just been trying to see the relationship between carbons in particular environments in a molecule and the
spectrum produced. We've had all the information necessary. Now let's make it a little more difficult - but we'll work from
much easier examples! In each example, try to work it out for yourself before you read the explanation. How could you tell
from just a quick look at a C-13 NMR spectrum (and without worrying about chemical shifts) whether you had propanone or
propanal (assuming those were the only options)?

Because these are isomers, each has the same number of carbon atoms, but there is a difference between the environments of
the carbons which will make a big impact on the spectra.
In propanone, the two carbons in the methyl groups are in exactly the same environment, and so will produce only a single
peak. That means that the propanone spectrum will have only 2 peaks - one for the methyl groups and one for the carbon in the
C=O group. However, in propanal, all the carbons are in completely different environments, and the spectrum will have three
peaks.

Example 3 : C 4 H10 O

There are four alcohols with the molecular formula C 4 H10 O .

Which one produced the C-13 NMR spectrum below?

You can do this perfectly well without referring to chemical shift tables at all.
In the spectrum there are a total of three peaks - that means that there are only three different environments for the
carbons, despite there being four carbon atoms.
In A and B, there are four totally different environments. Both of these would produce four peaks.
In D, there are only two different environments - all the methyl groups are exactly equivalent. D would only produce two
peaks.

Jim Clark 6/21/2021 4 https://chem.libretexts.org/@go/page/1809


That leaves C. Two of the methyl groups are in exactly the same environment - attached to the rest of the molecule in
exactly the same way. They would only produce one peak. With the other two carbon atoms, that would make a total of
three. The alcohol is C.

Example 4 :
This follows on from Example 3, and also involves an isomer of C4 H10 O but which isn't an alcohol. Its C-13 NMR
spectrum is below. Work out what its structure is.

Because we do not know what sort of structure we are looking at, this time it would be a good idea to look at the shift
values. The approximations are perfectly good, and we will work from this table:
carbon environment chemical shift (ppm)

C-C 0 - 50
C-O 50 - 100
C=C 100 - 150
C=O 150 - 200

There is a peak for carbon(s) in a carbon-oxygen single bond and one for carbon(s) in a carbon-carbon single bond. That
would be consistent with C-C-O in the structure.
It is not an alcohol (you are told that in the question), and so there must be another carbon on the right-hand side of the
oxygen in the structure in the last paragraph. The molecular formula is C4H10O, and there are only two peaks. The only
solution to that is to have two identical ethyl groups either side of the oxygen. The compound is ethoxyethane (diethyl
ether), CH3CH2OCH2CH3.

Example 5
Using the simplified table of chemical shifts above, work out the structure of the compound with the following C-13
NMR spectrum. Its molecular formula is C H O .
4 6 2

Let's sort out what we've got.


There are four peaks and four carbons. No two carbons are in exactly the same environment.
The peak at just over 50 must be a carbon attached to an oxygen by a single bond.
The two peaks around 130 must be the two carbons at either end of a carbon-carbon double bond.
The peak at just less than 170 is the carbon in a carbon-oxygen double bond.

Jim Clark 6/21/2021 5 https://chem.libretexts.org/@go/page/1809


Putting this together is a matter of playing around with the structures until you have come up with something reasonable.
But you can't be sure that you have got the right structure using this simplified table. In this particular case, the spectrum
was for the compound:

If you refer back to the more accurate table of chemical shifts towards the top of the page, you will get some better
confirmation of this. The relatively low value of the carbon-oxygen double bond peak suggests an ester or acid rather than
an aldehyde or ketone.
It can't be an acid because there has to be a carbon attached to an oxygen by a single bond somewhere - apart from the
one in the -COOH group. We've already accounted for that carbon atom from the peak at about 170. If it was an acid, you
would already have used up both oxygen atoms in the structure in the -COOH group. Without this information, though,
you could probably come up with reasonable alternative structures. If you were working from the simplified table in an
exam, your examiners would have to allow any valid alternatives.

Contributors and Attributions


Jim Clark (Chemguide.co.uk)

Jim Clark 6/21/2021 6 https://chem.libretexts.org/@go/page/1809


Diffusion Ordered Spectroscopy (DOSY)
Diffusion Ordered SpectroscopY (DOSY) utilizes magnetic field gradients to investigate diffusion processes occurring in solid
and liquid samples.

Theory
Spin Diffusion
In the classic formation of a spin-echo (i.e. 90o-τ -180o-2τ ) the intensity of the echo will damp according to T2 relaxation
processes. However, in the presence of inhomogeneities in the magnetic field, if a spin drifts to a location in which the
magnetic field is different, the Larmor frequency changes and the spin will not refocus with the remaining spins. We can then
re-write the bloch equation to account for diffusion processes
dM Mx + My Mz − M0
2
= γM B − − + D∇ M (1)
dt T2 T1

Typically, the changes in the Larmor frequency have a negligible effect. However, application of a Gradient, G, will induce a
large change in larmor frequencies experienced at places in the sample.

Magnetic Field Gradient Echo


The principle behind the DOSY experiment is the formation of a magnetic field gradient echo. This is a spin-echo in which a
magnetic field gradient is applied during the spin evolution. Thus spins that diffuse during this time will not refocus in the spin
echo, as the application of the gradient will impart their own magnetic field.

Pulse Program
The sequence consists of a conventional Hahn echo (spin echo) sequence in which two gradient pulses are applied at equal
timings after the 90 and 180 degree pulses. The gradients are opposite in magnitude. The basic principle relies on the diffusion
of spins in the sample. Initially the 90 degree pulse tips the magnetization into the x-y plane, where spins begin to precess with
their characteristic Larmor frequencies. The application of a gradient sometime after the 90 degree pulse, encodes a spatial
component to the spin. That is the gradient is not uniform over the sample and therefore, the processional frequencies will
change. Next the application of the 180 flips the magnetization to refocus the spins. However, the spins, due to the application
of the gradient will not refocus. The application of a gradient at the same time (but with opposite direction) will refocus the
spins at a total time of 2 τ . If a spin diffuses to a different place in the sample, the refocusing will not occur, leading the
dampening of the echo intensity.

Below is the basic pulse sequence for a pulsed field gradient experiment.

6/20/2021 1 https://chem.libretexts.org/@go/page/55275
Processing
Processing the data is fairly straight forward. Apply a Fourier Transformation along the F2 dimension. The F1 dimension, in
which either the time or the gradient was incremented remains in the time domain. The echo intensity of a given peak is then
described by:
2 2 2
ln μ(ga , tc ) − ln μ(0, tc ) = −Cn γ Dδ ga tD (2)

where μ((g , t ) is the echo amplitude after the gradient application, ln m(0, t ) is the echo amplitude with no gradient
a c c

applied, Cn is a constant that depends on the particular pulse sequence used, γ is the gyromagnetic ratio, D is the diffusion
coefficient, δ is the width of the applied gradient, and tD is the diffusion time.

6/20/2021 2 https://chem.libretexts.org/@go/page/55275
Magnetic Resonance Imaging
Magnetic resonance imaging is a widely used noninvasive medical imaging technique to visualize the inner part of human
body. It applied the basic principles of nuclear magnetic resonance (NMR) spectroscopy, which provides both chemical and
physical information of molecules.

Introduction
Since a large fraction of composition of the human body is water and fat, the abundance of hydrogen atoms in fat and water
makes the human body approximately 63% hydrogen atoms. MRI uses a powerful magnetic field to align the nuclear
magnetization of hydrogen atoms in water in the body. When Radio frequency (RF) fields are added to systematically alter the
alignment of this magnetization, the hydrogen nuclei produce a rotating magnetic field detectable by the scanner. Magnetic
resonance imaging primarily images the NMR signals provided by the hydrogen nuclei inside the body. Those signals can be
detected by gradient magnetic fields and build up enough information to construct an image of the body, especially important
since it can display the different properties between normal tissues and tumor. The ability of depicting inner structure of
substances also makes MRI of great use in physics, chemistry and many other areas.
Compared to PET and CT, MRI has a rather good resolution and is harmless to patient, since it doesn’t require radioactive
injections. This young and still growing technique should have a promising future. In 2003, there were approximately 10,000
MRI units worldwide, and approximately 75 million MRI scans per year performed.[2] As the field of MRI continues to grow,
so do the opportunities in MRI.

History of NMR and MRI


Back in the 1930s, Isidor Rabi conducted research on the nature of the force binding protons to atomic nuclei and eventually
discovered that the spin orientation of nuclei in a magnetic field changes when radio frequency is added. This was the first
human investigation on the interaction of atomic nuclei, magnetic field and radio frequency. He was thus awarded the Nobel
Prize for Physics in the year of 1944 for this great work.
A decade later, in 1946, Purcel of Harvard University and F. Block of Stanford University independently discovered that
magnetic nuclei in a magnetic field absorb and re-emit electromagnetic radiation. Their discovery of nuclear magnetic
resonance and further expansion on this technique won them the Nobel Prize for Physics in 1952.
Immediate after the discovery of nuclear magnetic resonance phenomena, people started to apply this technique on
investigation on nuclei structure and properties, such as the measurement of nuclear magnetic moment, as well as other
practical researches. Chemists developed NMR spectroscopy according to the influences of molecular structure on the
surround magnetic field of Hydrogen. NMR spectroscopy becomes a prevailing technique for molecular structure
determination and it develops from one dimensional 1H spectroscopy to 13 carbon spectroscopy, two dimensional
spectroscopy and other advanced methods. In the 1990s, scientist even applied this technique on protein, which made the
accurate determination of molecular structure of liquid phase protein possible. Nowadays, nuclear magnetic resonance
technique is widely used in various areas in chemistry, biology and medical research, from molecular structure and
composition analysis to medical diagnose.
In 1971, R. Damadian found out that both the NMR longitudinal relaxation time (T1) and transverse relaxation time (T2) in
tumor are longer than normal tissue, which means tumor and normal tissue can be distinguished by NMR in vivo. This work
provided a magnificent chance of bringing NMR technique into medical research. Magnetic resonance imaging was first
demonstrated on small test tube samples by Paul Lauterbur in 1973 and the first cross-sectional image of a living mouse was
published in January 1974. In 1975 Richard Ernst proposed magnetic resonance imaging using phase and frequency encoding,
and the Fourier Transform, which is the basis of current MRI techniques. While Richard Ernst was rewarded for his
achievements in pulsed Fourier Transform NMR and MRI with the Nobel Prize in Chemistry in 1991, Paul C. Lauterbur of the
University of Illinois and Sir Peter Mansfield of the University of Nottingham were awarded the Nobel Prize in Medicine for
their discoveries concerning magnetic resonance imaging in 2003. To date, MRI is widely used for tumor detection and
diagnose.

Basic Principles

7/4/2021 1 https://chem.libretexts.org/@go/page/1811
Spin physics
The hydrogen atom possesses a nuclear spin of 1/2. When placed in an external magnetic field,the spin vector of hydrogen
atoms aligns itself with the magnetic field and the single energy state will split into two energy states. A particle in the lower
energy state can absorb a photon, whose energy exactly matches the energy gap between the two states, and transit to the upper
state. The energy of the photon can be described as:
Where h stands for the Planck’s constant and ν is called the resonance frequency and the Larmor frequency in MRI. It depends
on the gyromagnetic ratio, γof the particleas:
For hydrogen, γ= 42.58 MHz / T.
Since the frequency of the photon lies in the radio frequency (RF) rangein NMR experiment, ν is between 60 and 800 MHz for
hydrogen nuclei. In clinical MRI, ν is typically between 15 and 80 MHz for hydrogen imaging.For example, at the most
commonly used field strength of 1.5 Tesla,ν is63.87MHz.

T1 processes
In the equilibrium state, the net magnetization vector lies along the direction of the external magnetic field Bo and is called the
equilibrium magnetization Mo. If we refer MZ as the longitudinal magnetizationand MX and MY as the transverse
magnetization, MZ equals M0 while both MX and MY equal zero at the equilibrium state. However, If we apply the nuclear spin
system to a radio frequency equal to the energy gap between the two spin states, the net magnetization can be changed and
eventually saturate the spin system, so that MZ reaches 0. The so called spin-lattice relaxation time (T1, also named as the
longitudinal relaxation time) describes how MZ returns to its equilibrium value. The equation depicts this process is shown
below:
Mz = Mo (1 - e-t/T1)
Obviously, T1 is the time to reduce the difference between the longitudinal magnetization (MZ) and its equilibrium value by a
factor of e.
Fig.1 T1 process (a) after a 90o pulse (b) after an 180o pulse
If the net magnetization is saturated to the –Z axis (by an 180o pulse) and we record its return to the equilibrium state, the
behavior can be described as:
Mz = Mo (1 - 2e-t/T1)

T2 processes
When the net magnetization was changed to the XY plane, after absorbing Energy from the radio frequency, it starts to
dephase as different spin packets rotates at its own Larmor frequency due to experiencing slightly different magnetic field. The
two main factors for this difference in magnetic field are molecular interactions and variations in B0. The former is considered
as a pure T2 molecular effect while the latter is said to lead to an inhomogeneous T2 effect. The combination of these two
factors is the realresult in the decay of transverse magnetizationandthe combined time constant is called T2 star. The
relationship is showed in the following equation:
1/T2* = 1/T2 + 1/T2inhomo.
To describes the return to equilibrium of the transverse magnetization, MXY, the time constant T2 is defined as in the following
equation.
MXY =MXYo e-t/T2
Fig.2 T2 process
T2 is called the spin-spin relaxation time or the transverse relaxation time and is always equal to or less than T1. The net
magnetization in the XY plane goes to zero and then the longitudinal magnetization grows in until we have Mo along Z.

Rotating frame, Bloch equations and pulse magnetic field


If we define a rotating frame of reference which rotates about the Z axis at the Larmor frequency, it will be much easy to
describe the behavior of the net magnetization vector in this frame than in the laboratory frame. In this case, the magnetization
vector rotating at the Larmor frequency in the laboratory frame will be stationary in our rotating frame.

7/4/2021 2 https://chem.libretexts.org/@go/page/1811
The Bloch equations are a set of coupled differential equations, used to describe the behavior of a magnetization vector under
any conditions. When properly integrated, the Bloch equations will yield the X', Y', and Z components of magnetization as a
function of time.
To create a magnetic field along the X axis, we can put a coil of wire around X axis and pass a current through it. An
alternating current can create an oscillating magnetic field, but when we see it in the rotating frame, it just provides a constant
magnetic field along the X’ axis. This works just as we move the coil about the rotating frame coordinate system at the Larmor
frequency. So the magnetic field generated by the coil passing an alternating current at the Larmor frequency is called the B1
magnetic field. We can generate this pulsed magnetic field but turning the current on and off. 90o pulse and 180o are the mostly
used pulses.

Magnetic field gradient


Magnetic field gradient makes it possible for different regions of spin to be exposed to a different magnetic field so that we are
able to image their positions. In the following sections, we will use Gx, Gy, and Gz for a magnetic field gradient in the x, y and
z directions. The strength of the magnetic field increases along the axis.
A magnetic field gradient in the z direction is highly important for slice selection, which is the selection of spins in a plane
through the object. When a 90o pulse is applied, together with our Gz, only spins in the certain slice will be rotated to the XY
plane. This is mainly because only spins in this slice matches the frequency provided by the RF, as showed in the following
figure.
Fig.3 slice selection magnetic field gradient [2]
The phase encoding gradient is a gradient to impart a specific phase angle to a transverse magnetization vector. After slice
selection, spins in that slice are rotated to the XY plane and if we consider them in the same chemical shift, they should have
the same Larmor frequency. When a gradient in the magnetic field along the X axis is applied, the vectors will process about
the direction of the applied magnetic field at the following frequency:
ν = γ ( Bo + x Gx) = νo + γ x Gx
So when we turn onthe phase encoding gradient, each transverse magnetization vector has its own unique Larmor
frequencyand the description of phase encoding is the same as frequency encoding.
In summary, the slice selection gradient is perpendicular to the slice planewhile the phase encoding gradient is along one of the
sides of the image planeand the frequency encoding gradient the other.

Pulse sequences
A pulse sequenceis defined as a set of RF pulses applied to the object for a specific form of NMR signal.The 90-FID (free
induction decay) sequence, spin-echo sequence and inversion recovery sequence are the most used ones.
In the 90-FID pulse sequence, net magnetization is rotated down into the X'Y' plane with a 90o pulseand the magnitude of the
vector decays with time.
The spin-echo sequence also uses a 90o pulse at first so that the magnetization rotates down into the X'Y' planeand the
transverse magnetization begins to dephase. However, an 180o pulse is then appliedshortly after the 90o pulse and it rotates the
magnetization by 180o about the X' axis. This creates a partially magnetization rephrase and produces the called echosignal.
Fig.4 (a) 90-FID sequence, (b) spin-echo sequence, (c) inversion recovery sequence
In the inversion recovery pulse sequence, an 180o pulse is first appliedto rotate the net magnetization down to the -Z axis.The
magnetization undergoes T1 process and returns toward its equilibrium position along the +Z axis. A 90o pulse is applied
before it reaches the equilibrium, so that the longitudinal magnetization is rotated into the XY plane. The dephasing process
then gives free induction decay.

Basic Imaging Techniques


Now we will talk about basic imaging techniques using various sequences. A good way to show how each imaging method
works is to use timing diagrams. A timing diagram normally includes the radio frequency, magnetic field gradients, and signal
as a function of time. For example, the simplest FT imaging sequence contains a 90o slice selective pulse, a slice selection
gradient, a phase encoding gradient, a frequency encoding gradient, and a signal. The slice selection gradientand the slice

7/4/2021 3 https://chem.libretexts.org/@go/page/1811
selection RF pulseare the first things to be turned on.When they are complete, a phase encoding gradient is applied to the
object. A frequency encoding gradient is turned on afterwards and a signal is recorded, which is in the form of a free induction
decay. To collect all the data needed for an image, the sequence of pulses is usually repeated 128 or 256 times. We define the
time between the repetitions of the sequence as the repetition time, TR. Each time the sequence is repeated, the magnitude of
the phase encoding gradient is changed in equal steps between the maximum amplitude of the gradient and the minimum
value. In this case, we can record 128 or 256 different free induction decaysfor resolving 128 or 256 locations in the phase
encoding direction. Those recorded signals described above can be Fourier transformed to create an image of the location of
spins.
A widely used parameter to quantify image quality is the signal-to-noise ratio (SNR). In an image, it is the ratio of the average
signal for the tissue to the standard deviation of the noise in the background. The SNR of an MRI image has the following
dependent factors: (1) the number of signal producing water molecules in the image voxel (2) the size of the signal each
molecule produces, (3) the quality of the signal detection, and (4) the amount of spurious noise and the statistics of noise
averaging.

Gradient-Echo Imaging
The gradient-echo sequence is intrinsically more sensitive to magnetic field inhomogeneities because of the use of the
refocusing gradient or wind-up gradient. Together with a slice selection gradient, the slice selective RF is turned on to creates a
rotation angle of 90o. The phase encoding gradient Gφ is applied right after. Here, a negative dephasing frequency encoding
gradient -Gf is on at the same time to cause the spins to be in phase at the center of the acquisition period. When the frequency
encoding gradient Gf is applied, an echo is produced, because the gradient refocuses the dephasing which occurred from the
dephasing gradient. This echo is named as a gradient echo, which is different from the echo created by an 180o pulse.
We define echo time (TE) as the time period from the start point of the RF pulse to the maximum of the signal. An equation to
describe the signal intensity versus time and spin density can be written as:
S = k ρ (1-exp(-TR/T1)) exp(-TE/T2*)
Fig.5 Gradient-Echo Imaging timing diagram

Spin-Echo Imaging
In spin-echo imaging, a spin-echo sequence is used and it displays the transverse relaxation time dependence to the signal.
This is especially useful for some tissues which have similar T1 values but different T2. So as showed in the timing diagram in
figure 5, a slice selective 90o pulse, together with the slice selection gradient, is applied to the object. After a time period of
TE/2, an 180o slice selective pulse is turned on in conjunction with the slice selection gradient. Between the two pulses, just
like in the previous Gradient-echo imaging method, the phase encoding gradient Gφ and frequency encoding gradient Gf are
applied at the same time. This time, Gf is necessary because it dephases the spins so that they will rephase by the center of the
echo. Finally, the frequency encoding gradient applied after the 180o pulse, during the time that echo is collected. The signal,
which is the collected echo, is as follows:
S = k ρ (1-exp(-TR/T1)) exp(-TE/T2)
Fig.6 Spin-Echo Imaging timing diagram

Inversion Recovery Imaging


An inversion recovery sequence is used in this imaging technique. It uses a spin-echo sequence to detect the signal so the RF
pulses are 180-90-180. An inversion recovery sequence which uses a gradient-echo signal detection is similar, with the
exception that a gradient-echo sequence is substituted for the spin-echo part of the sequence.
S = k ρ (1-2exp(-TI/T1)+exp(-TR/T1))
Fig.7 Inversion Recovery Imaging timing diagram

Multi-slice Imaging
Multi-slice imaging makes a volume of anatomy to be imaged in the shortest time possible. The time to obtain an image is
equal to the product of the TR value and the number of phase encoding steps. While TR takes time in seconds, the large steps
of phase encoding last much longer. So when we take a careful look at the timing diagram of the gradient-echo sequence
introduced above, we can easily find out that the most of the sequence time is unused. To make full use of the scanning time,

7/4/2021 4 https://chem.libretexts.org/@go/page/1811
other slices can be excited as show in figure 8. However, other RF frequencies of the 90o should be used in order to prevent the
interactions between those slices.
Fig.8 Multi-slice Imaging timing diagram

T1, T2, and ρ Images


Sometimes, properties of the spins in a tissue, such as the spin-lattice relaxation time (T1), spin-spin relaxation time (T2), and
the spin density (ρ) are searched for. There are several methods of calculating T1, T2, and ρ values. A T1 image can be created
from a series of images using the same pulse sequence with varying TR. The signal for a given pixel can be plotted for each
TR value and the best fit line from the spin-echo equation drawn through the data to find T1, as show in figure 10.
Similarly, to produce a T2 image, a series of images using a spin-echo pulse sequence with varying TE is recorded. The signal
for a given pixel can be plotted for each TE value and the best fit line from the spin-echo equation drawn through the data to
find T2, as shown in figure 12. The spin-echo equation is mentioned in the previous Pulse Sequences section. Once T1 and T2
are obtained, it’s easy to acquire spin density by using the spin echo signal equation and any spin echo signal.
Fig.9 a series of imaging recorded with various TR under spin-echo sequence[2]
Fig.10 signal – TR graph
Fig. 11 a series of imaging recorded with various TE under spin-echo sequence[2]
Fig. 12 signal – TE graph

Chemical contrast agent


Although MRI has a good resolution and is very sensitive compared with other imaging modalities, such as positron emission
tomography, chemical contrast agents (CA), substances which are introduced into the body to change the contrast between the
tissues, play a vital role in the imaging procedure. Contrast agents may be injected intravenously to enhance the appearance of
blood vessels, tumors or inflammation or directly injected into a joint in the case of arthrograms. Normally, they can be
divided into two main categories, namely T1 contrast agents and T2 contrast agents and they shorten the according relaxation
time, respectively. The most commonly used compounds for contrast enhancement are gadolinium-based or iron-based. The
former is paramagnetic and belongs to T1 contrast agents, as it changes the contrast by creating time varying magnetic fields
which promote spin-lattice and spin-spin relaxation of the water molecules. The latter is superparamagnetic and comes to the
second group, as it changes T2 of the water molecules around by distorting the Bo magnetic field. Several well-known contrast
agents are Magnevist (Gd-DPTA), Dotarem (Gd-DOTA) and superparamagnetic iron oxide (SPIO). Free Gadolinium ions in
the human body may arose a certain kind of disease called nephrogenic systemic fibrosis(NSF). That’s why Gd needs to be
connected with organic polymers with great care for safety reason.

Applications
In clinical practice, MRI is used to distinguish pathologic tissue, such as a brain tumor, from normal tissue.

Hardware
Fig. 13 Hardware overview[2]
Figure 13 is an overview of the magnetic resonance imaging scanning system. Computer plays the most important role in this
procedure as it controls all the components. The magnet produces the Bo field which is critical for the imaging process, along
with those gradient coils for gradients in Bo in the X, Y, and Z directions. The RF coil generates the B1 magnetic field for
rotating the spins by 90o, 180o, or any other value selected by the pulse sequence and also functions as the signal detector from
the spins within the body. The RF amplifier increases the pulses power from mW to kW while the gradient amplifier increases
the power of the gradient pulses to a sufficient level to drive the gradient coils. Since outer RF pulses may cause an additional
impact on imaging, this scan room is enclosure with RF shields.

Advantages and disadvantages


Despite of the rapid development and its improvement in the past several decades, MRI still has its shortcomings. The scan
time of MRI for cardinal tissues lasts more than 30min, which is much longer than a CT scan and the expense of an MRI
system is still high. Patients need to be stationary during the scan to prevent distortion of images. Besides, patients with

7/4/2021 5 https://chem.libretexts.org/@go/page/1811
pacemakers cannot have MRIs and claustrophobic patients cannot usually make it through a MRI. The machine makes a
tremendous amount of noise during a scan due to the rising electrical current in the wires of the gradient.
Even with these disadvantages, we cannot ignore the tremendous contribution MRI has made in clinical practice in the past
years. Because variations in T1 and T2 values are so much greater than variations in tissue density, MRI provides better soft-
tissue contrast than plain radiography or computed tomography (CT).[3] In addition, it is easy to take a MRI slice in any
direction, without changing the position of the patient. Compared to PET and CT, MRI is also harmless to the human body.
The high resolution of imaging makes MRI an effective imaging modality and its opportunity in the future should be even
more promising.

References
1. http://en.Wikipedia.org/wiki/MRI
2. The Basic of MRI. Dr. Joseph P. Hornak
3. Magnetic resonance imaging, R. Edelman et al. The New England Journal of Medicine, Vol.328, No.10
4. Magnetic resonance imaging,D.C. Lee et al. Chapter 16 inPRACTICAL SIGNAL AND IMAGE PROCESSING IN
CLINICAL CARDIOLOGY2010, Part 2, 251-273
5. Brain magnetic resonance imaging with contrast dependent on blood oxygenation, S. Ogawa et al. Proc. Natl. Acad. Sci.
USA 87 (1990)
6. The future technical development of MRI, SJ. Riederer. J Magn Reson Imaging 1996 Jan-Feb; 6(1):52-6

Contributors and Attributions


Tang Tang (UCD)

7/4/2021 6 https://chem.libretexts.org/@go/page/1811
NMR - Interpretation
Nuclear Magnetic Resonance (NMR) interpretation plays a pivotal role in molecular identifications. As interpreting NMR
spectra, the structure of an unknown compound, as well as known structures, can be assigned by several factors such as
chemical shift, spin multiplicity, coupling constants, and integration. This Module focuses on the most important 1H and 13C
NMR spectra to find out structure even though there are various kinds of NMR spectra such as 14N, 19F, and 31P. NMR
spectrum shows that x- axis is chemical shift in ppm. It also contains integral areas, splitting pattern, and coupling constant.

Strategy for Solving Structure


Here is the general strategy for solving structure with NMR:
1. Molecular formula is determined by chemical analysis such as elementary analysis
2. Double-bond equivalent (also known as Degree of Unsaturation) is calculated by a simple equation to estimate the
number of the multiple bonds and rings. It assumes that oxygen (O) and sulfur (S) are ignored and halogen (Cl, Br)
and nitrogen is replaced by CH. The resulting empirical formula is CaHb

3. Structure fragmentation is determined by chemical shift, spin multiplicity, integral (peak area), and coupling constants
( J, J)
1 2

4. Molecular skeleton is built up using 2-dimensional NMR spectroscopy.


5. Relative configuration is predicted by coupling constant (3J).

1H NMR
Chemical Shift
Chemical shift is associated with the Larmor frequency of a nuclear spin to its chemical environment. Tetramethylsilane
(TMS, (CH ) Si ) is generally used as an internal standard to determine chemical shift of compounds: δTMS=0 ppm. In other
3 4

words, frequencies for chemicals are measured for a 1H or 13C nucleus of a sample from the 1H or 13C resonance of TMS. It is
important to understand trend of chemical shift in terms of NMR interpretation. The proton NMR chemical shift is affect by
nearness to electronegative atoms (O, N, halogen.) and unsaturated groups (C=C,C=O, aromatic). Electronegative groups
move to the down field (left; increase in ppm). Unsaturated groups shift to downfield (left) when affecting nucleus is in the
plane of the unsaturation, but reverse shift takes place in the regions above and below this plane. 1H chemical shift play a role
in identifying many functional groups. Figure 1. indicates important example to figure out the functional groups.

Figure 1: 1H chemical shift ranges for organic compound

Chemical equivalence

6/30/2021 1 https://chem.libretexts.org/@go/page/1812
Protons with Chemical equivalence has the same chemical shift due to symmetry within molecule (C H 3 C OC H3 ) or fast
rotation around single bond (-CH3; methyl groups).

Spin-Spin Splitting
Spin-Spin splitting means that an absorbing peak is split by more than one “neighbor” proton. Splitting signals are separated to
J Hz, where is called the coupling constant. The spitting is a very essential part to obtain exact information about the number
of the neighboring protons. The maximum of distance for splitting is three bonds. Chemical equivalent protons do not result in
spin-spin splitting. When a proton splits, the proton’s chemical shift is determined in the center of the splitting lines.
Spin Multiplicity (Splitting pattern)
Spin Multiplicity plays a role in determining the number of neighboring protons. Here is a multiplicity rules: In case of A B m n

system, the multiplicity rule is that Nuclei of B element produce a splitting the A signal into nB + 1 lines. The general
formula which applies to all nuclei is 2 I + 1 , where I is the spin quantum number of the coupled element. The relative
n

intensities of the each lines are given by the coefficients of the Pascal’s triangle (Figure 2).

Figure 2: Pascal's triangle


First-order splitting pattern
The chemical shift difference in Hertz between coupled protons in Hertz is much larger than the J coupling constant:
Δν
≥8 (1)
J

Where Δν is the difference of chemical shift. In other word, the proton is only coupled to other protons that are far away in
chemical shift. The spectrum is called first-order spectrum. The splitting pattern depends on the magnetic field. The second-
order splitting at the lower field can be resolved into first-order splitting pattern at the high field. The first-order splitting
pattern is allowed to multiplicity rule (N+1) and Pascal’s triangle to determine splitting pattern and intensity distribution.

Example 1
The note is that structure system is A3M2X2. Ha and Hx has the triplet pattern by Hm because of N+1 rule. The signal of
Hm is split into six peaks by Hx and Ha (Figure3) The First order pattern easily is predicted due to separation with equal
splitting pattern.

Figure 3: An example of splitting pattern

High-order splitting pattern

6/30/2021 2 https://chem.libretexts.org/@go/page/1812
High-order splitting pattern takes place when chemical shift difference in Hertz is much less or the same that order of
magnitude as the j coupling.
Δv
≤ 10 (2)
J

The second order pattern is observed as leaning of a classical pattern: the inner peaks are taller and the outer peaks are shorter
in case of AB system (Figure 4). This is called the roof effect.

Figure 4: a) first-order pattern and b) second-order pattern of AB system


Here is other system as an example: A2B2 (Figure 5). The two triplet incline toward each other. Outer lines of the triplet are
less than 1 in relative area and the inner lines are more than 1. The center lines have relative area 2.

Figure 5: a) first-order pattern and b) second-order pattern of A2B2 system


Coupling constant (J Value)
Coupling constant is the strength of the spin-spin splitting interaction and the distance between the split lines. The value of
distance is equal or different depending on the coupled nuclei. The coupling constants reflect the bonding environments of the
coupled nuclei. Coupling constant is classified by the number of bonds:
Geminal proton-proton coupling (2JHH)
Germinal coupling generates through two bonds (Figure 6). Two proton having geminal coupling are not chemically
equivalent. This coupling ranges from -20 to 40 Hz. 2JHHdepends on hybridization of carbon atom and the bond angle and the
substituent such as electronegative atoms. When S-character is increased, Geminal coupling constant is increased:
2
Jsp1>2Jsp2>2Jsp3 The bond angle(HCH) gives rise to change 2JHH value and depend on the strain of the ring in the cyclic
systems. Geminal coupling constant determines ring size. When bond angle is decreased, ring size is decreased so that geminal
coupling constant is more positive. If a atom is replace to an electronegative atom, Geminal coupling constant move to
positive value.

Figure 6: Geminal coupling


3
Vicinal proton-proton coupling ( JHH)
Vicinal coupling occurs though three bonds (Figure 7.). The Vicinal coupling is the most useful information of dihedral angle,
leading to stereochemistry and conformation of molecules. Vicinal coupling constant always has the positive value and is
affected by the dihedral angle (?;HCCH), the valence angle (?; HCC), the bond length of carbon-carbon, and the effects of
electronegative atoms. Vicinal coupling constant depending on the dihedral angle (Figure 8) is given by the Karplus equation.
3 2
J = 7.0 − 0.5 cos ϕ + 4.5 cos ϕ (3)

6/30/2021 3 https://chem.libretexts.org/@go/page/1812
When ? is the 90o, vicinal coupling constant is zero. In addition, vicinal coupling constant ranges from 8 to 10 Hz at the and ?
=180o, where ?=0o and ?=180o means that the coupled protons have cis and trans configuration, respectively.

Figure 7: Vicinal coupling


The valence angle(?;Figure 8) also causes change of 3JHH value. Valence angle is related with ring size. Typically, when the
valence angle decreases, the coupling constant reduces. The distance between the carbons atoms gives influences to vicinal
coupling constant

Figure 8: a) Dihedral angle and b) valence angle


The coupling constant increases with the decrease of bond length. Electronegative atoms affect vicinal coupling constants so
that electronegative atoms decrease the vicinal coupling constants.

Integral
Integral is referred to integrated peak area of 1H signals. The intensity is directly proportionally to the number of hydrogen.
13
C NMR
Chemical Shift

Figure 9 shows typical 13C chemical shift regions of the major chemical class.

Figure 9: 13C Chemical shift range for organic compound


Spin-Spin splitting
Comparing the 1H NMR, there is a big difference thing in the 13C NMR. The 13C- 13 C spin-spin splitting rarely exit between
adjacent carbons because 13C is naturally lower abundant (1.1%)
13
C-1H Spin coupling: 13C-1H Spin coupling provides useful information about the number of protons attached a carbon
atom. In case of one bond coupling (1JCH), -CH, -CH2, and CH3 have respectively doublet, triplet, quartets for the 13C
resonances in the spectrum. However, 13C-1H Spin coupling has an disadvantage for 13C spectrum interpretation. 13C-1H
Spin coupling is hard to analyze and reveal structure due to a forest of overlapping peaks that result from 100% abundance
of 1H.

6/30/2021 4 https://chem.libretexts.org/@go/page/1812
Decoupling: Decoupling is the process of removing 13C-1H coupling interaction to simplify a spectrum and identify which
pair of nuclei is involved in the J coupling. The decoupling 13C spectra shows only one peak(singlet) for each unique
carbon in the molecule(Figure 10.). Decoupling is performed by irradiating at the frequency of one proton with continuous
low-power RF.

Figure 10. Decoupling in the 13C NMR


Distortionless enhancement by polarization transfer (DEPT): DEPT is used for distinguishing between a CH3 group, a
CH2 group, and a CH group. The proton pulse is set at 45o, 90o, or 135o in the three separate experiments. The different
pulses depend on the number of protons attached to a carbon atom. Figure 11. is an example about DEPT spectrum.

Figure 11. DEPT spectrum of n-isobutlybutrate

2-dimensional NMR spectroscopy (COSY)


COSY stands for COrrelation SpectroscopY. COSY spectrum is more useful information about what is being correlated.
1H-1H COSY (COrrelation SpectroscopY)
1 1
H- H COSY is used for clearly indicate correlation with coupled protons. A point of entry into a COSY spectrum is one of
the keys to predict information from it successfully. Relation of Coupling protons is determined by cross peaks(correlation
peaks) and in the COSY spectrum. In other words, Diagonal peaks by lines ar e coupled to each other. Figure 12 indicates that
there are correlation peaks between proton H1 and H2 as well as between H2 and H4. This means the H2 coupled to H1 and H4.

Figure 12. 1H-1H COSY spectrum

6/30/2021 5 https://chem.libretexts.org/@go/page/1812
1H-13C COSY (HETCOR)
1H-13C COSY is the heteronuclear correlation spectroscopy. The HETCOR spectrum is correlated 13C nuclei with directly
attached protons. 1H-13C coupling is one bond. The cross peaks mean correlation between a proton and a carbon (Figure 13).
If a line does not have cross peak, this means that this carbon atoms has no attached proton (e.g. a quaternary carbon atom)

Figure 13. 1H-13C COSY spectrum

References
1. Balc*, M., Basic p1 sH- and p13 sC-NMR spectroscopy. 1st ed.; Elsevier: Amsterdam ; Boston, 2005; p xii, 427.
2. Breitmaier, E., Structure elucidation by NMR in organic chemistry : a practical guide. 3rd rev. ed.; Wiley: Chichester, West
Sussex, England, 2002; p xii, 258.
3. Jacobsen, N. E., NMR spectroscopy explained : simplified theory, applications and examples for organic chemistry and
structural biology. Wiley-Interscience: Hoboken, N.J., 2007; p xv, 668.
4. Silverstein, R. M.; Webster, F. X., Spectrometric identification of organic compounds. 6th ed.; Wiley: New York, 1998; p
xiv, 482.

Outside Links
NMRShiftDB: a Free web database for NMR data : nmrshiftdb.chemie.uni-mainz.de/nmrshiftdb
NMR database from ACD/LAbs : www.acdlabs.com/products/spec_lab/exp_spectra/spec_libraries/aldrich.html
NMR database from John Crerar Library : http://crerar.typepad.com/crerar_lib...h_ir_nmr_.html

Problems
Draw the 1H NMR spectrum for 2-Hydroxypropane in CDCl3. Assume sufficient resolution to provide a first-order spectrum
and ignore vicinal proton-proton coupling(3JHH)

Solution
1) the structure of 2-hydoroxyporpane is drawn

Figure out which protons are chemically equivalent, i.e., two methyl (-CH3) groups are chemical equivalent.

6/30/2021 6 https://chem.libretexts.org/@go/page/1812
Figure1): chemical shift of methyl groups (Ha) : 1-2 ppm (?Ha=1.1 ppm); chemical shift of -CH- groups (Hb) moves to
downfield due to effect on aldehyde groups:2-3ppm ( ?Hb=2.4 ppm); chemical shift of aldehyde groups (Hc):9-10 ppm (?
Hc=9.6 ppm)
4) Splitting pattern is determined by (N+1) rule: Ha is split into two peaks by Hb(#of proton=1). Hb has the septet pattern by
Ha (#of proton=6). Hc has one peak.(Note that Hc has doublet pattern by Hb due to vicinal proton-proton coupling.)

Contributors and Attributions


You Jin Seo

6/30/2021 7 https://chem.libretexts.org/@go/page/1812
Integration in NMR
The intensity of the signal is proportional to the number of hydrogens that make the signal. Sometimes, NMR machines
display signal intensity as an automatic display above the regular spectrum. (The exact number of hydrogens giving rise to
each signal is sometimes also explicitly written above each peak, making our job a lot easier.) The intensity of the signal
allows us to conclude that the more hydrogens there are in the same chemical environment, the more intense the signal will be.

Introduction
We can get the following information from a 1H Nuclear Magnetic Resonance (NMR) structure:
1. The number of signals gives the number of non-equivalent hydrogens
2. Chemical shifts show differences in the hydrogens’ chemical environments
3. Splitting presents the number of neighboring hydrogens (N+1 rule)
4. Integration gives the relative number of hydrogens present at each signal
The integrated intensity of a signal in a 1H NMR spectrum (does not apply to 13C NMR) gives a ratio for the number of
hydrogens that give rise to the signal, thereby helping calculate the total number of hydrogens present in a sample.NMR
machines can be used to measure signal intensity, a plot of which is sometimes automatically displayed above the regular
spectrum. To show these integrations, a recorder pen marks a vertical line with a length that is proportional to the integrated
area under a signal (sometimes referred to as a peak)-- a value that is proportional to the number of hydrogens that are
accountable for the signal. The pen then moves horizontally until another signal is reached, at which point, another vertical
marking is made. We can manually measure the lengths by which the horizontal line is displaced at each peak to attain a ratio
of hydrogens from the various signals. We can use this technique to figure out the hydrogen ratio when the number of
hydrogens responsible for each signal is not written directly above the peak (look in the links section for an animation on how
to manually find the ratio of hydrogens as described here).
Now that we’ve seen how the signal intensity is directly proportionate to the number of hydrogens that give rise to that signal,
it makes sense to conclude that the more hydrogens of one kind there are in a molecule (equivalent hydrogens, so in the same
chemical environment), the more intense the corresponding NMR signal will be. Here's a model that may help clear up
some of the uncertainties. NMR model

Study tips
1. When trying to build a structure for an unknown molecule from NMR data, find a common divisor that reduces the
integrated areas of all signals to small whole numbers when the exact number of hydrogens responsible of each peak is
NOT given.
2. If, for example, the total number of hydrogens you count from your NMR spectrum is 5, but you know that there are
supposed to be 10 hydrogens in the molecular formula, then it is a sure sign that the your molecule is symmetrical. You
only find half the hydrogens in your NMR spectrum because the symmetry places hydrogens in the same chemical
environment-- giving rise to the same peaks. (Again, the spectrum simply gives you a ratio.)
3. A peak with 3 hydrogens commonly belongs to a -CH3 group.
4. More instructions on how to approach an NMR problem are given in the links below.

Outside Links
Animation: how to manually find the ratio of hydrogens in a spectrum
http://www.wfu.edu/~ylwong/chem/nmr/h1/integration.html
Clear cut instructions on how to approach a 1H NMR spectrum problem
http://www.chemhelper.com/nmrspechelp2.html

References
1. Schore, Neil E. and Vollhardt, K. Peter C. Organic Chemistry: Structure and Function. New York: Bleyer, Brennan, 2007.
(405-407)
2. UC Davis 118A Supplementary Booklet for the Laboratory/Discussion (Fall quarter 2008)_ Page 39

6/11/2021 1 https://chem.libretexts.org/@go/page/1813
Problems
1.) True or False? The number of hydrogens determines the intensity of a signal.
2.) Give the number of signals, the chemical shift value for each signal, and the number of integrating hydrogens for
CH3OCH2CH2OCH3?

3.)

4.)

Answers
1. False. The relative number of hydrogens determines the intensity of a signal. The signal given by the three hydrogens in
CH3CH2CHCl2 will not have the same intensity as the three hydrogens in ClCH2OCH3.
2. There are 2 signals. One is at 3.3 ppm (6 hydrogens); the other at 3.5 ppm (4 hydrogens).
3. a and d
4. c
5. d

6/11/2021 2 https://chem.libretexts.org/@go/page/1813
Pascal’s Triangle
The Pascal’s triangle is a graphical device used to predict the ratio of heights of lines in a split NMR peak. To construct the
Pascal’s triangle, use the following procedure.
Step 1: Draw a short, vertical line and write number one next to it.

Step 2: Draw two vertical lines underneath it symmetrically.

Step 3: Connect each of them to the line above using broken lines.

Step 4: Each of the two lines is connected to the single line above, which carries the number one. Therefore, write number one
next to each line.

Step 5: Draw three vertical lines symmetrically underneath the two lines.

Step 6: Connect each of them to the nearest line(s) above.

7/1/2021 1 https://chem.libretexts.org/@go/page/14971
Step 7: Each terminal line is connected to one line above, which carries the number one. Therefore, write number one next to
each of them. The internal line is connected to two lines above, each carrying the number one. 1 + 1 = 2; write number two
next to the internal line.

Continue the process as far down as necessary.

see also (n+1) Rule, coupling constant

Contributors and Attributions


Gamini Gunawardena from the OChemPal site (Utah Valley University)

7/1/2021 2 https://chem.libretexts.org/@go/page/14971
NMR Hardware
The theory sections in this wiki have been devoted to understanding how NMR works and we have made several attempts to
investigate atomic nuclei using Magnetic fields, coils, and RF fields. This section is devoted to developing and NMR
spectrometer from the theory side then introducing modifications to this simple spectrometer that makes NMR spectroscopy
much easier. It is the goal of this page to provide enough information about NMR hardware that one (given the proper
materials) could build an NMR spectrometer in their garage!

Basic NMR Spectrometer


Shown below is the most basic NMR spectrometer you can make. For an NMR spectrometer to be operational we first need a
magnet to split the degenerate nuclear energy levels of the sample. We also need a coil to apply an oscillating RF field which
will be used to drive transitions between the energy levels as well as detect the signals. Therfore we will also need and RF
oscillator and a oscilloscope to generate and detect the RF waves.

Magnet
The magnet is arguably the most important part of the NMR instrumentation. The magnetic field drives the splitting of the
energy levels and allows for a change in the distribution of energy levels of the degenerate ground state. As we have seen
previously, the larger the magnetic field, the larger the spin difference between the upper and lower levels allowing for a
stronger signal to be observed. In addition to the splitting of energy levels, several NMR interactions are dependent on the
magnetic field. Both the dipolar and CSA interactions are multiplied by an increase in magnetic field, while the Quarupolar
coupling interaction is decreased with increasing magnetic field.
Insert picture
Both superconducting and electromagnets magnets are comprised of loops of wires with a current passing through, which
results in the formation of a magnetic field. The "north pole" of the magnetic field (using the right hand rule) is generated in
the direction perpendicular to the direction of the current flow.
show equation

Superconducting Magnets
The magnet is typically composed of an alloy that is super conducting at low temperature. While the theory of super
conductivity is beyond the scope of this page, several unique properties of super conductors are important. This alloy is made
into a wire which is wrapped into a coil and has a current passed through it. This generates the magnetic field! The wires are
kept cooled with liquid helium at 4K, which significantly reduces the resistance of the wire to nearly zero. Therefore the
current passing through the wire onyl needs to happen once and the spectrometer does not need to be plugged in. The liquid
helium Dewar is insulated by a vacuum chamber and then a liquid nitrogen dwar which itself is insulated by another vacuum
chamber.

Electromagnets

Probe
The probe is a specialized piece of hardware that contians the coils and sample and therefore is responsible for tuning to the
correct frequency, applying RF pulses, receiving the signal from the nuclei. More specialized probes can also spin the sample,
vary the temperature, and change the angle of the sample and coil.
Lets first optimize the efficiency of the coil so that the maximum power is transferred to the sample from RF source and from
the sample to the receiver. Since both the power transfer from the RF source and the sample to the receiver is done in the coil
we can rite that the power is the

P =V I (1)

were P is power, V is the voltage, and I is the amperage. Now an EMF is generated in the coil when the magnetization
processes. ANd EMF is considered to be a pure voltage source and is measure in volts. However the coil has a finite resistance
to it which is related to the voltage by

6/11/2021 1 https://chem.libretexts.org/@go/page/13125
V = IR (2)

so the total voltage measurable voltage is then


Vtot = VEMF − I R (3)

Now measuring a voltage is fairly straightforward and will not be discussed. However The concept of impedance needs to be
addressed. Impedence is another type of resistance that stems from a phase lag between the current and the voltage.
Specifically, the current lags behind the voltage. As implied by the term "phase" the impedance is a complex quantity, with the
rela part being the resistance and the imaginary part being called the reactance. Reactance is the opposition of a circuit element
to a change in the current or voltage.
We can then show the Impedance, Z, is then

Z = \absZ e = R + iX (4)

where θ is the thase shift in polar coordinates, and X is the reactance.


Then
V = IZ (5)

Now our total voltage can be reordered to show


Vtot = VEMF − I Z (6)

Lets first start with our coil. The coil is an inductor has an induction of L given by the coil dimensions as
2 2
r n
L = (7)
9r + 10l

where r is the radius of the coil, n is the number of turns, and l is the length of the coil. In our coil we will have an oscillating
magnetic field which generates both an oscillating current and voltage which can be described as

V (t) = V0 cosωt (8)

V0
I (t) = sinω(t) (9)
ωL

we can then immediately see that the voltage and current are 90 degrees out of phase leading to a completely imaginary
impedence

ZL = iωL (10)

There are several points to this. First, we now see that the current is frequency dependent, that is at high frequencies the
current is very small, but at low frequencies the current is maximized. This has ramification for both application of power to
the coil and detection as high frequency nuclei will then be harder to detect! Remember that we are trying to measure a voltage
and that any decrease in current will directly decrease the voltage and therefore the signal that we receive! The inductor also
has a finite resistance, allowing it to behave as a resistor as well. For a oscillating voltage the current is then
V0
I (t) = cosωt (11)
R

assuming the time dependent voltage is of the form


V (t) = V0 cosωt (12)

as we can see there is no phase shift so the impedance of the resistor is then
ZR = R (13)

and the current is directly related to the voltage. We have now reached quite a predicament in the high frequency range. That is
the power scales down significantly when we go to high frequency. Therefore we need some way to "reverse" this effect. One
way to do this is to use a capacitor. The current flowing across a capacitor is given by
−C ωV0 sinωt (14)

6/11/2021 2 https://chem.libretexts.org/@go/page/13125
and we can see that this scales directly with frequency. Therefore capacitors allow high frequencies to pass through them and
block low frequencies. Since they are 90 degrees out of phase with the voltage, they exhibit the same reactance that inductors
do

RF Circuit
Matching and Tuning
Cross Diodes
Amplifier
Detector
A simple description of the NMR detector is given in the Detectors section. The function of the NMR detector is to detect the
nuclear transitions occurring in the experiment. As the magnetization is precessing, it generates an electromotive force (EMF).
The induced EMF is detected through the coil and sent to the receiver.

EM F = (15)
dt

where dϕ is the magnetic flux induced from the magnetization vector processing.
Let's assume our magnetization vector has length m. The time dependence of m is then
m(t) = |m|[sinψcos(ω0 t + ξ0 )ex + sinψsin(ω0 t + ξ0 )ey + cosψ ez (16)

for a spherical basis set. The resultant magnetization at point r from the precession is
m0
B(r, t) = [3(m(t) ⋅ er )er − m(t)] (17)
3
4πr

the magnetic flux through the coil, distance r away from magnetic vector, is given by a surface integral

ϕ(t) = ∫ ∫ B(r, t) ⋅ nrdrdθ (18)

Solving this integral results in


rcoil 2π
m0 dr
ϕ(t) = − ∫ ∫ dθm(t) ⋅ n (19)
2
4π 0 r 0

which can be simplified if the coil is positioned such that it is in the x-z or y-z plane (n=ey or ex). (check math).
Interestingly,using only one coil, we are only able to detect the the magnetization is precessing, not which direction it is
precessing. We circumvent this issue by employing quadrature detection.

Quadrature Detection
To find the direction of the magnetization precession, the NMR signal is split into 2 parts and applying a phase to each part.
Relation to rotating frame
Problems with imperfect phases
reference coherence pathways
describe how its done

Shims
The shims are used to obtain a homogenous field around the sample. They are named after the thin metal pieces NMR
technicians would insert into the magnetic field in the to manually make a homogenous magnetic field. Nowadays, the shims
are coils of wire which pass current to generate magnetic fields in certain directions. These shims are not just about the z, y
and z directions, the vary across a variety of planes to ensure a very homogenous field is achieved.

6/11/2021 3 https://chem.libretexts.org/@go/page/13125
If we want our magnetic field to be homogenous, then the space we want to be homogenous must be absent of any charges or
current. Therefore, a homogenous magnetic field obeys the Laplace Equation, more commonly written as
2 2 2
d d d
( + + ) B0 = 0 (20)
2 2 2
d x d y d z

Those familiar quantum mechanics may recognize this from the Schroedinger equation in the forms
2
∇ ψ (21)

which represents the potential energy in three dimensions. Solving this second order differential is indeed challenging. As with
any field, it can be represented by a linear of any complete orthogonal basis set. A particularly useful basis set is the spherical
harmonics basis set. B) under this basis set is described as
∞ n
r
n
B0 = ∑ ∑ Cnm ( ) Pnm (cos θ) cos[m(ϕ − ψnm )] (22)
a
n=0 m=0

where r is a distance of some point, a is the bore radius of the magnet, and P nm is the Legendre polynomial.
Zonal Harmonics
When m =0 the variation of B0 with (\phi\) goes away and the equation is reduced to

r n
B0 = ∑ Cn0 ( ) Pnm (cosθ) (23)
a
n=0

These inhomogeneities can be cancelled by application of the z shims.


Tesseral Harmonics
For the remainder of the m values, the off-z axis shims must be used.
We must then apply magnetic fields in all directions in order to cancel the inhomegenties in B0.Since modern shims are coils of
wire, the current must be adjusted to create the proper field to cancel to inhomogeneity. Commonly the shims are listed in
cartesian coordinated.
Conversion equation
Show the Chart of conversion
Shim Mixing
Unfortunately, the shims are not independent, that is adjusting one shim changes the magnetic fields of the previously set
shims.

Problems
Questions
1. Derive the magnetic flux assuming the coil is in the y-z plane
2. Derive the magnetic flux assuming the coil is in the x-z plane

Answers

6/11/2021 4 https://chem.libretexts.org/@go/page/13125
NMR Spectroscopy in Lab: Complications
There will be cases in which you already know what the structure might be. In these cases:
You should draw attention to pieces of data that most strongly support your expected structure. This approach will
demonstrate evaluative understanding of the data; that means you can look at data and decide what parts are more crucial
than others.
You should also draw attention to negative results: that is, peaks that might be there if this spectrum matched another,
possible structure, but that are in fact missing.
One of the most complicated problems to deal with is the analysis of a mixture. This situation is not uncommon when students
run reactions in lab and analyse the data.
Sometimes the spectra show a little starting material mixed in with the product.
Sometimes solvents show up in the spectrum.
As you might expect, the minor component usually shows up as smaller peaks in the spectrum. If there are fewer molecules
present, then there are usually fewer protons to absorb in the spectrum.
In this case, you should probably make two completely separate sets of data tables for your analysis, one for each
compound, or else one for the main compound and one for impurities.
Remember that integration ratios are really only meaningful within a single compound. If your NMR sample contains some
benzene (C6H6) and some acetone (CH3COCH3), and there is a peak at 7.15 that integrates to 1 proton and a peak at 2.10 ppm
integrating to 6 protons, it might mean there are 6 protons in acetone and 1 in benzene, but you can tell that isn't true by
looking at the structure. There must be six times as many acetone molecules as benzene molecules in the sample.
There are six protons in the benzene, and they should all show up near 7 ppm. There are six protons in acetone, and they
should all show up near 2 ppm. Assuming that small integral of 1H for the benzene is really supposed to be 6H, then the large
integral of 6H for the acetone must also represent six times as many hydrogens, too. It would be 36 H. There are only six
hydrogens in acetone, so it must represent six times as many acetone molecules as there are benzenes.
Similarly, if you have decided that you can identify two sets of peaks in the 1H spectrum, analysing them in different tables
makes it easy to keep the integration analysis completely separate too ; 1 H in one table will not be the same size integral as 1
H in the other table unless the concentrations of the two compounds in the sample are the same.
However, comparing the ratio of two integrals for two different compounds can give you the ratio of the two compounds in
solution, just as we could determine the ratio of benzene to acetone in the mixture described above.
We will look at two examples of sample mixtures that could arise in lab. Results like these are pretty common events in the
labIn the first example, a student tried to carry out the following reaction, a borohydride reduction of an aldehyde. The
borohydride should give a hydride anion to the C=O carbon; washing with water should then supply a proton to the oxygen,
giving an alcohol.

Her reaction produced the following spectrum.

Chris Schaller 7/4/2021 NMR Spectroscopy in Lab.1 https://chem.libretexts.org/@go/page/4191


(simulated data)
From this data, she produced the table below.

Notice how she calculated that ratio. She found a peak in molecule 1, the aldehyde, that she was pretty sure corresponded to
the aldehydic hydrogen, the H attached to the C=O; in other words, the CH=O. She found another peak from molecule 2, the
alcohol, that she was pretty sure represented the two hydrogens on the carbon attached to oxygen, the CH2-O.
The integrals for those two peaks are equal. They are both 2H in her table. However, she notes that within each molecule, the
first integral really represents 1H and the second represents 2H. That means there must be twice as many of molecule 1 as
there are molecule 2. That way, there would be 2 x CH=O, and its integral would be the same as the 1 x CH2-O in the other
molecule.
One way to approach this kind of problem is to:
choose one peak from each of the two compounds you want to compare.
decide how many hydrogens each peak is supposed to represent in a molecule. Is it supposed to be a CH2, a CH, a CH3?
divide the integral value for that peak by that number of hydrogens it is supposed to represent in a molecule.
compare the two answers (integral A / ideal # H) vs (integral B / ideal # H).
the ratio of those two answers is the ratio of the two molecules in the sample.
So there is twice as much aldehyde as alcohol in the mixture. In terms of these two compounds alone, she has 33% alcohol and
66% aldehyde. That's ( 1/(1+2) ) x100% for the alcohol, and ( 2/(1+2) ) x100% for the aldehyde. That calculation just
represents the amount of individual component divided by the total of the components she wants to compare.
There are a number of things to take note of here.
Her reaction really didn't work very well. She still has majority starting material, not product.
She will get a good grade on this lab. Although the experiment didn't work well, she has good data, and she has analyzed it
very clearly.
She has separated her data table into different sections for different compounds. Sometimes that makes it easier to analyze
things.

Chris Schaller 7/4/2021 NMR Spectroscopy in Lab.2 https://chem.libretexts.org/@go/page/4191


She has noted the actual integral data (she may have measured the integral with a ruler) and also converted it into a more
convenient ratio, based on the integral for a peak that she felt certain about.
She went one step further, and indicated the internal integration ratio within each individual compound.
She calculated the % completion of the reaction using the integral data for the reactant and product, and she made clear
what part of the data she used for that calculation. A similar procedure could be done if a student were just trying to
separate two components in a mixture rather than carry out a reaction.
She also calculated the overall purity of the mixture, including a solvent impurity that she failed to remove.
However, CHCl3 is not included in her analysis of purity. CHCl3 really isn't part of her sample; it was just present in the
NMR solvent, so it doesn't represent anything in the material she ended up with at the end of lab.
Another student carried out a similar reaction, shown below. He also finished the reaction by washing with water, but because
methanol is soluble in water, he had to extract his product out of the water. He chose to use dichloromethane for that purpose.

He obtained the following data.

From this data, he constructed the following table.

There are some things to learn about this table, too.


Does the integration ratio really match the integral data? Or is this just wishful thinking?
This table might reflect what he wants to see in the data. But what else could be in the data?
CHCl3 is often seen in NMR spectra if CDCl3 is used for the NMR sample. It's there, at 7.2 ppm.
"Leftover" or residual solvent is very common in real lab data. There it is, CH2Cl2 from the extraction, at 5.4 ppm.
What about water? Sometimes people don't dry their solutions properly before evaporating the solvent. There is probably
water around 1.5 to 1.6 ppm here.
This student might not get a very good grade; the sample does not even show up in the spectrum, so he lost it somewhere. But
his analysis is also poor, so he will really get a terrible grade.

Contributors and Attributions


Chris P Schaller, Ph.D., (College of Saint Benedict / Saint John's University)

Chris Schaller 7/4/2021 NMR Spectroscopy in Lab.3 https://chem.libretexts.org/@go/page/4191


Pulse Sequences
A pulse sequence is a succinct visual representation of the pulses and delays used in a certain NMR experiment. Depending on
the experiment there may be hundreds of pulses! This page is dedicated to understanding what a pulse sequence is and how to
understand the pictorial representations, as well as terms commonly used to describe parts of pulse sequences.

How to Read and Understand a Pulse Sequence


The most basic pulse sequence is a single pulse followed by detection of the signal. This is shown in figure 1. The pulse is
represented by the large black bar and the acquisition or detection period is denoted by the squiggly line which represents the
free induction decay (FID) of thr signal generated. The arrow below the sequence is the direction in which time is increasing.
In all pulse sequences, time increases from left to right.

Figure 1. A basic pulse sequence.

Pulses
A pulse is a collection of oscillating waves with a broad range of frequencies (~1/4*pulse length) used to rotated the bulk
magnetization. Pulses are described using two parameters, the angle and the phase. Typically, the angle and the phase are
shown above the pulse. The angle can vary from 0°-180° While the phase can take on values from 0°-360°. Commonly the
phase of the pulse is referred to the axis along which the B field is applied. For instance, shown below are the effects of a 90
1

pulse with a phase along x axis. This rotates the magnetization down along the y axis (right hand rule). While this example is
fairly straightforward, pulse phases can become incredibly complex. More information about pulse phases will be covered in
the phase cycling section. Pulses may be refereed to by just their angle ("a nintey pulse"), their angle in radians (a "pi over
π

two pulse") or their angle and phase "ninety x", (90°)x.

Figure 2. The effect of a (90o)x

Acquisition
Now we move further to the right, increasing in time and see a squiggly line. This line represents the signal, which shows
decay, as expected from T2 and T2* relaxation.

Multiple Pulse Sequences


Most pulse sequences have more than one pulse, which add a degree of complexity in to interpretations of the pulse sequence.
Using multiple pulses have many advantageous including separation of NMR interactions, measuring relaxation timescales,
and signal enhancement. Two additional features are present in multiple pulse experiments, delays and loops.

7/2/2021 1 https://chem.libretexts.org/@go/page/8023
Figure 3. Multi-pulse pulse sequence.
Delays
Delays are times when pulsing and detection are not occurring and are very important in pulse sequences, as separation of
NMR interactions require precise timings. Additionally, creating proper delays can preserve select information as well as
enhance signals. Delays are often referred to as "times". Below is a list of common delays:
Relaxation Delay: Amount of time needed to wait before the next experiment can begin. Typically this is on the order 3-5
T1.
Rotor Synchronized Delay (pulses): These delays are set so pulses remain centered in time with an integer multiple of the
rotor period. Delays will then be:
1
Delay = n ∗ τr − (Pi + Pf ) (1)
2

where n is an integer number, τ is the spinning speed, Pi is the pulse length of the pulse before the delay and Pf is the pulse
r

length of the pulse after the delay.


Evolution (time) Delay: This is time during which the magnetization is precessing to satisfy the an equation relating to the
separation of a variable.
Storage (time) Delay: This is the time at which the bulk magnetization is stored along the external magnetic field to allow
for sample manipulation, such a hopping or flipping.
Loops
Loops are repeating parts of the pulse sequence. They are typically employed in sequences to boost the S/N.
Echoes
Echoes are appear in pulse sequences as 2 FIDs placed in opposite directions. Figure 3 shows what an echo looks like during
the CPMG loop.

7/2/2021 2 https://chem.libretexts.org/@go/page/8023
Phase Cycling and Coherence Transfer Pathways
In NMR simply throwing a string of pulses with random phases is a recipe for disaster. Phase cycling is a way to choose the
proper coherence pathway in order to acquire the NMR signal which has the interactions you are trying to acquire.
Additionally, phase cycling eliminates artifacts that can appear in the NMR spectrum which are not real NMR signals.

NMR Artifact
Two problems are commonly encountered during detection of the NMR signal. This first occurs when the detector should be
detecting zero but actually records a non-zero value. This is known as a receiver baseline error, commonly known as a zero
glitch. The other type of artifact occurs from errors in the quadrature detection of the NMR signal. NMR spectra can contain
both of these and can severely complicate spectra. Fortunately, we can eliminate these by employing phase cycling.

Zero Glitch
If the detector is turned on and the voltage is non-zero, then a spike will form in the exact center of the spectrum. A zero glitch
is easy to check for by changing the carrier frequency. If the spike moves to the center of the spectrum, you have a zero glitch.

Quadrature Ghost (Image)


From our discussion on the NMR receiver, quadrature detection is used. In this, the frequency is split and a reference phase is
superimposed on one of the carrier frequencies. Successful recombination of the frequencies is of the form
S = cos +i sin (1)

and the resulting Fourier Transform is a peak at a certain frequency. However, when the phase is not set correctly, or the
efficiency of the hardware changes, the signal now has a different phase (not 90°). This results in a signal at the exact opposite
frequency.

Coherence Transfer Pathway


Conferences are developed by the application of pulses of a phase ϕ . As our discussion from the Bloch equations, three
different conferences can develop p=-1,0,+1 with the application of the pulse, depending on the efficiency of the coherence
transfer. We also saw that the relaxation processes decay the coherence. Additionally, the coherence will not change without
the application of a pulse. We can then construct a coherence transfer diagram which illustrates the coherence path as a
function of the pulse sequence.

As you can see from above, there are multiple coherences at the end of the pulse sequence. In fact, for a spin=1/2 system with
n pulses, 3n coherence pathways evolve! We must then select the proper coherence to detect! Luckily, most receivers are
equipped to only detect on the -1 coherence. This limits detection to only a fraction of the total coherences that evolve.
Therefore, we must develop a way to selectively obtain the coherences with the interactions we want to probe.

Nested Phase Cycling

6/20/2021 1 https://chem.libretexts.org/@go/page/17016
Phase cycling is a way to selectively extract a single coherence that has evolved with a selected pathway. The only way to
change a coherence is to apply a pulse by phase ϕ . Therefore, we must know the coherence pathway we are interested in and
we need to set the phases of the pulses and receiver accordingly. Luckily it is fairly easy to calculate which phases are needed.
We know that the coherence must be =-1 at the end of the pulse sequence for the signal to be detected by the receiver. The
master equation is listed below
MT

ϕrec = − ∑ Δp ϕm (2)

m=1

where ϕ is the receiver phase, Δp is the desired change in coherence and ϕ is the phase of the pulse m.
R m

If we only ever used one pulse, then only three coherences would evolve. However, with multiple pulses, multiple signals will
be at the -1 coherence during detection. We can average these by changing the phase ϕ . The phase should be incremented, at
m

most by

Δϕm = (3)
nm

where
(max) (min)
nm = Δpm − Δpm +1 (4)

Let's consider the pulse sequence in figure 1. Let's also assume we do not have a perfect detector, and some of the
magnetization from the +1 and 0 coherences can be detected.

Figure 2. Desired coherence pathway for a 1-pulse experiments.


(max) (min)
Our desired pathway is 0 to -1. Δp 1
= 1 and Δp = −1 Then n1=3. Therefore we need to increment
1
ϕ by steps of
120^∘ to create the phase cycle. Our desired Δp = −1 − 0 = −1 . Thus
ϕrec = ϕ1 (5)

Our full phase cycle is for both ϕ and ϕ = 0, 120, 240. Therefore 3 scans are needed to complete the phase cycle.
rec 1

Experiments should have a multiple of 3 to make sure a proper spectrum is acquired. Let's take a look at a more complex pulse
sequence, the Hahn echo sequence. The pulse sequences with the desired pathway is outline below.

Figure 3. The Hahn echo coherence transfer pathway with the desired pathway highlighted in blue. Coherences which evolve,
but are not desired are green.
We must once again calculate the the master equation, however we now have to consider the second pulse. Lets first break the
sequence up into 2 steps, the first step will be for ϕ and the second for ϕ . For ϕ :
1 2 1

Δp1 = 1 − 0 = 1 (6)

6/20/2021 2 https://chem.libretexts.org/@go/page/17016
nm == 1 − (−1) + 1 = 3 (7)



Δϕ1 = = 120 (8)
nm

For ϕ :2

Δp2 = −1 − 1 = −2 (9)

nm == 1 − (−1) + 1 = 3 (10)

2π ∘
Δϕ1 = = 120 (11)
nm

Then

ϕrec = −(ϕ1 − 2 ϕ2 ) (12)

and the phase cycle would be


∘ ∘ ∘ ∘ ∘ ∘ ∘ ∘ ∘
ϕ1 = 0 120 240 0 120 240 0 120 240
∘ ∘ ∘ ∘ ∘ ∘ ∘ ∘ ∘
ϕ2 = 0 0 0 120 120 120 240 240 240 (13)

∘ ∘ ∘ ∘ ∘ ∘ ∘ ∘ ∘
ϕrec = 0 240 120 240 120 0 120 0 240

As you can see from above, this is a nested phase cycle, in which the three steps of ϕ are incremented successively, while ϕ
1 2

is incremented only after completion of a full phase cycle is completed for ϕ while this creates a robust phase cycle, there are
1

several shortcuts which enable the phase cycles to be shortened. For instance, if no quadrature ghosts are detected, meaning
only to -1 coherence is observed, then only one pulse needs to be phase cycled. This gives
∘ ∘ ∘
ϕ1 = 0 120 240
(14)
∘ ∘ ∘
ϕrec = 0 240 120

or
∘ ∘ ∘
ϕ2 = 0 120 240
(15)
∘ ∘ ∘
ϕrec = 0 240 120

Note that the phases for ϕ and ϕ are obtained when the opposing ϕ =0
1 2

Phase Cycling with Coherences>1


Until now, we have considered a spin=1/2 nucleus with coherences = -1, 0 ,+1, which we readily obtained from the spherical
representation of the Bloch equations. However, the Bloch equations are only valid for a single uncouple spin, which is not the
case for most systems. Therefore, access to higher coherences are possible for coupled spin=1/2 nuclei as well as quadrupolar
nuclei.

6/20/2021 3 https://chem.libretexts.org/@go/page/17016
Solid State Experiments
The solid state NMR community has developed hundreds of techniques to investigate different parts of the NMR Hamiltonian.
This section gives a brief, by no means comprehensive of a few techniques that are helpful in probing select interactions.
Additional resources useful for understanding 2D NMR are available.

Magic Angle Spinning (MAS) NMR


For a spin 1/2 nucleus, the CSA and the dipolar coupling significantly broaden the solids lineshape, even in crystalline
compounds. By inspecting the Hamiltonians of each interaction, the 3cos(θ) − 1 term appears. If this term was set to zero the
dipolar coupling would go to zero. The CSA would only be dependent on ϕ and η . Rapid rotation about this angle would
average all ϕ leading to a purely isotropic lineshapes. This is the logic behind (MAS) NMR. In this experiment, the sample is
physically rotated around the magic angle at very high spinning speeds, on the order of kHz. The magic angle can be solved
for numerically by,
3 cos θ − 1 = 0 (1)

θ = 54.74 = Magic Angle (2)

The chemical shift can be related to the individual orientation of the crystal in the rotor to the PAS using Euler angles, α, β, γ.
This conversion gives a chemical shift of

−ω0 σiso + [ A1 cos(ωr t + γ) + B1 sin(ωr t + γ)] + [ A2 cos(2 ωr t + 2γ) + B2 sin(ωr t + 2γ] (3)

where
2 – 2 2
P AS P AF P AS P AF
A1 = √2 sin βcosβ[ cos α(σxx − σzz ) + sin α(σyy − σzz )] (4)
3

2 – P AS P AF
B1 = √2 sin α cos α sin β(σxx − σzz ) (5)
3

1
2 2 2 P AS P AF 2 2 2 P AS P AF
A2 = ((cos β cos α − sin α)(σxx − σzz ) + (cos β sin α − cos α)(σyy − σzz )) (6)
3

2
P AS P AF
B2 = − sin α cos α cos β(σxx − σzz ) (7)
3

This expression is useful when considering incomplete averaging of the CSA by MAS. Incomplete averaging of the CSA
occurs when the spinning frequency is less than the size of the CSA. The first term of this equation is the isotropic component
and the remaining terms are all dependent on the spinning frequency ω . When the CSA averaging is incomplete, these terms
r

describe rotation echoes observed in the FID. Rotational echoes form as the crystallite moves with the rotor. Due to the
rotation, the orientation of the crystallite changes which changes the shielding around the nucleus. Once the sample returns to
its original starting position, the next rotation will have the crystallite sampling the same orientation for the same time,
imposing a separate FID on the FID. The additional FIDs, denoted by the red arrows, are known as rotational echoes and are
denoted as sharp spike in the time dimension, shown in the figure below. Rotational echoes are easy to identify as they appear
in the FID as sharp spikes spaced every . When the sample is spinning faster than the anisotropy, the shielding lags behind
1

ωr

the spinning speed and the crystallite shielding will not experience the same set of values over the same time period prevention
rotational echo formation.

Figure 9. Rotational echoes in the FID are denoted by the red arrows. After a FT the rotational echoes become equally spaced
spinning sidebands at integer multiples of the spinning speed (8kHz).

Derrick Kaseman 6/18/2021 1 https://chem.libretexts.org/@go/page/1814


Figure 10. Effect of spinning speed on the CSA ( Δ = 200ppm )of 77Se.
Fourier transforming this results in characteristic spinning sidebands which appear at integer multiples of the spinning speed.
Spinning sideband patterns, shown above, provide information about the CSA and are therefore helpful to have when
performing structural studies. In this figure, δ = 200ppm and η = 0.5 .However, if a compound has multiple sites all with
large CSA values, the spectrum becomes complicated and the isotropic peaks cannot be identified a priori. To identify the
isotropic positions, a Herzfeld Berger analysis of the sideband patters may be performed. In a Herzfeld Berger analysis, spectra
are collected at different spinning speeds. The more spinning speeds at which the spectrum is collected, the easier the
assignment of the peaks. The sidebands appear at integer multiples of the rotor speed, while the isotropic peaks remain
constant. Thus to find the peaks you only need to locate those peaks that do not move when the spinning speed is changed. The
relative intensities of the isotropic peaks will not change, while the sideband patterns will change intensities depending on the
spinning speed. However, an isotropic peak may be overlapping with a sideband, giving artificial intensity to that peak.
As alluded to earlier, the CSA parameters may be derived from the sideband patterns. When doing this one must consider a
few things:
1. The dipolar coupling. If the spinning speed is not large enough to average the dipolar coupling between nuclei this can give
additional sidebands which when fit lead to inaccurate determination of CSA
2. At least 4 sidebands in order to have enough degrees of freedom to accurately fit the chemical shift tensor. The chemical
shift tensor is 3x3 matrix. If the matrix is diagonalized, there are 6 remaining values; 3 isotropic and 3 anisotropic. The
anisotropy may be fit using 3 values in which case you need 3+1 degrees of freedom to fit accurately.
Below is a typical pulse sequence used for MAS experiments. It is sometimes called pulse and acquire, as only a pulse is π

used and the spectrum is immediately acquired. The pulse doesn't need to be . In fact, normally a pulse with a smaller tip
π

angle is used in order to acquire more scans in a shorter time. The optimum tip angle at which the signal is maximized at the
shortest time is called the Ernst angle.

Figure 11. Pulse sequence for MAS experiment

Hahn Echo
Derrick Kaseman 6/18/2021 2 https://chem.libretexts.org/@go/page/1814
The Hahn echo sequence is named after Erwin Hahn, who in the 1950's discovered that by applying a pulse with one phase
and then another pulse with a different phase, the signal would refocus and form what is known as an echo. There are many
different types of echoes including solids echoes, Solomon echoes, and Hahn echoes. The Hahn echo is particularly important
for spin =1/2 nuclei. By applying a pulse waiting a set delay and then pulsing with a π pulse results in an echo forming. The
π

delay between the and the π pulses will be the time at which the echo reaches its maximum value, as shown in the pulse
π

sequence below. In this experiment the delay is an integer multiple of the rotor period, or rotor-synchronized. This
synchronization is important in order to retain any information about the CSA. Obviously if the rotor is no longer rotor
synchronized the CSA will refocus at a different time yielding different (fake) interactions. The Hahn echo is often used to
shift the signal out in time to negate any pulse ringdown effects that may add artificial broadening to the spectrum.

Figure 12. Pulse sequence for Hahn echo experiment


To understand how an echo forms an analogy to runners on a track (shown below) is used. Initially, all of the runners (circles)
are at the starting line. Each runner (different colored circle) represents one spin in the ensemble. After the race starts, the
runners begin to race around the track, separating based on speed (arrows). This is equivalent to T2 relaxation in the NMR
system. However, as theses runners begin to lose their coherence a whistle is blown and the runners all turn around and run
back toward the starting line at EXACTLY the same speed as before. When they reach the starting line they are all in phase
again. This is the echo that is seen in the NMR experiment. Thus the Hahn echo can be used to measure T2 processes.

Figure 13. Graphic depiction of how a Hahn echo forms. Initially, each runner (spin) is at the starting line (in phase). After
some time the runners begin to separate based on their speed (spins precess at slightly different speed). At some time a whistle
is blown and the runner turn and race back to the start at exactly the same speed (π pulse). After the same amount of time the
runners come back to the starting line (echo formation).

Solids Echo
In quadrupolar nuclei, the refocusing of the spins into echoes becomes challenging as there are multiple energy levels and
certain transitions are forbidden Δm ≠ ±1 . Consequently, treating the quadrupolar nuclei with the Hahn echo sequence
results in incomplete refocusing (Some of the runners do not hear the whistle and turn around). This leads to different
interactions which need to be discussed more thoroughly. However, if instead a pulse is used instead of the π pulse, the
π

central transitions will experience the spin flip and reform a solids (quadrupolar) echo. The main advantage of this type of
echo is that it includes no broadening to do satellite transitions and is therefore more quantitative than a Hahn echo for
quadrupolar nuclei.

Derrick Kaseman 6/18/2021 3 https://chem.libretexts.org/@go/page/1814


Figure 14. Solids echo pulse sequence

Multiple Quantum MAS (MQMAS)

Figure 16. MQMAS pulse sequence


MQMAS is a high resolution 2D NMR technique that allows for an isotropic spectrum of a quadrupolar nucleus to be
detected. Frydman and Harwood showed that by manipulating spin coherences using pulses, the second and fourth rank
Legendre Polynomials can be averaged over the entire experiment. Separating the experiment into 2 time domains t1 and t2
and setting the Legendre polynomials to zero yields
C4 (I , m1 )t1 + C4 (I , m2 )t2 = 0 (8)

.
Experimentally, t1 can be chosen and the resulting echo will form at t2. However, m2 is fixed as the final transition needs to be
1/ so the signal can be detected. Solving for t yields:
2 2

t2 = kt1 (9)

C4 (I , m1 )
k = (10)
1
C4 (I , )
2

In order to complete this experiment, the MQ transition must first be excited. This can be accomplished using a very short π

12

pulse. After waiting t1, the spins are then reconverted into the central transition so they can be detected. The resultant echo
consists of spins which have satisfied the averaging of the fourth rank polynomials and therefore only contain isotropic
information. Next, t1 is varied and the resultant echo is measured. The signal then evolves on a diagonal (k) which is then
sheared an FT in both dimensions giving a v1' correlated with the isotropic and anisotropic components.

Figure 17. Processing of the raw MQMAS data. After a 2D FT and shear the isotropic and anisotropic components of the
signal can be correlated. Initially the signal evolves on a diagonal proportional to k.
in this figure
34 νiso − 60 ν0
ν1 = (11)
9

and
ν2 = νiso + 3 ν0 − 21 ν4 (θ, ϕ) (12)

Two Dimensional Phase Adjusted Spinning Sidebands (2D PASS)


Derrick Kaseman 6/18/2021 4 https://chem.libretexts.org/@go/page/1814
Figure 18. 2D PASS pulse sequence. The timings between the pulses are modulated to satisfy the PASS equations.

The 2D PASS experiment is an experiment which separates the CSA from the isotropic chemical shift. The pulse sequence
consists of a pulse followed by a train of 5 π pulses which are spaced at timings given by the PASS equations. The PASS
π

equations are complex and will not be discussed further. Timings for the pulses may be found in the paper published by
Antzutkin et al. in 1995.
The goal of this experiment is to separate the CSA into a separate dimension from the isotropic chemical shift. The basic
principle is that the CSA may be separated out by changing the timings between the pi pulses. The change in the timings
allows for the isotropic part of the chemical shift to evolve as a function of the timings and the CSA is directly acquired in
each 1D slice. The two are then related to one another by a shearing transformation. The data acquisition and evolution of the
CSA and isotropic chemical shift are shown in the figure below.

Figure 19. Processing of the 2D data. After a 2D FT and a shearing transformation, the anisotropic and isotropic components
of the CSA can be related.

Dynamic Angle Spinning (DAS)


Unlike MQMAS which manipulated the spin of the nuclei, DAS manipulates the orientation of the spins in the rotor. As can be
seen in the quadrupolar equations, no one angle can allow for complete averaging the of the quadrupolar interactions.
However, solving these equations individually leads to sets of angles (left as an exercise for the reader).
Several techniques have been developed in order to exploit the different angles.
Double Rotation: This technique places a smaller rotor oriented at a different orientation inside a rotor which spins at the
magic angle. The rotor inside also spins, which averages 2 of the above interactions. This is an extremely challenging
experiment and the inner rotor is only able to spin at a max of 3 kHz.
Magic Angle Flipping: In this technique, the rotor is physically reoriented by a stepper motor during the experiment.
Commonly, the experiment is performed at the magic angle and then is reoriented to 90 degrees to reintroduce the anisotropy.
In this way the anisotropy can be correlated to the isotropic chemical shift, much like the 2D PASS experiment. However,
when using a quadrupolar nucleus, the sample can be reoriented to another "magic" angle which will average the remainder of
the quadrupolar interaction. In order for a successful experiment τ , the time it takes to switch the rotor must be shorter than T1
or no signal will be observed.

Derrick Kaseman 6/18/2021 5 https://chem.libretexts.org/@go/page/1814


Figure 20. Pulse sequence and rotor position for the DAS experiment.

References
MAS References
1. Fyfe,C. A., Massbruger, H., Yannoni, C.S. J. Mag. Res. 36 (1979) 61-68 doi:10.1016/0022-2364(79)90215-4
2. Herzfled, J., Berger, A.E., J. Chem Phys. 73 (1980) 6021-6030 doi:10.1063/1.440136

MQMAS References
1. Goldbourt, A., Madhu, P., Chemical Monthly. 133 (2002) 1497-1534. doi:10.1007/s00706-002-0502-y
2. Medek, A., Frydman, L., J. Braz. Chem Soc. 10 (1999) 263-277. doi:10.1021/ja00156a015
3. Frydman, L., Harwood , J.S., J. Am.Chem. SOc. 117 (1995) 5367 ISSN:0103-5053

2D PASS References
1. Dixon, W.T., J. Chem. Phys. 77 (1982) 1800-1809
2. Antzutkin, O. N., Shekar, S.C., Levitt, M. H., J. Magnetic Rsonance A. 115 (1995) 7-19 doi:10.1006/jmra.1995.1142
3. Kaseman, D. C., Hung, I., Gan, Z., Sen, S. J. Phys. Chem. B. 117 (2013)949-954 doi:10.1021/jp311320t
4. Kaseman, D. C. Endo, T. Sen, S. J. Non-Cryst. Solids. 359 (2013) 33-39 doi:10.1016/j.jnoncrysol.2012.09.030
5. Hung, I., Edwards, T, Sen, S., Gan, Z., J. Mag Res. 221 (2012) 103-109 doi:10.1016/j.jmr.2012.05.013
6. Fayon, F., Bessada, C., Douy, A. Massiot, D. J. Mag. Res. 137 (1999) 116-121 doi:10.1006/jmre.1998.1641
7. Davis, M. C., Shookman, K. M., Sillaman, J. D., Grandinetti, P. J., J Mag Res. 210 (2011) 51-58
doi:10.1016/j.jmr.2011.02.008

Echo References
1. Anderson, A. G., Garwin, R. L., Hahn, E.L., Horton, J.W., Tucker, G.L., Walker, R.M., J. Appl. Phys. 26 (1955) 1324-1338
doi:10.1063/1.1721903
2. Hahn, E. L., Phys. Rev. 80 (1950) 580-594 doi:10.1103/PhysRev.80.580 doi:10.1103/PhysRev.93.639
3. Hahn, E. L., Herzog, B., Phys. Rev. 93 (1954) 639-640 doi 10.1103/PhysRev.93.639
4. Hahn, E. L., Maxwell, D. E., Phys Rev. 84 (1951) 1246-1247 doi:10.1103/PhysRev.84.1246
5. Hahn, E. L., Maxwell, D. E., Phys Rev 88 (1952) 1070-1084 doi:10.1103/PhysRev.88.1070
6. Rowan, L. G., Hahn, E. L., Mims, W. B., Phys Rev. 137 (1965) A61 doi:10.1103/PhysRev.137.A61

CPMG References
1. Carr, H. Y., Purcell, E. M., Phys Rev. 94 (1954) 630-638 doi:10.1103/PhysRev.94.630
2. Meiboom, S., Gill, D., Review of Scientific Instruments. 29 (1958) 688-691 doi:10.1063/1.1716296
3. Hung, I., Edwards, T, Sen, S., Gan, Z., J. Mag Res. 221 (2012) 103-109 doi:10.1016/j.jmr.2012.05.013
4. Baltisberger, J. H., Waler, B. J., Keeler, E. G., Kaseman, D. C., Sanders, K. J., Grandinetti, P. J., J Chem Phys. 136 (2012)
doi:10.1063/1.4728105

DAS References
1. Chmelka, B.F., Mueller, K. T., Pines, A., Stebbins, J., Wu, Y., Zwanziger, J. W., Nature. 339 (1989) 42-43
doi:10.1038/339042a0
2. Eastman, M. A., Grandinetti, P.J., Lee, Y. K., Pines, A., J. Mag. Res. 98 (1990) 333-341 doi:10.1016/0022-2364(92)90136-
u
3. Farnan, I., Grandinetti, P. J., Baltisberger, J. H., Stebbines, J. H., Werner, U., Eastman, M.A., Pines, A., Nature. 358 (1992)
31-35 doi:10.1038/358031a0

Derrick Kaseman 6/18/2021 6 https://chem.libretexts.org/@go/page/1814


4. Mueller, K. T., Baltisberger, J. H., Wooten, E. W., Pines, A., J. Phys Chem. 96 (1992) 7001-7004 doi:10.1021/j100196a028
5. Mueller, K. T., Chingas, G.C., Pines, A., Review Scientific Instruments. 62 (1991) 1445-1452 doi:10.1063/1.1142465
6. Meuller, K. T., Sun, B.Q., Chingas, G. C., Zwanziger, J. W., Terao, T., Pines, A. J Mag. Res. 86 (1990) 470-487
doi:10.1016/0022-2364(90)90025-5
7. Meuller, K.T., Wooten, E.W., Pines, A., J. Mag. Res, 92 (1991) 620-627 doi:10.1016/0022-2364(91)90359-2

Contributors
Derrick Kaseman (UC Davis)

Derrick Kaseman 6/18/2021 7 https://chem.libretexts.org/@go/page/1814


Car-Purcell-Meiboom-Gill (CPMG) Echo Train Acquisition
CPMG is a NMR technique commonly employed to enhance the S/N of an NMR spectrum

Introduction
CPMG is a very useful technique for SSNMR spectroscopists. It uses a train of π pulses to refocus inhomogeneous broadening
of the nuclear spins. This is very beneficial and can be used to enhance S/N, measure diffusion, measure T2 processes, and
reduce experimental time. CPMG was initally developed by Carr and Purcell, at Rutgers and Harvard, respectivley. Their work
was added on to by Meiboom and Gill, resulting in a pulse sequence bearing each of their names.

Pulse Sequence
Like the Hahn echo, a π pulse is placed after the last pulse in the NMR experiment, which refocuses the spins leading to echo
formation. For CPMG echo train acquisition, M π pulses are applied every 2ntr (if the sample is spinning), resulting in M
echoes. The number of echoes which can be acquired is directly related to T2 processes. Neglecting pulse imperfections, the
echo tops will diminish in intensity due to coherence losses between spins, which is homogenous T2, as the \(\pi) pulses
refocus the inhomogeneous T2 due to the varying magnetic field experienced by the sample.

Figure 15. CPMG Pulse Sequence

CPMG Signal
CPMG Sidebands
Truncation

Data Processing
After acquisition, the issue becomes the processing of the data. There are two schools of thoughts in this. First, some people
directly FT the data. This leads to a spiklet pattern in which the spikes are due to the echoes which are spaced at 1/2ntr. The
spikelets map out the NMR spectrum. There are some debate as to whether these spikelet patterns are arbitrarily increasing the
signal to noise. The other school of though is to stack the echoes on top of one another. This then adds all the signals together
and one spikelet-less spectrum is obtained with the S/N increased by the number of echoes obtained.

7/4/2021 1 https://chem.libretexts.org/@go/page/1815
Magic Angle Hopping (MAH)
Upon the advent of the MAS experiment, solid state NMR became particularly powerful because line shapes that were
convoluted and overlapping due to the CSA were able to be resolved. While advantageous, the MAS, assuming the same
rotation >> CSA completely averages the CSA, which destroys potentially useful information. The other limitation of MAS is
when the CSA>the sample rotation resulting in sideband formation. The Magic Angle Hopping experiment was developed by
Bax et al to separate the isotropic contributions of the chemical shift while preserving the anisotropic contributions in a
separate dimension. This is accomplished by rotating the sample in successive 120 increments about the magic angle.

Pulse Sequence
The basic pulse sequence is shown below. It consists of 5 90 pulses followed by acquisition.

Figure 1. The Magic Angle Hopping Experiment. The experiment consists of 3 equal evolution segments and 2 projection
segments, where the rotor is moved while magnetization is stored about the z-axis.
Initially, the bulk magnetization M is along the external magnetic field direction, here the z-axis. This is denoted in the pulse
0

sequence and vector magnetization diagram (below) as 1. The 90 pulse rotates the magnetization into the x-y plane. (If the

t1
pulse is of phase "-y" then the magnetization is along the x-axis!) Next, (2) the magnetization is allowed to process for . The
3

angle which the M rotates is α . Application of a 90 pulse rotates the magnetization along the z-axis the rotor (depicted by
0

the red circle) is then incremented 120 (3). Repeating the process of steps 1-3 allows for 1 more evolution period and rotation

of 120 . The rotor is now 240 from the initial starting position. Finally, the magnetization evolves for one more evolution
∘ ∘

period until it is detected (6).

7/4/2021 1 https://chem.libretexts.org/@go/page/16966
Figure 2. Vector diagram of the bulk magnetization M and the rotor position. M is shown in blue, while the rotor is the red
0 0

circle. The black line shows the evolution of the rotor as a function of the time elapsed during the MAH experiment. The
numbers correspond to the numbers in the pulse sequence in Figure 1.
The MAH hopping experiment needs 4 experiments per t1 increment in order to completely average the CSA contributions and
obtain a pure isotropic dimension. Additionally, the MAH experiment is a 2D experiment in which t1 is incremented to create
the second dimension.

Mathematical Treatment of Magnetization Evolution


Lets consider the picture of the vector magnetization. Initially, the magnetization lies along the z-axis (1). Application of a
− y pulse places the bulk magnetization, M0 along the x-axis, where it evolves for a period of time (2). The precession in
π

the x-y plane is dependent on the frequency of the individual spin for the given sample orientation and position in the rotor,
and the amount of time. More simply,
t1
α = ωa (1)
3

The magnetization is the placed along the z-axis, forcing the transverse magnetization to de-phase and the rotor is incremented
120 . The spin that was initially in position A is now in position B, rotated 120 about the magic angle (3). Repeating the
∘ ∘

same process described above we now obtain (4-6)


t1
β = ωb (2)
3

and
t1
γ = ωc (3)
3

At the end of t1, the x component of the magnetization is

7/4/2021 2 https://chem.libretexts.org/@go/page/16966
M0 cos α cos β cos γ (4)

and the total magnetization in the x-y plane becomes


t1
M1 (t1 =)M0 cos α cos βexp(i ωc ) (5)
3

Next the phases of the pulses are incremented in such a way the following magnetization are obtained during t1
t1
M2 (t1 ) = i M0 sin α cos βexp(i ωc ) (6)
3

t1
M3 (t1 ) = i M0 cos α sin βexp(i ωc ) (7)
3

t1
M4 (t1 ) = −M0 sin α sin βexp(i ωc ) (8)
3

Using Eulers formula summation of the above equations results in


t1
M0 exp(i ωa + ωb + ωc ) (9)
3

where
ωa + ωb + ωc T rσ (σ11 + σ22 + σ33 )
= = = ωiso (10)
3 3 3

Thus the signal only contains the isotropic shielding resonance frequencies. Consider the sum of the signal to include both t1
and t2, the magnetization is
M = M0 exp(i ωi t1 )exp(i ωc t2 )exp(−[ t1 + t2 ]/ T2 ) (11)

Data Analysis
The CSA components still evolve during t2, which correspond to the static sample. Consequently CSA information is
contained in the second dimension. The spectral width of the second dimension is defined by the 1/t1 increment, while the
number of points in the second dimension is the number of t1 points. The spectrum below is what one obtains after Fourier
Transforming in both dimensions.

7/4/2021 3 https://chem.libretexts.org/@go/page/16966
Figure 3. Representative spectrum of a MAH experiment. Projections of the isotropic and anisotropic dimensions are shown
on the right and left, respectively of the 2D plot. The anisotropic line shapes for each site are shown above the anisotropic
projection.
Each CSA pattern may be extracted and simulated to extract the CSA tensor parameters, while fitting the isotropic spectrum
will give the relative fractions of each site.

References
1. Bax, A., Szeverenyi, M., Maciel, G. E. Journal of Magnetic Resonance. 52 147-152 (1983)

Contributors and Attributions


Derrick C. Kaseman (UC Davis)

7/4/2021 4 https://chem.libretexts.org/@go/page/16966
Solid State NMR Experimental Setup

7/7/2021 1 https://chem.libretexts.org/@go/page/1816
NMR: Structural Assignment

(n+1) Rule

Background to C-13 NMR

Determine Structure with Combined Spectra

High Resolution Proton NMR Spectra


This page describes how you interpret simple high resolution nuclear magnetic resonance (NMR) spectra. It assumes
that you have already read the background page on NMR so that you understand what an NMR spectrum looks like and
the use of the term "chemical shift". It also assumes that you know how to interpret simple low resolution spectra.

Integration in Proton NMR

Interpreting C-13 NMR Spectra

Introduction to Proton NMR

Low Resolution Proton NMR Spectra

More About Electronics

Multiplicity in Proton NMR

NMR11. More About Multiplicity

6/29/2021 NMR.1 https://chem.libretexts.org/@go/page/7821


NMR14. More Practice with NMR Spectroscopy

NMR2. Carbon-13 NMR

NMR3. Symmetry in NMR

NMR4. 13C NMR and Geometry

NMR5. 13C NMR and Electronics

NMR8. Chemical Shift in 1H NMR

NMR Appendix. Useful Charts for NMR identification

6/29/2021 NMR.2 https://chem.libretexts.org/@go/page/7821


(n+1) Rule
The (n+1) Rule, an empirical rule used to predict the multiplicity and, in conjunction with Pascal’s triangle, splitting pattern of
peaks in 1H and 13C NMR spectra, states that if a given nucleus is coupled (see spin coupling) to n number of nuclei that are
equivalent (see equivalent ligands), the multiplicity of the peak is n+1. eg. 1:

The three hydrogen nuclei in 1, H , H , and H , are equivalent. Thus, 1H NMR spectrum of 1 H s only one peak. H , H ,
a b c a a b

and H are coupled to no hydrogen nuclei. Thus, for H , H , and H , n=0; (n+1) = (0+1) = 1. The multiplicity of the peak of
c a b c

H , H , and H is one. The peak H s one line; it is a singlet. eg. 2:


a b c a

There are two sets of equivalent hydrogen nuclei in 2:


Set 1: H a

Set 2: H , H
b c

Thus, the 1H NMR spectrum of 2 H s two peaks, one due to H and the other to H and H .
a a b c

The peak of H : There are two vicinal hydrogens to H : H and H . H and H are equivalent to each other but not to H .
a a b c b c a

Thus, for H , n=2; (n+1) = (2+1) = 3. The multiplicity of the peak of H is three. The peak H s three lines; from the Pascal’s
a a a

triangle, it is a triplet.
The peak of H and H : There is only one vicinal hydrogen to H and H : H . H is not equivalent to H and H . Thus, for
b c b c a a b c

H and H , n=1; (n+1) = (1+1) = 2. The multiplicity of the peak of H and H is two. The peak H s two lines, from the
b c b c a

Pascal’s triangle, it is a doublet.


To determine the multiplicity of a peak of a nucleus coupled to more than one set of equivalent nuclei, apply the (n+1) Rule
independently to each other.
eg:

7/7/2021 1 https://chem.libretexts.org/@go/page/14970
There are three set of equivalent hydrogen nuclei in 3:
Set 1: H a

Set 2: H b

Set 3: H c

peak of H :a

multiplicity of the peak of H = 2 × 2 = 4 . To determine the splitting pattern of the peak of


a Ha , use the Pascal’s triangle,
based on the observation that, for alkenyl hydrogens, J > J
cis . gem

The peak of H is a doublet of a doublet.


a

peak of H :b

7/7/2021 2 https://chem.libretexts.org/@go/page/14970
multiplicity of the peak of H = 2 × 2 = 4 . To determine the splitting pattern of the peak of
b Hb , use the Pascal’s triangle,
based on the observation that, for alkenyl hydrogens, J >J
trans .
gem

The peak of H is a doublet of a doublet.


b

peak of H :c

multiplicity of the peak of H = 2 × 2 = 4 . To determine the splitting pattern of the peak of


c Hc , use the Pascal’s triangle
based on the observation that, for alkenyl hydrogens, J >J
trans . cis

The peak of H is a doublet of a doublet.


c

Contributors and Attributions


Gamini Gunawardena from the OChemPal site (Utah Valley University)

7/7/2021 3 https://chem.libretexts.org/@go/page/14970
Background to C-13 NMR
This page describes what a C-13 NMR spectrum is and how it tells you useful things about the carbon atoms in
organic molecules.

The background to C-13 NMR spectroscopy


Nuclear magnetic resonance is concerned with the magnetic properties of certain nuclei. On this page we are
focussing on the magnetic behavior of carbon-13 nuclei.

Carbon-13 nuclei as little magnets


About 1% of all carbon atoms are the C-13 isotope; the rest (apart from tiny amounts of the radioactive C-14) is C-
12. C-13 NMR relies on the magnetic properties of the C-13 nuclei. Carbon-13 nuclei fall into a class known as
"spin ½" nuclei for reasons which don't really need to concern us at the introductory level this page is aimed at (UK
A level and its equivalents). The effect of this is that a C-13 nucleus can behave as a little magnet. C-12 nuclei
don't have this property.
If you have a compass needle, it normally lines up with the Earth's magnetic field with the north-seeking end
pointing north. Provided it isn't sealed in some sort of container, you could twist the needle around with your fingers
so that it pointed south - lining it up opposed to the Earth's magnetic field. It is very unstable opposed to the Earth's
field, and as soon as you let it go again, it will flip back to its more stable state.

Because a C-13 nucleus behaves like a little magnet, it means that it can also be aligned with an external magnetic
field or opposed to it.
Again, the alignment where it is opposed to the field is less stable (at a higher energy). It is possible to make it flip
from the more stable alignment to the less stable one by supplying exactly the right amount of energy.

The energy needed to make this flip depends on the strength of the external magnetic field used, but is usually in
the range of energies found in radio waves - at frequencies of about 25 - 100 MHz. (BBC Radio 4 is found between
92 - 95 MHz!) If you have also looked at proton-NMR, the frequency is about a quarter of that used to flip a
hydrogen nucleus for a given magnetic field strength.
It's possible to detect this interaction between the radio waves of just the right frequency and the carbon-13
nucleus as it flips from one orientation to the other as a peak on a graph. This flipping of the carbon-13 nucleus
from one magnetic alignment to the other by the radio waves is known as the resonance condition.
The importance of the carbon's environment
What we've said so far would apply to an isolated carbon-13 nucleus, but real carbon atoms in real bonds have
other things around them - especially electrons. The effect of the electrons is to cut down the size of the external

Jim Clark 6/20/2021 1 https://chem.libretexts.org/@go/page/3736


magnetic field felt by the carbon-13 nucleus.

Suppose you were using a radio frequency of 25 MHz, and you adjusted the size of the magnetic field so that an
isolated carbon-13 atom was in the resonance condition.
If you replaced the isolated carbon with the more realistic case of it being surrounded by bonding electrons, it
wouldn't be feeling the full effect of the external field any more and so would stop resonating (flipping from one
magnetic alignment to the other). The resonance condition depends on having exactly the right combination of
external magnetic field and radio frequency.
How would you bring it back into the resonance condition again? You would have to increase the external magnetic
field slightly to compensate for the shielding effect of the electrons.
Now suppose that you attached the carbon to something more electronegative. The electrons in the bond would be
further away from the carbon nucleus, and so would have less of a lowering effect on the magnetic field around the
carbon nucleus.

The external magnetic field needed to bring the carbon into resonance will be smaller if it is attached to a more
electronegative element, because the C-13 nucleus feels more of the field. Even small differences in the
electronegativities of the attached atoms will make a difference to the magnetic field needed to achieve resonance.

Summary
For a given radio frequency (say, 25 MHz) each carbon-13 atom will need a slightly different magnetic field applied
to it to bring it into the resonance condition depending on what exactly it is attached to - in other words the
magnetic field needed is a useful guide to the carbon atom's environment in the molecule.

Features of a C-13 NMR spectrum


The C-13 NMR spectrum for ethanol
This is a simple example of a C-13 NMR spectrum. Don't worry about the scale for now - we'll look at that in a
minute.

There are two peaks because there are two different environments for the carbons. The carbon in the CH3 group is
attached to 3 hydrogens and a carbon. The carbon in the CH2 group is attached to 2 hydrogens, a carbon and an
oxygen. The two lines are in different places in the NMR spectrum because they need different external magnetic
fields to bring them in to resonance at a particular radio frequency.

Jim Clark 6/20/2021 2 https://chem.libretexts.org/@go/page/3736


The C-13 NMR spectrum for a more complicated compound
This is the C-13 NMR spectrum for 1-methylethyl propanoate (also known as isopropyl propanoate or isopropyl
propionate).

This time there are 5 lines in the spectrum. That means that there must be 5 different environments for the carbon
atoms in the compound. Is that reasonable from the structure?

Well - if you count the carbon atoms, there are 6 of them. So why only 5 lines? In this case, two of the carbons are
in exactly the same environment. They are attached to exactly the same things. Look at the two CH3 groups on the
right-hand side of the molecule.
You might reasonably ask why the carbon in the CH3 on the left isn't also in the same environment. Just like the
ones on the right, the carbon is attached to 3 hydrogens and another carbon. But the similarity isn't exact - you
have to chase the similarity along the rest of the molecule as well to be sure.
The carbon in the left-hand CH3 group is attached to a carbon atom which in turn is attached to a carbon with two
oxygens on it - and so on down the molecule.
That's not exactly the same environment as the carbons in the right-hand CH3 groups. They are attached to a
carbon which is attached to a single oxygen - and so on down the molecule.
We'll look at this spectrum again in detail on the next page - and look at some more similar examples as well. This
all gets easier the more examples you look at.
For now, all you need to realise is that each line in a C-13 NMR spectrum recognises a carbon atom in one
particular environment in the compound. If two (or more) carbon atoms in a compound have exactly the same
environment, they will be represented by a single line.

The need for a standard for comparison - TMS


Before we can explain what the horizontal scale means, we need to explain the fact that it has a zero point - at the
right-hand end of the scale. The zero is where you would find a peak due to the carbon-13 atoms in
tetramethylsilane - usually called TMS. Everything else is compared with this.

You will find that some NMR spectra show the peak due to TMS (at zero), and others leave it out. Essentially, if you
have to analyse a spectrum which has a peak at zero, you can ignore it because that's the TMS peak.
TMS is chosen as the standard for several reasons. The most important are:
It has 4 carbon atoms all of which are in exactly the same environment. They are joined to exactly the same
things in exactly the same way. That produces a single peak, but it's also a strong peak (because there are lots
of carbon atoms all doing the same thing).

Jim Clark 6/20/2021 3 https://chem.libretexts.org/@go/page/3736


The electrons in the C-Si bonds are closer to the carbons in this compound than in almost any other one. That
means that these carbon nuclei are the most shielded from the external magnetic field, and so you would have
to increase the magnetic field by the greatest amount to bring the carbons back into resonance.
The net effect of this is that TMS produces a peak on the spectrum at the extreme right-hand side. Almost
everything else produces peaks to the left of it.

The chemical shift


The horizontal scale is shown as (ppm). is called the chemical shift and is measured in parts per million - ppm.
A peak at a chemical shift of, say, 60 means that the carbon atoms which caused that peak need a magnetic field
60 millionths less than the field needed by TMS to produce resonance. A peak at a chemical shift of 60 is said to
be downfield of TMS. The further to the left a peak is, the more downfield it is.

Solvents for NMR spectroscopy


NMR spectra are usually measured using solutions of the substance being investigated. A commonly used solvent
is CDCl3. This is a trichloromethane (chloroform) molecule in which the hydrogen has been replaced by its isotope,
deuterium.
CDCl3 is also commonly used as the solvent in proton-NMR because it doesn't have any ordinary hydrogen nuclei
(protons) which would give a line in a proton-NMR spectrum. It does, of course, have a carbon atom - so why
doesn't it give a potentially confusing line in a C-13 NMR spectrum? In fact it does give a line, but the line has an
easily recognizable chemical shift and so can be removed from the final spectrum. All of the spectra from the
SDBS have this line removed to avoid any confusion.

Contributors and Attributions


Jim Clark (Chemguide.co.uk)

Jim Clark 6/20/2021 4 https://chem.libretexts.org/@go/page/3736


Determine Structure with Combined Spectra
There are a many ways we can use NMR spectroscopy to analyse compounds. One common application is in determination of
an unknown structure. Given the MS, IR, 13C and 1H NMR spectra, what might be the structure of an unknown sample?
It is often easiest to start with the IR spectrum.
identify at least three peaks in the IR spectrum. Which peaks seem to tell you the most information about this compound?
don't think with your head; think with your hands. Write down ideas on the spectrum.
if you are working on a formal proof of structure, on a class test or a lab report, you may be required to enter your data in a
table correlating wavenumber with peak assignment:

cm-1 asst

For example, a student might obtain the following IR spectrum.

From that information, she constructs the following table. She might even write this table, by hand, directly on her spectrum.
She makes useful notes on the edges, and might even include some guesses, which she later crosses out, but does not erase.
She is assisted in this task by consulting an IR table, that suggests what some of these peaks might mean.

Remember:
make special note of what atoms are present in the compound: C, H, N, O...
also note your initial ideas about specific functional groups that may be present.
if you are unsure of an assignment, put a question mark beside it to signal this uncertainty.
some data may need to be discarded later if it is not consistent with other data.
Look at the 13C spectrum.
How many different carbons are there, based on the number of peaks in the spectrum? This is the first step in estimating
the molecular formula.
Do you have reason to believe there is symmetry in the structure? In the entire compound or just part of it? Adjust the
number of carbons you think you are dealing with.
As in IR spectroscopy, begin assigning peaks, either on the spectrum or, if required, in a table:

ppm asst

For example, a student might obtain the following 13C NMR spectrum:

From that information, she puts together the following table:

Remember:
you will be able to assign all peaks in the NMR spectrum, not just a few like in IR.
You can get more information on the formula from the 1H NMR spectrum and the mass spectrum.
In the 1H NMR spectrum, what does the sum of the integrations suggest about the number of hydrogens?
At this point you may have a formula, CxHyOz. What would be the mass of a compound with this formula?
Compare this mass to the mass spectrum. Does it match?

Chris Schaller 6/5/2021 1 https://chem.libretexts.org/@go/page/4190


If not, consider whether a common atom (such as oxygen) is missing from your formula, or if there might be symmetry that
you have missed.
Does the mass spectrum suggest the presence of a nitrogen (odd molecular weight) or chlorine or bromine (isotope
pattern)?
Once you have the molecular formula, the number of possible structures is automatically limited. The number of units of
unsaturation can help you narrow down the possibilities.
The ratio of C : H in a saturated, acyclic hydrocarbon is n : 2n+2.
Each pair of H missing from the formula corresponds to a multiple bond or a ring.
The presence of oxygen (which is divalent) does not alter the C : H ratio.
The presence of halogen (which is monovalent) means there is one H missing.
The presence of a nitrogen (which is trivalent) means there is an extra H in the formula.
In other words, C4H8O has one ring or double bond just like C4H8, and so do C4H7Br and C4H9N.
As in 13C NMR, you should be able to assign all peaks in the 1H NMR spectrum. You may be able to do so by making notes
on the spectrum. If you think you know the structure, you may be able to draw it and note which peak belongs with which
proton.
A formal proof of structure might require a table of assignments.

ppm int mult partial structure assignment

This table demonstrates your ability to read the spectrum. Can you decide what ratio of protons is suggested by the integral
line? Can you decide whether a peak is a quartet?
The partial structure column should explain the shift, integration and multiplicity for the peak in that row. It should not
show any other information from elsewhere in the structure. This restriction forces you to demonstrate a thorough
understanding of the data in a way that "getting the right answer" does not.
The partial structure column is best filled in with drawings, not words. The drawing is a partial structure.
Because the partial structure will show the protons absorbing at the shift in that row as well the neighbouring protons, you
need to distinguish between them in your picture. Most people circle or underline or make bold the protons that show up at
the shift given in that row.
When finished with the partial structure column, you should be able to link the partial structures together to make an entire
structure in the assignment column.
An example of a spectrum and its accompanying data table is given below. Here is the spectrum:

Here is a data table:

Things to note:
This student has used two integration columns instead of just one.
The first column shows the integral measured from the spectrum. She probably used a ruler.
The second column, which she called int(n), contains a convenient ratio taken from the raw data. This ratio is easier to use
in her assignments.
Also note that the peak at 9.7 ppm does not have a very good integral. There is either a "phasing" or a "level & tilt"
problem here that can be corrected using the NMR software, but this is sometimes difficult to do. If she had taken an
automatic printout of this integral measurement, she would have gotten a strange number; in this case, it would be about -5,
because the end of the integral line is lower than the start. It clearly isn't a negative number of hydrogens, though. She has
instead measured the vertical rise in the integral and recorded that; it isn't perfect, but is a fair estimate in this case.

Contributors and Attributions


Chris P Schaller, Ph.D., (College of Saint Benedict / Saint John's University)

Chris Schaller 6/5/2021 2 https://chem.libretexts.org/@go/page/4190


High Resolution Proton NMR Spectra
This page describes how you interpret simple high resolution nuclear magnetic resonance (NMR) spectra. It assumes that you
have already read the background page on NMR so that you understand what an NMR spectrum looks like and the use of the
term "chemical shift". It also assumes that you know how to interpret simple low resolution spectra.

The difference between high and low resolution spectra


What a low resolution NMR spectrum tells you
The number of peaks tells you the number of different environments the hydrogen atoms are in.
The ratio of the areas under the peaks tells you the ratio of the numbers of hydrogen atoms in each of these environments.
The chemical shifts give you important information about the sort of environment the hydrogen atoms are in.
High resolution NMR spectra
In a high resolution spectrum, you find that many of what looked like single peaks in the low resolution spectrum are split into
clusters of peaks.

1 peak a singlet

2 peaks in the cluster a doublet

3 peaks in the cluster a triplet

4 peaks in the cluster a quartet

You can get exactly the same information from a high resolution spectrum as from a low resolution one - you simply treat each
cluster of peaks as if it were a single one in a low resolution spectrum. But in addition, the amount of splitting of the peaks
gives you important extra information.

Interpreting a high resolution spectrum


The n+1 rule
The amount of splitting tells you about the number of hydrogens attached to the carbon atom or atoms next door to the one you
are currently interested in. The number of sub-peaks in a cluster is one more than the number of hydrogens attached
to the next door carbon(s). So - on the assumption that there is only one carbon atom with hydrogens on next door
to the carbon we're interested in.

singlet next door to carbon with no hydrogens attached

doublet next door to a CH group

triplet next door to a CH2 group

quartet next door to a CH3 group

Using the n+1 rule


What information can you get from this NMR spectrum?

Assume that you know that the compound above has the molecular formula C4H8O2.

Jim Clark 6/11/2021 1 https://chem.libretexts.org/@go/page/3740


Treating this as a low resolution spectrum to start with, there are three clusters of peaks and so three different environments for
the hydrogens. The hydrogens in those three environments are in the ratio 2:3:3. Since there are 8 hydrogens altogether, this
represents a CH2 group and two CH3 groups. What about the splitting?
The CH2 group at about 4.1 ppm is a quartet. That tells you that it is next door to a carbon with three hydrogens
attached - a CH3 group.
The CH3 group at about 1.3 ppm is a triplet. That must be next door to a CH2 group. This combination of these
two clusters of peaks - one a quartet and the other a triplet - is typical of an ethyl group, CH3CH2. It is very
common.
Finally, the CH3 group at about 2.0 ppm is a singlet. That means that the carbon next door doesn't have any hydrogens
attached.
So what is this compound? You would also use chemical shift data to help to identify the environment each group was in, and
eventually you would come up with:

Alcohols
Where is the -O-H peak? This is very confusing! Different sources quote totally different chemical shifts for the
hydrogen atom in the -OH group in alcohols - often inconsistently. For example:

The Nuffield Data Book quotes 2.0 - 4.0, but the Nuffield text book shows a peak at about 5.4.
The OCR Data Sheet for use in their exams quotes 3.5 - 5.5.
A reliable degree level organic chemistry text book quotes1.0 - 5.0, but then shows an NMR spectrum for ethanol with a
peak at about 6.1.
The SDBS database (used throughout this site) gives the -OH peak in ethanol at about 2.6.
The problem seems to be that the position of the -OH peak varies dramatically depending on the conditions - for example,
what solvent is used, the concentration, and the purity of the alcohol - especially on whether or not it is totally dry.

A clever way of picking out the -OH peak


If you measure an NMR spectrum for an alcohol like ethanol, and then add a few drops of deuterium oxide, D2O, to the
solution, allow it to settle and then re-measure the spectrum, the -OH peak disappears! By comparing the two spectra, you can
tell immediately which peak was due to the -OH group.
The reason for the loss of the peak lies in the interaction between the deuterium oxide and the alcohol. All alcohols, such as
ethanol, are very, very slightly acidic. The hydrogen on the -OH group transfers to one of the lone pairs on the oxygen of the
water molecule. The fact that here we've got "heavy water" makes no difference to that.

Jim Clark 6/11/2021 2 https://chem.libretexts.org/@go/page/3740


The negative ion formed is most likely to bump into a simple deuterium oxide molecule to regenerate the alcohol -
except that now the -OH group has turned into an -OD group.

Deuterium atoms don't produce peaks in the same region of an NMR spectrum as ordinary hydrogen atoms, and so the peak
disappears.
You might wonder what happens to the positive ion in the first equation and the OD- in the second one. These get lost into the
normal equilibrium which exists wherever you have water molecules - heavy or otherwise.

The lack of splitting with -OH groups


Unless the alcohol is absolutely free of any water, the hydrogen on the -OH group and any hydrogens on the next door carbon
don't interact to produce any splitting. The -OH peak is a singlet and you don't have to worry about its effect on the next door
hydrogens.

The left-hand cluster of peaks is due to the CH2 group. It is a quartet because of the 3 hydrogens on the next door CH3 group.
You can ignore the effect of the -OH hydrogen. Similarly, the -OH peak in the middle of the spectrum is a singlet. It
hasn't turned into a triplet because of the influence of the CH2 group.

Equivalent hydrogen atoms


Hydrogen atoms attached to the same carbon atom are said to be equivalent. Equivalent hydrogen atoms have no effect on
each other - so that one hydrogen atom in a CH2 group doesn't cause any splitting in the spectrum of the other one.
But hydrogen atoms on neighboring carbon atoms can also be equivalent if they are in exactly the same environment. For
example:

These four hydrogens are all exactly equivalent. You would get a single peak with no splitting at all. You only have to
change the molecule very slightly for this no longer to be true.

Because the molecule now contains different atoms at each end, the hydrogens are no longer all in the same environment. This
compound would give two separate peaks on a low resolution NMR spectrum. The high resolution spectrum would show that
both peaks subdivided into triplets - because each is next door to a differently placed CH2 group.

Contributors and Attributions


Jim Clark (Chemguide.co.uk)

Jim Clark 6/11/2021 3 https://chem.libretexts.org/@go/page/3740


Integration in Proton NMR
There is additional information obtained from 1H NMR spectroscopy that is not typically available from 13C NMR
spectroscopy. Chemical shift can show how many different types of hydrogens are found in a molecule; integration reveals the
number of hydrogens of each type. An integrator trace (or integration trace) can be used to find the ratio of the
numbers of hydrogen atoms in different environments in an organic compound.
An integrator trace is a computer generated line which is superimposed on a proton NMR spectra. In the diagram, the
integrator trace is shown in red.

An integrator trace measures the relative areas under the various peaks in the spectrum. When the integrator trace
crosses a peak or group of peaks, it gains height. The height gained is proportional to the area under the peak or
group of peaks. You measure the height gained at each peak or group of peaks by measuring the distances shown
in green in the diagram above - and then find their ratio.
For example, if the heights were 0.7 cm, 1.4 cm and 2.1 cm, the ratio of the peak areas would be 1:2:3. That in turn
shows that the ratio of the hydrogen atoms in the three different environments is 1:2:3.

Figure NMR16.1H NMR spectrum of ethanol with solid integral line. Source: Spectrum taken in CDCl3 on a Varian Gemini
2000 Spectrometer with 300 MHz Oxford magnet.
Looking at the spectrum of ethanol, you can see that there are three different kinds of hydrogens in the molecule. You can also
see by integration that there are three hydrogens of one type, two of the second type, and one of the third type -- corresponding
to the CH3 or methyl group, the CH2 or methylene group and the OH or hydroxyl group. That information helps narrow down
the number of possible structures of the sample, and so it makes structure elucidation of an unknown sample much easier.
integration reveals the ratio of one type of hydrogen to another within a molecule.
Integral data can be given in different forms. You should be aware of all of them. In raw form, an integral is a horizontal line
running across the spectrum from left to right. Where the line crosses the frequency of a peak, the area of the peak is
measured. This measurement is shown as a jump or step upward in the integral line; the vertical distance that the line rises is
proportional to the area of the peak. The area is related to the amount of radio waves absorbed at that frequency, and the
amount of radio waves absorbed is proportional to the number of hydrogen atoms absorbing the radio waves.
Sometimes, the integral line is cut into separate integrals for each peak so that they can be compared to each other more easily.

Figure NMR17.1H NMR spectrum of ethanol with broken integral line. Source: Spectrum taken in CDCl3 on a Varian Gemini
2000 Spectrometer with 300 MHz Oxford magnet.
Often, instead of displaying raw data, the integrals are measured and their heights are displayed on the spectrum.

Figure NMR18.1H NMR spectrum of ethanol with numerical integrals.


Source: Spectrum taken in CDCl3 on a Varian Gemini 2000 Spectrometer with 300 MHz Oxford magnet.

Chris Schaller 6/11/2021 1 https://chem.libretexts.org/@go/page/4200


Sometimes the heights are "normalized". They are reduced to a lowest common factor so that their ratios are easier to
compare. These numbers could correspond to numbers of hydrogens, or simply to their lowest common factors. Two peaks in
a ratio of 1H:2H could correspond to one and two hydrogens, or they could correspond to two and four hydrogens, etc.

Figure NMR19.1H NMR spectrum of ethanol with normalized integral numbers.


Source: Spectrum taken in CDCl3 on a Varian Gemini 2000 Spectrometer with 300 MHz Oxford magnet.

Problem NMR.6.
Sketch a predicted NMR spectrum for each of the following compounds, with an integral line over each peak.

Problem NMR.7.
Measure the integrals in the following compounds. Given the integral ratios and chemical shifts, can you match each peak to a
set of protons?

Contributors and Attributions


Chris P Schaller, Ph.D., (College of Saint Benedict / Saint John's University)
Jim Clark (Chemguide.co.uk)

Chris Schaller 6/11/2021 2 https://chem.libretexts.org/@go/page/4200


Interpreting C-13 NMR Spectra
This page takes an introductory look at how you can get useful information from a C-13 NMR spectrum.

Typical chemical shifts in C-13 NMR spectra


In the table, the "R" groups won't necessarily be simple alkyl groups. In each case there will be a carbon atom
attached to the one shown in red, but there may well be other things substituted into the "R" group.

carbon environment chemical shift (ppm)

C=O (in ketones) 205 - 220


C=O (in aldehydes) 190 - 200
C=O (in acids and esters) 170 - 185
C in aromatic rings 125 - 150
C=C (in alkenes) 115 - 140
RCH2OH 50 - 65
RCH2Cl 40 - 45
RCH2NH2 37 - 45
R3CH 25 - 35
CH3CO- 20 - 30
R2CH2 16 - 25
RCH3 10 - 15

If a substituent is very close to the carbon in question, and very electronegative, that might affect the values given
in the table slightly. For example, ethanol has a peak at about 60 because of the CH2OH group. It also has a peak
due to the RCH3 group. The "R" group this time is CH2OH. The electron pulling effect of the oxygen atom
increases the chemical shift slightly from the one shown in the table to a value of about 18.

Example 1: Ethanol

Remember that each peak identifies a carbon atom in a different environment within the molecule. In this case there are two peaks
because there are two different environments for the carbons. The carbon in the CH3 group is attached to 3 hydrogens and a carbon.
The carbon in the CH2 group is attached to 2 hydrogens, a carbon and an oxygen. So which peak is which?
You might remember from the introductory page that the external magnetic field experienced by the carbon nuclei is affected by the
electronegativity of the atoms attached to them. The effect of this is that the chemical shift of the carbon increases if you attach an
atom like oxygen to it. That means that the peak at about 60 (the larger chemical shift) is due to the CH2 group because it has a
more electronegative atom attached.

Example 2: But-3-en-2-one

This is also known as 3-buten-2-one (amongst many other things!)

Jim Clark 5/28/2021 1 https://chem.libretexts.org/@go/page/3737


Here is the structure for the compound:
You can pick out all the peaks in this compound using the simplified table above.
The peak at just under 200 is due to a carbon-oxygen double bond. The two peaks at 137 and 129 are due to the carbons at
either end of the carbon-carbon double bond. And the peak at 26 is the methyl group which, of course, is joined to the rest of the
molecule by a carbon-carbon single bond.
If you want to use the more accurate table, you have to put a bit more thought into it - and, in particular, worry about the values which
don't always exactly match those in the table!
The carbon-oxygen double bond in the peak for the ketone group has a slightly lower value than the table suggests for a ketone.
There is an interaction between the carbon-oxygen and carbon-carbon double bonds in the molecule which affects the value
slightly. This isn't something which we need to look at in detail for the purposes of this topic.
You must be prepared to find small discrepancies of this sort in more complicated molecules - but don't worry about this for exam
purposes at this level. Your examiners should give you shift values which exactly match the compound you are given.
The two peaks for the carbons in the carbon-carbon double bond are exactly where they would be expected to be. Notice that they
aren't in exactly the same environment, and so don't have the same shift values. The one closer to the carbon-oxygen double bond
has the larger value.
And the methyl group on the end has exactly the sort of value you would expect for one attached to C=O. The table gives a range
of 20 - 30, and that's where it is.
One final important thing to notice. There are four carbons in the molecule and four peaks because they are all in different
environments. But they aren't all the same height. In C-13 NMR, you cannot draw any simple conclusions from the heights of the
various peaks.

Example 3: Isopropyl propanoate

1-methylethyl propanoate is also known as isopropyl propanoate or isopropyl propionate.

Here is the structure for 1-methylethyl propanoate:


Two simple peaks
There are two very simple peaks in the spectrum which could be identified easily from the second table above.
1. The peak at 174 is due to a carbon in a carbon-oxygen double bond. (Looking at the more detailed table, this peak is due to the
carbon in a carbon-oxygen double bond in an acid or ester.)
2. The peak at 67 is due to a different carbon singly bonded to an oxygen. Those two peaks are therefore due to:

Jim Clark 5/28/2021 2 https://chem.libretexts.org/@go/page/3737


If you look back at the more detailed table of chemical shifts, you will find that a carbon singly bonded to an oxygen has a range of
50 - 65. 67 is, of course, a little bit higher than that.
As before, you must expect these small differences. No table can account for all the fine differences in environment of a carbon in a
molecule. Different tables will quote slightly different ranges. At this level, you can just ignore that problem!
Before we go on to look at the other peaks, notice the heights of these two peaks we've been talking about. They are both due to a
single carbon atom in the molecule, and yet they have different heights. Again, you can't read any reliable information directly from
peak heights in these spectra.
The three right-hand peaks
From the simplified table, all you can say is that these are due to carbons attached to other carbon atoms by single bonds. But
because there are three peaks, the carbons must be in three different environments. The more detailed table is more helpful.
The easiest peak to sort out is the one at 28. If you look back at the table, that could well be a carbon attached to a carbon-oxygen
double bond. The table quotes the group as CH3CO-, but replacing one of the hydrogens by a simple CH3 group won't make much
difference to the shift value.
The right-hand peak is also fairly easy. This is the left-hand methyl group in the molecule. It is attached to an admittedly complicated
R group (the rest of the molecule). It is the bottom value given in the detailed table.
The tall peak at 22 must be due to the two methyl groups at the right-hand end of the molecule - because that's all that's left. These
combine to give a single peak because they are both in exactly the same environment.
If you are looking at the detailed table, you need to think very carefully which of the environments you should be looking at. Without
thinking, it is tempting to go for the R2CH2 with peaks in the 16 - 25 region. But you would be wrong!
The carbons we are interested in are the ones in the methyl group, not in the R groups. These carbons are again in the environment:
RCH3. The R is the rest of the molecule.
The table says that these should have peaks in the range 10 - 15, but our peak is a bit higher. This is because of the presence of the
nearby oxygen atom. Its electronegativity is pulling electrons away from the methyl groups - and, as we've seen above, this tends to
increase the chemical shift slightly.
Once again, don't worry about the discrepancies. In an exam, perhaps your examiners will just want you to have learnt the simple
table above - in which case, they can't expect you to work out which peak is which in a complicated spectrum of this sort. Or they will
give you tables of chemical shifts - in which case, they will give you values which match the peaks in the spectra.

Working out structures from C-13 NMR spectra


So far, we've just been trying to see the relationship between carbons in particular environments in a molecule and
the spectrum produced. We've had all the information necessary. Now let's make it a little more difficult - but we'll
work from much easier examples!
In each example, try to work it out for yourself before you read the explanation.

Example 1
How could you tell from just a quick look at a C-13 NMR spectrum (and without worrying about chemical shifts)
whether you had propanone or propanal (assuming those were the only options)?

Because these are isomers, each has the same number of carbon atoms, but there is a difference between the
environments of the carbons which will make a big impact on the spectra.
In propanone, the two carbons in the methyl groups are in exactly the same environment, and so will produce only
a single peak. That means that the propanone spectrum will have only 2 peaks - one for the methyl groups and
one for the carbon in the C=O group. However, in propanal, all the carbons are in completely different
environments, and the spectrum will have three peaks.

Example 2
Thare are four alcohols with the molecular formula C4H10O.

Jim Clark 5/28/2021 3 https://chem.libretexts.org/@go/page/3737


Which one produced the C-13 NMR spectrum below?

You can do this perfectly well without referring to chemical shift tables at all.
In the spectrum there are a total of three peaks - that means that there are only three different environments for the
carbons, despite there being four carbon atoms.
In A and B, there are four totally different environments. Both of these would produce four peaks.
In D, there are only two different environments - all the methyl groups are exactly equivalent. D would only produce
two peaks.
That leaves C. Two of the methyl groups are in exactly the same environment - attached to the rest of the molecule
in exactly the same way. They would only produce one peak. With the other two carbon atoms, that would make a
total of three. The alcohol is C.

Example 3
This follows on from Example 2, and also involves an isomer of C4H10O but which isn't an alcohol. Its C-13 NMR
spectrum is below. Work out what its structure is.

Because we don't know what sort of structure we are looking at, this time it would be a good idea to look at the shift
values. The approximations are perfectly good, and we will work from this table:
carbon environment chemical shift (ppm)

C-C 0 - 50
C-O 50 - 100
C=C 100 - 150
C=O 150 - 200

There is a peak for carbon(s) in a carbon-oxygen single bond and one for carbon(s) in a carbon-carbon single
bond. That would be consistent with C-C-O in the structure. It isn't an alcohol (you are told that in the question),
and so there must be another carbon on the right-hand side of the oxygen in the structure in the last paragraph.

Jim Clark 5/28/2021 4 https://chem.libretexts.org/@go/page/3737


The molecular formula is C4H10O, and there are only two peaks. The only solution to that is to have two identical
ethyl groups either side of the oxygen.
The compound is ethoxyethane (diethyl ether), CH3CH2OCH2CH3.

Example 4
Using the simplified table of chemical shifts above, work out the structure of the compound with the following C-13
NMR spectrum. Its molecular formula is C4H6O2.

Let's sort out what we've got.


There are four peaks and four carbons. No two carbons are in exactly the same environment.
The peak at just over 50 must be a carbon attached to an oxygen by a single bond.
The two peaks around 130 must be the two carbons at either end of a carbon-carbon double bond.
The peak at just less than 170 is the carbon in a carbon-oxygen double bond.
Putting this together is a matter of playing around with the structures until you have come up with something
reasonable. But you can't be sure that you have got the right structure using this simplified table.
In this particular case, the spectrum was for the compound:

If you refer back to the more accurate table of chemical shifts towards the top of the page, you will get some better
confirmation of this. The relatively low value of the carbon-oxygen double bond peak suggests an ester or acid
rather than an aldehyde or ketone.
It can't be an acid because there has to be a carbon attached to an oxygen by a single bond somewhere - apart
from the one in the -COOH group. We've already accounted for that carbon atom from the peak at about 170. If it
was an acid, you would already have used up both oxygens in the structure in the -COOH group.
Without this information, though, you could probably come up with reasonable alternative structures. If you were
working from the simplified table in an exam, your examiners would have to allow any valid alternatives.

Contributors and Attributions


Template:ContribCalark

Jim Clark 5/28/2021 5 https://chem.libretexts.org/@go/page/3737


Introduction to Proton NMR
This page describes what a proton NMR spectrum is and how it tells you useful things about the hydrogen atoms in
organic molecules.

The background to NMR spectroscopy


Nuclear magnetic resonance is concerned with the magnetic properties of certain nuclei. On this page we are
focusing on the magnetic behaviour of hydrogen nuclei - hence the term proton NMR or 1H-NMR.1H NMR
spectroscopy is used more often than 13C NMR, partly because proton spectra are much easier to obtain than carbon spectra.
The 13C isotope is only present in about 1% of carbon atoms, and that makes it difficult to detect. The 1H isotope is almost
99% abundant, which helps make it easier to observe. Another advantage is that 1H NMR spectroscopy gives more
information than 13C NMR, as you will find out later.
Note that in this discussion, the word "proton" is used for "hydrogen atom", because it is the proton in the nucleus of the 1H
isotope that is observed in these experiments. Although 2H (deuterium) and 3H (tritium) are also NMR-active, they absorb at
frequencies that are different from the ones used in 1H NMR. The 1H isotope is also much more common than the other two,
so 1H NMR spectroscopy is more conveniently done than 2H NMR spectroscopy.

Hydrogen atoms as little magnets


If you have a compass needle, it normally lines up with the Earth's magnetic field with the north-seeking end
pointing north. Provided it isn't sealed in some sort of container, you could twist the needle around with your fingers
so that it pointed south - lining it up opposed to the Earth's magnetic field. It is very unstable opposed to the Earth's
field, and as soon as you let it go again, it will flip back to its more stable state.

Hydrogen nuclei also behave as little magnets and a hydrogen nucleus can also be aligned with an external
magnetic field or opposed to it. Again, the alignment where it is opposed to the field is less stable (at a higher
energy). It is possible to make it flip from the more stable alignment to the less stable one by supplying exactly the
right amount of energy.

The energy needed to make this flip depends on the strength of the external magnetic field used, but is usually in
the range of energies found in radio waves - at frequencies of about 60 - 100 MHz. (BBC Radio 4 is found between
92 - 95 MHz!)
It's possible to detect this interaction between the radio waves of just the right frequency and the proton as it flips
from one orientation to the other as a peak on a graph. This flipping of the proton from one magnetic alignment to
the other by the radio waves is known as the resonance condition.

The importance of the hydrogen atom's environment

Jim Clark 6/20/2021 1 https://chem.libretexts.org/@go/page/3738


What we've said so far would apply to an isolated proton, but real protons have other things around them -
especially electrons. The effect of the electrons is to cut down the size of the external magnetic field felt by the
hydrogen nucleus.

Suppose you were using a radio frequency of 90 MHz, and you adjusted the size of the magnetic field so that an
isolated proton was in the resonance condition.
If you replaced the isolated proton with one that was attached to something, it wouldn't be feeling the full effect of
the external field any more and so would stop resonating (flipping from one magnetic alignment to the other). The
resonance condition depends on having exactly the right combination of external magnetic field and radio
frequency.
How would you bring it back into the resonance condition again? You would have to increase the external magnetic
field slightly to compensate for the effect of the electrons. Now suppose that you attached the hydrogen to
something more electronegative. The electrons in the bond would be further away from the hydrogen nucleus, and
so would have less effect on the magnetic field around the hydrogen.

The external magnetic field needed to bring the hydrogen into resonance will be smaller if it is attached to a more
electronegative element, because the hydrogen nucleus feels more of the field. Even small differences in the
electronegativities of the attached atom or groups of atoms will make a difference to the magnetic field needed to
achieve resonance.

Summary
For a given radio frequency (say, 90 MHz) each hydrogen atom will need a slightly different magnetic field applied
to it to bring it into the resonance condition depending on what exactly it is attached to - in other words the
magnetic field needed is a useful guide to the hydrogen atom's environment in the molecule.

Features of an NMR spectrum


A simple NMR spectrum looks like this:

The peaks
There are two peaks because there are two different environments for the hydrogens - in the CH3 group and
attached to the oxygen in the COOH group. They are in different places in the spectrum because they need slightly
different external magnetic fields to bring them in to resonance at a particular radio frequency.

Jim Clark 6/20/2021 2 https://chem.libretexts.org/@go/page/3738


The sizes of the two peaks gives important information about the numbers of hydrogen atoms in each environment.
It isn't the height of the peaks that matters, but the ratio of the areas under the peaks. If you could measure the
areas under the peaks in the diagram above, you would find that they were in the ratio of 3 (for the larger peak) to
1 (for the smaller one).
That shows a ratio of 3:1 in the number of hydrogen atoms in the two environments - which is exactly what you
would expect for CH3COOH.
The need for a standard for comparison - TMS
Before we can explain what the horizontal scale means, we need to explain the fact that it has a zero point - at the
right-hand end of the scale. The zero is where you would find a peak due to the hydrogen atoms in
tetramethylsilane - usually called TMS. Everything else is compared with this.

You will find that some NMR spectra show the peak due to TMS (at zero), and others leave it out. Essentially, if you
have to analyse a spectrum which has a peak at zero, you can ignore it because that's the TMS peak.
TMS is chosen as the standard for several reasons. The most important are:
It has 12 hydrogen atoms all of which are in exactly the same environment. They are joined to exactly the same
things in exactly the same way. That produces a single peak, but it's also a strong peak (because there are lots
of hydrogen atoms).
The electrons in the C-H bonds are closer to the hydrogens in this compound than in almost any other one.
That means that these hydrogen nuclei are the most shielded from the external magnetic field, and so you
would have to increase the magnetic field by the greatest amount to bring the hydrogens back into resonance.
The net effect of this is that TMS produces a peak on the spectrum at the extreme right-hand side. Almost
everything else produces peaks to the left of it.

The chemical shift


The horizontal scale is shown as (ppm). is called the chemical shift and is measured in parts per million - ppm.
A peak at a chemical shift of, say, 2.0 means that the hydrogen atoms which caused that peak need a magnetic
field two millionths less than the field needed by TMS to produce resonance. A peak at a chemical shift of 2.0 is
said to be downfield of TMS. The further to the left a peak is, the more downfield it is.

Solvents for NMR spectroscopy


NMR spectra are usually measured using solutions of the substance being investigated. It is important that the
solvent itself does not contain any simple hydrogen atoms, because they would produce confusing peaks in the
spectrum. There are two ways of avoiding this. You can use a solvent such as tetrachloromethane, CCl4, which
does not contain any hydrogen, or you can use a solvent in which any ordinary hydrogen atoms are replaced by its
isotope, deuterium - for example, CDCl3 instead of CHCl3. All the NMR spectra used on this site involve CDCl3 as
the solvent.
Deuterium atoms have sufficiently different magnetic properties from ordinary hydrogen that they do not produce
peaks in the area of the spectrum that we are looking at.

Contributors and Attributions


Jim Clark (Chemguide.co.uk)
Chris P Schaller, Ph.D., (College of Saint Benedict / Saint John's University)

Jim Clark 6/20/2021 3 https://chem.libretexts.org/@go/page/3738


Low Resolution Proton NMR Spectra
This page describes how you interpret simple low resolution nuclear magnetic resonance (NMR) spectra. It
assumes that you have already read the background page on NMR so that you understand what an NMR
spectrum looks like and the use of the term "chemical shift".

Difference between high and low resolution spectra

A low resolution spectrum looks much simpler because it cannot distinguish between the individual peaks in the
various groups of peaks.

The numbers against the peaks represent the relative areas under each peak. That information is extremely
important in interpreting the spectra.

Interpreting a low resolution spectrum


Using the total number of peaks
Each peak represents a different environment for hydrogen atoms in the molecule. In the methyl propanoate
spectrum above, there are three peaks because there are three different environments for the hydrogens.
Remember that methyl propanoate is CH3CH2COOCH3. The hydrogens in the CH2 group are obviously in a
different environment from those in the CH3 groups. The two CH3 groups aren't in the same environment either.
One is attached to a CH2 group, the other to an oxygen.

Using the areas under the peaks


The ratio of the areas under the peaks tell you the ratio of the numbers of hydrogens in the various environments.
In the methyl propanoate case, the areas were in the ratio of 3:2:3, which is exactly what you want for the two
differently placed CH3 groups and the CH2 group.
You will probably be told the relative areas under the peaks - especially if you are only looking at low resolution
spectra, but it is just possible that you might have to work them out. NMR spectrometers have a device which
draws another line on the spectrum called an integrator trace (or integration trace). You can measure the relative
areas from this trace.

Using chemical shifts


The position of the peaks tells you useful things about what groups the various hydrogen atoms are in. The
important shifts for the groups present in methyl propanoate are:

Jim Clark 7/7/2021 1 https://chem.libretexts.org/@go/page/3739


Showing these groups on the low resolution spectrum gives:

Questions
1. An organic compound was known to be one of the following. Use its low resolution NMR spectrum to decide
which it is.

Notice that there are three peaks showing three different environments for the hydrogens. That eliminates methyl
ethanoate as a possibility because that would only give two peaks - due to the two differently situated CH3 group
hydrogens.
Does the ratio of the areas under the peaks help? Not in this case - both the other compounds would have three
peaks in the ratio of 1:2:3. Now you need to look at the chemical shifts:

Checking the positions of the various hydrogens in the two possible compounds against the chemical shift table
gives you this pattern of shifts:

Comparing these with the actual spectrum means that the substance was propanoic acid, CH3CH2COOH.

2. How would you use low resolution NMR to distinguish between the isomers propanone and propanal?

Jim Clark 7/7/2021 2 https://chem.libretexts.org/@go/page/3739


The propanone would only give one peak in its NMR spectrum because both CH3 groups are in an identical
environment - both are attached to -COCH3. The propanal would give three peaks with the areas underneath in the
ratio 3:2:1. You could refer to the chemical shift table above to decide where the peaks are likely to be found, but it
isn't really necessary.
3. How many peaks would there be in the low resolution NMR spectrum of the following compound, and what
would be the ratio of the areas under the peaks?

All the CH3 groups are exactly equivalent so would only produce 1 peak. There would also be peaks for the
hydrogens in the CH2 group and the COOH group. There would be three peaks in total with areas in the ratio 9:2:1.

Contributors and Attributions


Jim Clark (Chemguide.co.uk)

Jim Clark 7/7/2021 3 https://chem.libretexts.org/@go/page/3739


More About Electronics
Methoxybenzene or anisole has six carbons, but only four peaks in the spectrum because of symmetry. These peaks are all
above 100 ppm, but some peaks are as far downfield as 160 ppm.

Figure NMR9.13C NMR spectrum of methoxybenzene (anisole).


Source: SDBSWeb : http://riodb01.ibase.aist.go.jp/sdbs/ (National Institute of Advanced Industrial Science and Technology of
Japan, 15 August 2008)
usually, a trigonal planar carbon shows up in the downfield half of the spectrum.
the region from about 100 to 200 ppm can be thought of as the sp2 window.
an electronegative atom moves a peak further downfield within the sp2 window.
Benzaldehyde has peaks between 130 and 140 ppm, as well as one near 190 ppm. Just as in the sp3 region of the spectrum,
when a carbon is attached to an electronegative element, it moves further downfield, and since the carbonyl (or C=O) carbon
in the aldehyde has two bonds to oxygen, it shows up considerably downfield. The carbonyl carbon in some ketones can show
up as far as 210 ppm.

Chris Schaller 7/7/2021 1 https://chem.libretexts.org/@go/page/4197


Figure NMR10. 13C NMR spectrum of benzaldehyde.
Source: SDBSWeb : http://riodb01.ibase.aist.go.jp/sdbs/ (National Institute of Advanced Industrial Science and Technology of
Japan, 15 August 2008)
Carbonyl carbons may show up even further downfield than 200 ppm.
Carbonyl carbons have a great deal of positive charge and low electron density.

Contributors and Attributions


Chris P Schaller, Ph.D., (College of Saint Benedict / Saint John's University)

Chris Schaller 7/7/2021 2 https://chem.libretexts.org/@go/page/4197


Multiplicity in Proton NMR
Another type of additional data available from 1H NMR spectroscopy is called multiplicity or coupling. Coupling is useful
because it reveals how many hydrogens are on the next carbon in the structure. That information helps to put an entire
structure together piece by piece.
In ethanol, CH3CH2OH, the methyl group is attached to a methylene group. The 1H spectrum of ethanol shows this
relationship through the shape of the peaks. The peak near 3.5 ppm is the methylene group with an integral of 2H.
This peak is split into four smaller peaks, evenly spaced, with taller peaks in the middle and shorter on the outside.
This pattern is called a multiplet, and specifically a quartet. A quartet means that these hydrogens have three neighbouring
hydrogens on adjacent carbons.

Figure NMR20.1H NMR spectrum of ethanol. Source: Modified from SDBSWeb : http://riodb01.ibase.aist.go.jp/sdbs/
(National Institute of Advanced Industrial Science and Technology of Japan, 15 August 2008)
The integral of 2H means that this group is a methylene, so it has two hydrogens. The carbon bearing these two hydrogens can
have two other bonds.

There could be two hydrogens on one neighbouring carbon and one on another. Otherwise, all three hydrogens could be on one
neighbouring carbon.

However, the shift of 3.5 ppm means that this carbon is attached to an oxygen. Mutliplicity usually only works with hydrogens
on neighbouring carbons. If there is an oxygen on one side of the methylene, all three neighbouring hydrogens must be on a
carbon on the other side.

Alternatively, look at the spectrum the other way around. The peak at 1 ppm is the methyl group with an integral of 3H.
The peak is split into three smaller ones, evenly spaced, with a taller one in the middle and shorter ones on the outside.
This pattern is called a triplet. A triplet means that these hydrogens have two neighbouring hydrogens on adjacent carbons.

Chris Schaller 5/23/2021 1 https://chem.libretexts.org/@go/page/4188


The neighbouring H could be on two different neighbouring carbons or both on the same one.

But this group is a methyl; the carbon already has three bonds, so it can have only one neighbouring carbon. It is next to a
methylene group.
The number of lines in a peak is always one more than the number of hydrogens on the neighboring carbon. The triplet for the
methyl peak means that there are two neighbors on the next carbon (3 - 1 = 2H); the quartet for the methylene peak indicates
that there are three hydrogens on the next carbon (4 - 1 = 3H). Table NMR 1 summarizes coupling patterns that arise when
protons have different numbers of neighbors.

# of lines ratio of lines term for peak # of neighbors

1 - singlet 0
2 1:1 doublet 1
3 1:2:1 triplet 2
4 1:3:3:1 quartet 3
5 1:4:6:4:1 quintet 4
6 1:5:10:10:5:1 sextet 5
7 1:6:15:20:15:6:1 septet 6
8 1:7:21:35:35:21:7:1 octet 7
9 1:8:28:56:70:56:28:8:1 nonet 8

The third peak in the ethanol spectrum is usually a "broad singlet." This is the peak due to the OH. You would expect it to be a
triplet because it is next to a methylene. Under very specific circumstances, it does appear that way. However, coupling is
almost always lost on hydrogens bound to heteroatoms (OH and NH). The lack of communication between an OH or NH and
its neighbours is related to rapid proton transfer, in which that proton can trade places with another OH or NH in solution. This
exchange happens quite easily if there are even tiny traces of water in the sample.
In summary, multiplicity or coupling is what we call the appearance of a group of symmetric peaks representing one hydrogen
in NMR spectroscopy.
A proton can absorb at different frequencies because of the influence of neighbouring hydrogens.
Protons on one carbon atom are affected by different protons on the next carbon atom, provided those two carbons are
directly attached to each other.
Stated another way, these neighboring hydrogens must be three bonds away (and so this phenomenon is sometimes called
"three-bond coupling").
When a proton is coupled, the number of neighbouring hydrogens is one less than the number of peaks in the multiplet.
There are limitations on coupling:
coupling doesn't occur between hydrogens of the same type ("equivalent hydrogens"). In the proton spectrum of ethane,
CH3-CH3, you would only observe one singlet.
coupling often doesn't occur across heteroatoms such as oxygen and nitrogen. The OH peak in ethanol may be a singlet
instead of a triplet, although there are two hydrogens on the neighboring carbon. The methylene peak in ethanol may be a
quartet instead of a quintet, even though there are actually four neighboring hydrogens: three on the attached methyl and
one on the attached hydroxyl

Contributors and Attributions


Chris P Schaller, Ph.D., (College of Saint Benedict / Saint John's University)

Chris Schaller 5/23/2021 2 https://chem.libretexts.org/@go/page/4188


NMR11. More About Multiplicity
The n + 1 rule (number of lines in a multiplet = number of neighboring H + 1) will work for the majority of problems you may
encounter. Occasionally, you may see more complicated coupling. The spectrum of methyl acrylate is a good example. There
are a couple of points to note in this spectrum, beginning with the number of peaks.

Figure NMR21.1H NMR spectrum of methyl acrylate.


Source: Simulated spectrum.
the vinyl portion of the spectrum (hydrogens adjacent to the C=C double bond) shows three peaks, not just two.
although there are two protons on one vinyl carbon, each H is different because of cis/trans relationships: one H is cis to
the carbonyl and the other is trans to it.
these hydrogens are symmetry inequivalent.
In addition, there is a problem with coupling in the vinyl region.
the proton adjacent to the carbonyl, at 6 ppm, has two neighbours and should give a triplet.
instead, the peak from this proton shows four lines, not three.
this peak is not a quartet; the middle two peaks are not three times as tall as the edge ones.
the lines within this multiplet are symmetrically spaced, but not evenly; the middle space is smaller than the spaces on the
edges.
This pattern is called a "doublet of doublets." The two symmetry-inequivalent neighbors on the other end of the double bond
each act as if the other one isn't there. They couple to the proton next to the carbonyl independently, each one splitting the peak
for this proton into a separate doublet.
There are a few cases in which this independent coupling will occur rather than the (n+1) type coupling we saw first.
Generally, independent coupling occurs when protons are not freely rotating. That can happen if one of the protons is attached
to a double-bonded carbon, because we can't rotate around a double bond. It may also happen with protons that are directly
attached to the carbons of a ring.
To see why this happens, you need to know more about coupling.
in coupling, magnetic information is shared between two protons. How fully this information is shared depends on the
angle between the hydrogens as you look down the connecting C-C bond.
for (n+1) coupling, the rotational angle about the C-C bond connecting the two protons (or 'dihedral angle' between the
coupled hydrogens) must be unrestricted. The bond must be able to rotate freely.
if the bond can rotate freely, any proton on the neighboring carbon can assume any dihedral angle with the proton being
observed. That means coupling information from one neighboring proton is indistinguishable from another, and so all the
neighboring protons affect the observed proton equally.

Chris Schaller 6/29/2021 1 https://chem.libretexts.org/@go/page/4189


Figure NMR22. A sawhorse projection (top) and Newman projection of ethanol (bottom), showing the angular relationship
between hydrogens on neighboring carbons. The angle between two groups on neighboring carbons is called the dihedral
angle. Because of free rotation about the C-C bond, all three hydrogens on the methyl have an equivalent spatial relationship to
the hydrogens on the CH2 group.
Sometimes coupling information is depicted as an arrow. This arrow stands for the coupling constant between two protons.
The coupling constant is related to the spin of a hydrogen atom. The spin (related to magnetic moment) can be aligned with the
external magnetic field (we will show it pointing up) or else against it; no other possibilities are allowed.

If there are two neighboring hydrogens, both spins could be aligned with the external field, both could be aligned against it, or
one could be aligned each way. That means there are three different magnetic combinations that will each have a different
effect on the observed proton: increased magnetic field, decreased magnetic field, and no net effect (canceling out).

These three combinations result in the observed proton absorbing at three different frequencies, because the frequency it
absorbs is sensitive to the magnetic field it experiences. Note that there are two ways to arrive at the middle possibility, with
one neighbour spin up and the other spin down. Statistically this possibility is twice as likely as either both spins up or both
spins down. It is thus twice as likely that the observed proton experiences that effect, and so the middle line in a triplet is twice
as high as the other two lines.

Chris Schaller 6/29/2021 2 https://chem.libretexts.org/@go/page/4189


Figure NMR23. Effects of neighboring protons on an observed peak. This case assumes all the neighboring protons have an
equivalent effect (they have the same coupling constant with the observed proton).
However, the size of that arrow, the coupling constant, is only the same for two neighboring hydrogens if they have the same
spatial relationship with the observed hydrogen. That isn't always true.
if the dihedral angle is limited, complex coupling occurs.
complex coupling occurs because one neighboring hydrogen can only adopt a limited range of dihedral angles. Another
neighboring proton can adopt a limited range of dihedral angles as well, but these two ranges do not overlap.
As a result, the two coupling constants are different. We can depict that situation using arrows of different lengths for the two
neighboring proton spins. Each spin can be either up or down, but now two opposing spins do not cancel out. The result is four
spin combinations of equal probability, not just three.

The doublet of doublets is four lines of about equal heights. The distance between the two pairs of lines on each edge

Chris Schaller 6/29/2021 3 https://chem.libretexts.org/@go/page/4189


Figure NMR23. Effects of neighboring protons on an observed peak. This case assumes neighboring protons have inequivalent
effects (they have differing coupling constants with the observed proton).
The dihedral angle is limited in only a few specific cases:
there is a double bond. Double bonds cannot rotate because that would require breaking the pi bond. In a pi bond, p orbitals
on the two carbons must remain coplanar in order to overlap and form a bond.
there is a ring. In a ring, there can't be complete rotation about a bond because the ring would twist into a pretzel.
in some cases, there can be such large groups of atoms on either end of a bond that it is difficult to rotate the bond without
having these groups crash into each other.

Contributors and Attributions


Chris P Schaller, Ph.D., (College of Saint Benedict / Saint John's University)

Chris Schaller 6/29/2021 4 https://chem.libretexts.org/@go/page/4189


NMR14. More Practice with NMR Spectroscopy
Contributors and Attributions
Chris P Schaller, Ph.D., (College of Saint Benedict / Saint John's University)

Chris Schaller 6/20/2021 1 https://chem.libretexts.org/@go/page/4192


NMR2. Carbon-13 NMR
Contributors and Attributions
Chris P Schaller, Ph.D., (College of Saint Benedict / Saint John's University)

Chris Schaller 6/29/2021 1 https://chem.libretexts.org/@go/page/4193


NMR3. Symmetry in NMR
Butane shows two different peaks in the 13C NMR spectrum, below. Note that:
the chemical shifts of these peaks are not very different from methane. The carbons in butane are in a similar environment to
the one in methane.
there are two distinct carbons in butane: the methyl, or CH3, carbon, and the methylene, or CH2, carbon.
the methyl carbon absorbs slightly upfield, or at lower shift, around 10 ppm.
the methylene carbon absorbs at slightly downfield, or at higher shift, around 20 ppm.
other factors being equal, methylene carbons show up at slightly higher shift than methyl carbons.

Figure NMR2. Simulated 13C NMR spectrum of butane (showing only the upfield portion of the spectrum).
In the 13C NMR spectrum of pentane (below), you can see three different peaks, even though pentane just contains methyl
carbons and methylene carbons like butane. As far as the NMR spectrometer is concerned, pentane contains three different
kinds of carbon, in three different environments. That result comes from symmetry.

Figure NMR3.13C NMR spectrum of pentane.


Source: SDBSWeb : http://riodb01.ibase.aist.go.jp/sdbs/ (National Institute of Advanced Industrial Science and Technology of
Japan, 15 August 2008)
Symmetry is an important factor in spectroscopy. Nature says:
atoms that are symmetry-inequivalent can absorb at different shifts.
atoms that are symmetry-equivalent must absorb at the same shift.
To learn about symmetry, take a model of pentane and do the following:
make sure the model is twisted into the most symmetric shape possible: a nice "W".

Chris Schaller 6/20/2021 1 https://chem.libretexts.org/@go/page/4194


choose one of the methyl carbons to focus on.
rotate the model 180 degrees so that you are looking at the same "W" but from the other side.
note that the methyl you were focusing on has simply switched places with the other methyl group. These two carbons are
symmetry-equivalent via two-fold rotation.

Wire Frame

Ball & Stick

Spacefilling

Animation NMR1. A three-dimensional model of pentane. Grab the model with the mouse and rotate it so that you are
convinced that the second and fourth carbons are symmetry-equivalent, but the third carbon is not.
By the same process, you can see that the second and fourth carbons along the chain are also symmetry-equivalent. However,
the middle carbon is not; it never switches places with the other carbons if you rotate the model. There are three different sets
of inequivalent carbons; these three groups are not the same as each other according to symmetry.

Contributors and Attributions


Chris P Schaller, Ph.D., (College of Saint Benedict / Saint John's University)

Chris Schaller 6/20/2021 2 https://chem.libretexts.org/@go/page/4194


NMR4. 13C NMR and Geometry
Contributors and Attributions
Chris P Schaller, Ph.D., (College of Saint Benedict / Saint John's University)

Chris Schaller 6/2/2021 1 https://chem.libretexts.org/@go/page/4195


NMR5. 13C NMR and Electronics
Contributors and Attributions
Chris P Schaller, Ph.D., (College of Saint Benedict / Saint John's University)

Chris Schaller 6/20/2021 1 https://chem.libretexts.org/@go/page/4196


NMR8. Chemical Shift in 1H NMR
The trends here are exactly the same as in carbon spectra. Wherever the carbon goes, it takes the proton with it. By analogy
with carbon spectra,
hydrogens on sp3 carbons usually show up in the upfield half of the spectrum, about 0 to 5 ppm.
hydrogens on sp2 carbons usually show up in the downfield half of the spectrum, about 5 to 10 ppm.
within these two halves of the spectrum, electronegative atoms attached to the same carbon as a proton will draw that
proton downfield.

Figure NMR11.1H NMR spectrum of hexane.


Source: Simulated spectrum.

Figure NMR12.1H NMR spectrum of 1-hexene.


Source: Simulated spectrum.

Figure NMR13.1H NMR spectrum of butanal.


Source: Simulated spectrum.
As before, there are also hydrogens on linear carbons, although they are much less common than tetrahedral or trigonal
carbons.
hydrogens on sp carbons show up between 2 and 6 ppm.
Remember, these are general rules that you should know. There will occasionally be exceptions; the proton in a carboxylic
acid may be seen at 12 ppm, and the proton in chloroform shows up at 7 ppm although it is attached to a tetrahedral carbon.
(World-record shifts occur for hydrogens attached to transition metals: "late" metals like ruthenium or rhodium can move
hydrogen peaks all the way up to -20 ppm, but "early" metals like tantalum can move them down as far as 25 ppm.)

Contributors and Attributions

Chris Schaller 7/4/2021 1 https://chem.libretexts.org/@go/page/4199


Chris P Schaller, Ph.D., (College of Saint Benedict / Saint John's University)

Chris Schaller 7/4/2021 2 https://chem.libretexts.org/@go/page/4199


NMR Appendix. Useful Charts for NMR identification
Table of 13C NMR Frequencies Common in Organic Compounds.

Note that effects are additive: two or more electron-withdrawing groups move the absorbance further to the left than just one
group.
Table of 1H NMR Frequencies Common in Organic Compounds.

This chart shows the frequancies of protons that are attached to carbons. In general, protons follow the trend seen in the carbon
to which they are attached. Note again the additive effects of multiple attached groups.
This table does not include OH (or NH) protons. Protons attached to heteroatoms are more difficult to pinpoint because their
locations in the spectrum are much less specific. Instead, they may be found across a very broad range.
Table of 1H NMR Frequencies of OH Common in Organic Compounds.

Chris Schaller 6/22/2021 1 https://chem.libretexts.org/@go/page/4201


Contributors and Attributions
Chris P Schaller, Ph.D., (College of Saint Benedict / Saint John's University)

Chris Schaller 6/22/2021 2 https://chem.libretexts.org/@go/page/4201


NMR - Theory
Nuclear magnetic resonance plays an important role in the fields chemistry, materials science, physics and engineering. It is
becoming a more and more useful method to probe the structure of molecules. The primary object of this module is to
understand the fundamental concepts of NMR. It is assumed that the reader already understands the quantum numbers
associated with electrons. For a more basic understanding of how NMR works, the reader is directed to the NMR introduction
page.

Introduction
Nuclear magnetic resonance, NMR, is a physical phenomenon of resonance transition between magnetic energy levels,
happening when atomic nuclei are immersed in an external magnetic field and applied an electromagnetic radiation with
specific frequency. By detecting the absorption signals, one can acquire NMR spectrum. According to the positions, intensities
and fine structure of resonance peaks, people can study the structures of molecules quantitatively. The size of molecules of
interest varies from small organic molecules, to biological molecules of middle size, and even to some macromolecules such
as nucleic acids and proteins. Apart from these commonly utilized applications in organic compound, NMR also play an
important role in analyzing inorganic molecules, which makes NMR spectroscopy a powerful technique.
But the a major question still remains- Why does NMR work? This module will begin by developing the concept of nuclear
spin then moving into a discussion about energy levels and the relative populations and the interactions of a nucleus with the
magnetic field.

Nuclear Spin Origins


The concept of spin is regularly addressed in subatomic particle physics. However, to most people spin seems like an abstract
concept. This is due to the fact there is no macroscopic equivalent of what spin is. However, for those people who have taken
an introduction to chemistry course have seen the concept of spin in electrons. Electrons are subatomic particles which have
spin intrinsic to them. The nucleus is not much different. Spin is just another form of angular momentum. The nucleus consists
of protons and neutrons and neutrons and protons are comprised of subatomic particles known as quarks and gluons. The
neutron has 2 quarks with a -e/3 charge and one quark with a +2e/3 charge resulting in a total charge of 0. The proton however,
has 2 quarks with +2e/3 charge and only one quark with a -e/3 charge giving it a net positive charge. Both protons and
neutrons are spin=1/2.

Figure 1: The atomic nucleus (black) of 2H. The proton (green) and neutron (red) are composed of quarks (purple and teal)
which have a charge and spin (arrow).
For any system consisting of n multiple parts, each with an angular momentum the total angular momentum can be described
by J where
J = | J1 + J2 +. . . +Jn |, | J1 + J2 +. . . +Jn | − 1, . . . | J1 − J2 −. . . −Jn | (1)

2
a = mx (2)

Here are some examples using the isotopes of hydrogen


1
H = 1 proton so J=1/2
2
H = 1 proton and 1 neutron so J = 1 or 0.

6/10/2021 1 https://chem.libretexts.org/@go/page/1819
For larger nuclei, it is not immediately evident what the spin should be as there are a multitude of possible values. For the
remainder of the discussion we will attribute the spin of the nucleus, I, to be an intrinsic value. There are some rules that the
nuclei do follow with respect to nuclear spin. They are summarized in the table below.
Table 1. General rules for determination of nuclear spin quantum numbers
Mass Number Number of Protons Number of Neutrons Spin (I) Example

Even Even 0 16
O (3)

Even
Odd Odd Integer (1,2,...) 2
H (4)

Even Odd Half-Integer (1/2, 3/2,...) 13


C (5)

Odd
Odd Even Half-Integer (1/2, 3/2,...) 15
N (6)

Nuclear Spin Angular Momentum and Quantum Numbers


As mentioned above, spin is a type of angular momentum. Nuclear spin angular momentum was first reported by Pauli in 1924
and will be described here. Analogous to the angular momentum commonly encountered in electron, the angular momentum is
a vector which can be described by a magnitude L and a direction, m. The magnitude is given by
−−−−−−−
L = ℏ√ I (I + 1) (7)

The projection of the vector on the z axis (arbitrarily chosen), takes on discretized values according to m, where
m = −I , −I + 1, −I + 2, . . . +I (8)

The angular momentum along the z-axis is now


Iz = mℏ (9)

Pictorially, this is represented in the figure below for three values of I .

Figure 2: The quantized angular momentum values for a I=1/2 (left) I=1(middle) and I=2 (right). The magnitude is denoted by
the arrow while the projection along the z-axis is denoted by the circle.
The quantum numbers of the nucleus are denoted below.
Interaction Symbol Quantum Numbers

Nuclear Spin Angular Momentum I 0<I<9/2 by 1/2

−−−−−−−
Spin Angular Momentum Magnitude L L = ℏ√ I (I + 1) (10)

Spin Angular Momentum Direction m m = −I , −I + 1, −I + 2, . . . +I (11)

Magnetic Moment of a Nucleus


We have now established that the nucleus has spin which can be denoted using specific quantum numbers. As with all charged
particles, if the nucleus is moved in a loop it will generate a magnetic field. The magnetic moment μ is related to the angular

6/10/2021 2 https://chem.libretexts.org/@go/page/1819
momentum of the nucleus by
μ = γI (12)

where γ is the gyromagnetic ratio, a proportionality constant unique to each nucleus. The table below shows some of the
gyromagnetic ratios for some commonly studies nuclei.

The net or bulk magnetization of the sample is given by M and is the sum of each individual magnetic vector, or

M = ∑μ (13)

since these magnetic moments are vectors and are randomly aligned, the bulk magnetization arising from the nucleus is zero.
There may be unpaired electrons which give rise to paramagnetic, anti ferromagnetic, or ferromagnetic properties. However, if
an external magnetic field is applied, the nuclei will align either with or against the field and result in a non-zero bulk
magnetization.

Nuclear Energy Levels in a Magnetic Field


We now understand why the nucleus has a magnetic moment associated with it. Now we are getting to the crux of
NMR, the use of an external magnetic field. Initially, the nucleus is in the nuclear ground state which is degenerate.
The degeneracy of the ground state is 2I+1. The application of a magnetic field splits the degenerate 2I+1 nuclear
energy levels. The energy of a particular level is
E = −μ ⋅ B0 (14)

where B is the external magnetic field. Along the z-direction, which we assume the magnetic field is applied,
0

E = −μB0 (15)

by substitution,

E = −mℏγB0 (16)

The magnitude of the splitting therefore depends on the size of the magnetic field. In most labs this magnetic field is
somewhere between 1 and 21T. Those spins which align with the magnetic field are lower in energy, while those that align
against the field are higher in energy.

6/10/2021 3 https://chem.libretexts.org/@go/page/1819
Figure 3: Splitting of the energy levels for a I=1/2 (black dashed lines), I= 3/2 (blue dashed lines), and I=5/2 (red dashed
lines). Note how the energy level is dependent on the applied magnetic field. The offset of the red and blue lines is for
illustrative purposes only.

Energy Level Spin Distribution


In the absence of a magnetic field the magnetic dipoles are oriented randomly and there is no net magnetization (vector sum of
µ is zero). Application of an external magnetic field, as was shown above, creates distinct energy levels based on the spin
angular momentum of the nucleus. Each energy level is populated by the spins which have the same angular momentum. To
illustrate this, consider a I=1/2 system. There are two energy levels, +1/2 and -1/2, which are populated by spins that have
aligned against or with the external magnetic field, respectively.

Figure 4: Spins configurations according to applied magnetic field


The energy separation between these states is relatively small and the energy from thermal collisions is sufficient to place
many nuclei into higher energy spin states. The number of nuclei in each spin state can be described by the Boltzmann
distribution. The Boltzmann equation expresses the relationship between temperature and the related energy as shown below.

Nupper −ΔE −h ν

= e kT
= e kT
(17)
Nlower

Where Nupper and Nlower represent the population of nuclei in upper and lower energy states, E is the energy difference
between the spin states, k is the Boltzmann constant (1.3805x10-23 J/Kelvin ) and T is the temperature in K. At room
temperature, the number of spins in the lower energy level, N lower, slightly outnumbers the number in the upper level, N upper.

Selection Rules
The selection rule in NMR is

6/10/2021 4 https://chem.libretexts.org/@go/page/1819
Δm = ±1 (18)

For a nucleus with I=1/2 there is only one allowed transition. For nuclei with I > 1/2 , there are multiple transitions which can
take place. Consider the case of I=3/2. The following transitions can take place
3 1
− ↔ − (19)
2 2

1 1
− ↔ (20)
2 2

1 3
↔ (21)
2 2

which is illustrated below.

The transition from


3 1
− ↔ − (22)
2 2

1 3
↔ (23)
2 2

are known as satellite transitions, while the


1 1
− ↔ (24)
2 2

transition is known as the central transition. The central transition is primarily observed in an NMR experiment. For more
information about satellite transitions please look at quarupole interactions.

The NMR Experiment


During the NMR experiment several things happen to the nucleus, the bulk magnetization is rotated from the z axis into the xy
plane and then allowed to relax back along the z-axis. A full theoretical explanation for a single atom was developed by Bloch
into a set of equations known as the Bloch equations. From the NMR experiment chosen a variety of information can be
gleaned by studying different interactions.

References
1. Claude H. Yoder, Charles D. Schaeffer,Introduction to multinuclear NMR : theory and application, Benjamin/Cummings
Pub. Co., Menlo Park, Calif., 1987.
2. Robin K. Harris, Nuclear magnetic resonance spectroscopy : a physicochemical view, Pitman, Marshfield, Mass., 1983.
3. Jeremy K.M. Sanders and Brian K. Hunter, Modern NMR spectroscopy : a guide for chemists, Oxford University Press,
New York, 1993.
4. David A.R. Williams ; editor, David J. Mowthorpe, Nuclear magnetic resonance spectroscopy, Published on behalf of
ACOL, London, by J. Wiley, New York, 1986.
5. Frank A. Bovey, Lynn Jelinski, Peter A. Mirau, Nuclear magnetic resonance spectroscopy, Academic Press, San Diego,
1988.

6/10/2021 5 https://chem.libretexts.org/@go/page/1819
6. R.J. Abraham, J. Fisher and P. Loftus, Introduction to NMR spectroscopy, Wiley, New York, 1988.
7. J.W. Akitt, NMR and chemistry : an introduction to modern NMR spectroscopy, Chapman & Hall, London; New York,
1992.

Contributors
Derrick Kaseman (UC Davis), Sureyya OZCAN, Siyi Du

6/10/2021 6 https://chem.libretexts.org/@go/page/1819
Bloch Equations
In 1946 Felix Bloch, co-discoverer of NMR, proposed a set of equations to described the time dependence of the net
magnetization during the course of the NMR experiment. These equations are known as the Bloch equations and give insights
into many processes in NMR. The Bloch equations follow first order kinetics and the derivations are first-order differentials.

Bloch Equations in the Lab Frame


It has been shown that nuclei in placed in a magnetic field precess with a characteristic Larmor frequency, ω.

ω0 = −γ(1 − σ)B0 (1)

which can be made time-dependent by considering a time-dependent magnetic field B(t).


In addition we have also seen that there is a bulk magnetization, based on Boltzmann statistics, which lies along the direction
of the applied magnetic field, B . This bulk magnetization can then be thought of preces\sing at the Larmor frequency.
0

Assuming that the B is along the Z-axis, we can describe the time dependence of the magnetization M , by
0

dM
= ω(t)xM (t) − [R][M (t) − Me q] (2)
dt

where R is a rotation matrix, and M eq is the magnetization at equilibrium along the z-axis. The rotation matrix accounts for
relaxation processes and is given by
1
⎛ 0 0 ⎞
T2
⎜ ⎟
⎜ ⎟
⎜ 1 ⎟
R =⎜ 0 0 ⎟ (3)
⎜ T2 ⎟
⎜ ⎟
⎜ 1 ⎟

⎝ 0 0 ⎠
T1

The [R][M (t) − M ] term then describes the decay of the magnetization in the x-y plane and the growth of the equilibrium
eq

magnetization, due to T2 and T1 effects, repectively.


Expansion of the equation yeilds the magnetization in each direction
∗ Mx (t)
dMx (t) ⎛ ⎞
⎛ ⎞ [ Mx (t)ωy (t) − My ωz (t)] −
dt ⎜ T2 ⎟
⎜ ⎟
⎜ ⎟ ⎜ ⎟
∗ ⎜ ⎟
⎜ dMy (t) ⎟ My (t)
= ⎜ ⎟ (4)
⎜ ⎟ ⎜ [ Mx (t)ωz (t) − Mz ωx (t)] − ⎟
⎜ dt ⎟ ⎜ T2 ⎟
⎜ ⎟ ⎜ ⎟
⎜ ∗ ⎟
dMz (t) ⎜ [ Mz (t) − Meq ] ⎟
⎝ ⎠ ⎝ [ My (t)ωx (t) − Mx ωy (t)] − ⎠
dt T1

Assuming that we have a static magnetic field these equations simplify to



dMx (t) Mx (t)
⎛ ⎞ ⎛ ⎞
−ω0 My (t) −
dt ⎜ T2 ⎟
⎜ ⎟
⎜ ⎟ ⎜ ⎟

⎜ dMy (t) ⎟ ⎜ My (t) ⎟
⎜ ⎟
⎜ ⎟ =⎜ ω0 Mx (t) − ⎟
(5)
⎜ dt ⎟ ⎜ T2 ⎟
⎜ ⎟ ⎜ ⎟
⎜ ∗ ⎟ ⎜ ⎟
dMz (t) [ Mz (t) − Meq ]
⎝ ⎠ ⎝ − ⎠
dt T1

As the magnetization is preces\sing about the Z axis as it recovers from the x-y plane it is realized that the time dependance
will follow an oscillatory pattern which can be described u\sing trigonometric functions. Therefore at a given time the
magnetization in any direction will be

6/5/2021 1 https://chem.libretexts.org/@go/page/1820
−t
⎛ ⎞
T2
⎜ [ Mx (0) cos ω0 t − My (0) sin ω0 t] e ⎟
Mx (t) ⎜ ⎟
⎛ ⎞ ⎜ −t ⎟
⎜ ⎟
⎜ My (t) ⎟ = ⎜ ⎟ (6)
⎜ [ M (0) cos ω t − M (0) sin ω t] e T2 ⎟
y 0 x 0
⎝ ⎠ ⎜ ⎟
Mz (t) ⎜ ⎟
⎜ −t −t ⎟

⎝ ⎠
Mz (0)e T1 + Meq (1 − e T1 )

Where M x,y,z (0) is the magnetization at t=0.

What Do The Equations Describe?


From this derivation we are able to describe the motion of the bulk magnetization of the nuclei after a pulse as a function of
time accounting for relaxation.

Bloch Equations in the Rotating Frame


π
One can imagine that application of a pulse in which the B- field now becomes time dependent would be challenging to
2
calculate u\sing the Bloch equations in the lab frame. Therefore we now shift our discussion to the rotating frame. It should be
stated here that the insertion of the to denote the magnetization projection on the rotating frame.

The frequency of rotation ω , is equivalent to spins precession frequency, which will be the Larmor frequency. From this we
rot

can describe the change of the magnetization as



d M dM
= − ωrot xM (7)
dt dt

dM
We know that is given by the bloch equations we derived for the lab frame that is:
dt

dM
= ω(t)xM (t) − [R][M (t) − Me q] (8)
dt

Substituting this into our equation, we obtain



d M
= ω(t)XM − ωrot xM − [R][M − Meq ] (9)
dt

which simplifies (u\sing cross product relations) to



d M
= ωef f (t)xM − [R][M − Meq ] (10)
dt

where
ωef f (t) = ω − ωrot (11)

Noting that the relation between the lab frame frame and the rotating frame is the common z-axis, we can define the Bloch
equations in the rotating frame

∗ Mx (t)
dMx (t) ⎛ ∗ ∗ ⎞
⎛ ⎞ γ[ My (t)Bz (t) − Mz (t)By (t)] −
dt ⎜ T2 ⎟
⎜ ⎟
⎜ ⎟ ⎜ ∗

∗ ⎜ ⎟
⎜ dMy (t) ⎟ My (t)
= ⎜ ∗ ∗ ⎟ (12)
⎜ ⎟ ⎜ γ[ Mz (t)Bx (t) − Mx (t)Bz (t)] − ⎟
⎜ dt ⎟ ⎜ T2 ⎟
⎜ ⎟ ⎜ ⎟
⎜ ∗ ⎟ ∗ ∗
dMz (t) ⎜ [ Mz (t) − Meq ] ⎟
∗ ∗
⎝ ⎠ ⎝ γ[ Mx (t)By (t) − My (t)Bx (t)] − ⎠
dt T2

Which can be simplified to give

6/5/2021 2 https://chem.libretexts.org/@go/page/1820

∗ Mx (t)
dMx (t) ⎛ ∗ ∗ ∗ ⎞
⎛ ⎞ −Ω My (t) − γ Mz (t)By (t)] −
dt ⎜ T2 ⎟
⎜ ⎟
⎜ ⎟ ⎜ ⎟
∗ ∗
⎜ dMy (t) ⎟ ⎜ My (t) ⎟
⎜ ∗ ∗ ∗ ⎟
⎜ ⎟ =⎜ −γ[ Mx (t) + γ Mz (t)Bx (t)] − ⎟
(13)
⎜ dt ⎟ ⎜ T2 ⎟
⎜ ⎟
⎜ ⎟
⎜ ∗ ⎟ ⎜ ∗ ∗

dMz (t) [ Mz (t) − Meq ]
∗ ∗ ∗ ∗
⎝ ⎠ ⎝ γ[ Mx (t)By (t) − My (t)Bx (t)] − ⎠
dt T2

where
Ω = ω0 − ωrot (14)

Effects of RF pulses
During the NMR experiment, we can choose which direction we can apply our magnetic field during the pulse. For simplicity
we assume that we apply the pulse along the +x axis. The nutation frequency is then defined as
ω1 =∣ γ B1 (t)∣ (15)

where B is the magnetic field applied along rhe x-axis. Application of this to the Bloch equations in the rotating frame we
1

obtain

−Mx
⎛ ∗
ω1 Mz sin ϕ − ⎞


T2
Mx (t) ⎜ ⎟
⎛ ⎞ ⎜ ∗ ⎟
⎜ −My ⎟
∗ ∗
⎜ My (t) ⎟ = ⎜ −ω1 Mz cos ϕ − ⎟ (16)
⎜ T2 ⎟
⎝ ∗ ⎠ ⎜ ⎟
Mz (t) ∗ ∗
⎜ (Mz − Me q) ⎟
∗ ∗
⎝ −ω1 Mz sin ϕ + ω1 Mz cos ϕ − ⎠
T1

where ϕ is the phase of the pulse (0=x axis and 270=-y axis). The RF pulse is much shorter than T1, T2, or ω . Therefore we 0

can neglect any terms containing these values.THe equations then simplify to
∗ ∗
Mx (t) ω1 Mz sin ϕ
⎛ ⎞ ⎛ ⎞
∗ ∗
⎜ My (t) ⎟ = ⎜ −ω1 Mz cos ϕ ⎟ (17)

⎝ ∗ ⎠ ⎝ ∗ ∗ ⎠
Mz (t) −ω1 Mz sin ϕ + ω1 Mz cos ϕ

Thereby application of a pulse of phase zero will result in the following result for the magnetization as a function of time

dMx (t)
⎛ ⎞

⎜ dt ⎟ 0
⎜ ∗ ⎟ ⎛ ⎞
⎜ dMy (t) ⎟ ∗
⎜ ⎟ = ⎜ γ B1 (t) Mz ⎟ (18)
⎜ ⎟
⎜ dt ⎟ ⎝ ∗ ⎠
⎜ ∗ ⎟ −γ B 1 (t)M z
dMz (t)
⎝ ⎠
dt

Or more generally as
∗ ∗
Mx (t) Mx
⎛ ⎞ ⎛ ⎞
∗ ∗ ∗
⎜ My (t) ⎟ = ⎜ My cos ω1 t − Mz sin ω1 t ⎟ (19)

⎝ ∗ ⎠ ⎝ ∗ ∗ ⎠
Mz (t) Mz cos ω1 t − My sin ω1 t

We can also describe the effect of a pulse of any phase on a Cartesian basis
1 1
⎛ ∗ ∗ ∗ ∗ ⎞
∗ Mx (1 + cos ω1 t) + (Mx cos 2ϕ + My sin 2ϕ)(1 − cos ω1 t) + Mz sin ϕ sin ω1 t
Mx (t)
⎛ ⎞
⎜ 2 2 ⎟
∗ ⎜ ⎟
⎜ My (t) ⎟ = ⎜ 1 1 ⎟ (20)
∗ ∗ ∗ ∗
⎜ My (1 + cos ω1 t) − (My cos 2ϕ + Mx sin 2ϕ)(1 − cos ω1 t) + Mz cos ϕ sin ω1 t ⎟
⎝ ∗ ⎠ ⎜ 2 2 ⎟
Mz (t)
⎝ ∗ ∗ ∗ ⎠
Mz cos ω1 t − (Mx sin ϕ − My cpsϕ) sin ω1 t

Steady State Approximation

6/5/2021 3 https://chem.libretexts.org/@go/page/1820
If the movement of the magnetization is slow, then
∗ ∗
dMx dMy dMz
= = =0 (21)
dt dt dt

When can then solve for the magnetizations in each direction


2
2πγ B1 M0 T Ω
∗ 2
Mx = − (22)
2 2 2 2 2
1 + 4π T Ω +γ B T1 T2
2 1

γ B1 M0 T2

My = − (23)
2 2 2
1 + 4π T Ω2 + γ 2 B T1 T2
2 1

2 2 2
M0 [1 + 4 π T Ω ]
2
Mz = (24)
2 2 2 2 2
1 + 4π T Ω +γ B T1 T2
2 1

Free Precession
In the absences of any magnetization applied in the x-y plane, the change in the magnetization is due to relaxation effects of T1
and T2 processes.
The Bloch equations in the rotating fram then reduce to

dMx (t)
∗ Mx (t)
⎛ ⎞ ⎛ ∗ ⎞
−(ωrot − ω0 )My (t) −
dt ⎜ T2 ⎟
⎜ ⎟
⎜ ⎟
⎜ ∗ ⎟ ∗
⎜ dMy (t) ⎟ ⎜ My (t) ⎟
⎟ =⎜ ⎟

⎜ (ωrot − ω0 )Mx (t) − (25)
⎜ ⎟
⎜ dt ⎟ ⎜ T2 ⎟
⎜ ⎟ ⎜ ⎟
⎜ ∗ ⎟ ⎜ ∗ ∗ ⎟
dMz (t) −[ Mz (t) − Meq ]
⎝ ⎠ ⎝ ⎠
dt T1

which can be solved to give


−t
⎛ ⎞

sin Ωt] e T2
∗ ∗
⎜ cos Ωt − [ Mx (0) My ⎟

Mx (t) ⎜ ⎟
⎛ ⎞ ⎜ −t ⎟
∗ ⎜ ⎟
⎜ My (t) ⎟ = ⎜ ⎟ (26)
⎜ [ M (0) cos Ωt + M sin Ωt] e T2
∗ ∗ ⎟
y x
⎝ ∗ ⎠ ⎜ ⎟
Mz (t) ⎜ ⎟
⎜ −t −t ⎟

⎝ ⎠
Mz (0)e T1 + Meq (1 − e T1

Bloch Equations in the Spherical Reference Frame


The effect of an RF pulse on the bulk magnetization using a spherical basis set is important concept. Assuming we have an
arbitrary phase of the pulse ϕ then the effect of the pulse is
+1∗ ∗ iϕ −1∗ 2iϕ
M (1 + cosω1 t) M e (isinω1 t) M e (1 − cosω1 t)
+1∗ 0
M = − + (27)

2 √2 2

+1∗ −iϕ −1∗ iϕ


M e (isinω1 t) M e (isinω1 t)
0∗ ∗
M = +M cosω1 t − (28)
– 0 –
√2 √2

+1∗ −i2ϕ ∗ −iϕ +1∗


M e (1 − cosω1 t) M e (isinω1 t) M (1 + cosω1 t)
−1∗ 0
M = + + (29)

2 √2 2

or more simply
1

pi ∗ pf ∗ −iΔpϕ
M = ∑ xp , pf (t)M e (30)
i

pf =−1

where x p0 , p1 (t) is the efficiency of the transfer of magnetization of from pi to pf and Δp = p f − pi

6/5/2021 4 https://chem.libretexts.org/@go/page/1820
The evolution of the magnetization when we consider the effects of relaxation then
p∗ p∗ ipΩt
M = y(t)M e (31)

where y(t) is the decay of the coherence.

References
Outside Links
http://bouman.chem.georgetown.edu/nmr/bloch/bloch.htm
http://mrsrl.stanford.edu/~brian/bloch/
http://www.chm.davidson.edu/vce/NMR/...Equations.html
comp.uark.edu/~jgeabana/blochapps/index.html

Contributors and Attributions


Derrick C. Kaseman (UC Davis)

6/5/2021 5 https://chem.libretexts.org/@go/page/1820
Larmor Precession
When placed in a magnetic field, charged particles will precess about the magnetic field. In NMR, the charged nucleus, will
then exhibit precessional motion at a characterisitc frequency known as the Larmor Frequency. The Larmor fequency is
specific to each nucleus. The Larmor fequency is measured during the NMR experiment, as it is dependent on the magnetic
field that the nucleus experineces.

Spinning Top Analogy


Often it is difficult in NMR to understand the microscopic processes that are occurring. However, precession is easily observed
on the macroscopic scale, as toy tops. When a top is spun, it rotates about a central axis (Figure 1). The angular momentum of
the top (L) is aligned along this central axis. If the top is set at an angle, the central axis will move in a circle. The top now
spinning along its own central axis precesses around in a circle around earths gravitational field.

Figure 1: This diagram explains the precession of a top. The torque created by the force of gravity applied at the center of
gravity and the reactive force applied where the top touches the table causes the top to precess. Image used wtih permision
from Wikipedia (Credit: Xavier Snelgrove).
Atomic nuclei contain intrinsic spin. The nucleus, like a top, will spin along an axis, which is the direction of the angular
momentum for the nucleus. The spin of the nucleus can be related to the magnetic moment of the nucleus through the relation
μ = γI (1)

where
μ is the magnetic moment and
γ is a proportionality constant known as the gyromagnetic moment.
This constant may be positive of negative, depending on if the nucelus precesses clockwise or counterclockwise, respectively.
The nuclear magnetic moment will couple to the external magnetic field, which produces a torque on the nucleus and causes
the precession around the magnetic field. This is analgous to the macroscopic tops in that the gravitational force couples with
the mass of the top. In the absence of friction, the top would precess forever! The frequency of precession is known as the
Larmor frequency, ν where0

ν0 = γ B0 (2)

The effect is illustrated below:

5/28/2021 1 https://chem.libretexts.org/@go/page/1822
Ensemble Effects
The net magnetization for a sample is the sum of the individual magnetic moments in the sample

M = ∑ μi (3)

we have already defined the magnetic moment for a nucleus with spin I. The magnetization can then be written as

M = γJ (4)

where J is the net spin angular momentum.


The torque, T, of the sample will then be
dJ
T = (5)
dt

Subsituting M for J we obtain


T = M xB (6)

and finally
dM
= γM xB (7)
dt

then
dM
= γM Bsinθ (8)
dt

since M and B are parallel the sin term drops out. We want to know the rate at whcih the magnetization is changing with
respect to time so we take the second derivative and the result is the Larmor frequency

ω0 = γB (9)

since M and B are both vector quantities, the cross product with B in only the Z direction i.e. (B=(0, 0, B0)) then we obtain the
Larmor frequency
ω0 = γ B0 (10)

References
1. Duer, M.J., Solid State NMR Spectroscopy: Principles and Applications. Blackwell Science Ltd. USA. 2002
2. Fukushima, E., Roeder, S.B.W., Experimental Pulse NMR A Nuts and Bolts Approach. Perseus Books Publishing, USA.
1981

5/28/2021 2 https://chem.libretexts.org/@go/page/1822
Contributors and Attributions
Derrick Kaseman (UC Davis)

5/28/2021 3 https://chem.libretexts.org/@go/page/1822
NMR Interactions
When an atomic nucleus is placed in a magnetic field, the ground state will split into different energy levels proportional to the
strength of the magnetic field. This effect is known as Zeeman splitting. While the Zeeman interaction is useful for identifying
different types of nuclei placed in magnetic fields, structural and dynamic information may be obtained by considering other
magnetic and electronic interactions coupling with the nucleus. These interactions are perturbations to the Zeeman interaction.
The full NMR Hamiltonian may therefore be expressed as
^ ^ ^ ^ ^ ^
H = H Zeeman + H J + H C S + H DD + H Q (1)

where HZeeman is the Zeeman interaction, HJ is the J coupling, HCS is the chemical shift coupling, HDD is the dipolar coupling,
and HQ is the quadrupolar coupling. The relative magnitude of these interactions is shown in the table below. The Zeeman
interaction is the largest, followed by the quadrupolar interactions which are on the order of MHz. The chemical shift and the
dipolar coupling are on the order of kHz while the scalar coupling is the smallest which is only tens of Hz. Clearly, some of
these interactions are more pronounced than others.
Interaction Magnitude (Hz)

Zeeman 108

Quadrupolar 106
Chemical Shift 103
Dipole 103
J 10

Table 1. Magnitude of different NMR interactions


In the liquid state, the dipolar and anisotropic contribution to the chemical shift are averaged due to the molecular reorientation
occurring in liquids. The averaging of these interactions gives the characteristically narrow isotropic peaks. Additionally, liquid state
NMR primarily looks at spin 1/2 nuclei (13C, 1H) which eliminates any quadrupole interactions. Only the J coupling and isotropic part
of the chemical shift remains. In the solid state, molecular reorientation does not occur and the solids may have a variety of bond
lengths and angles of a given chemical site. These factors broaden in the NMR spectrum with the broadest peaks over 1MHz wide!

Contributors and Attributions


Derrick Kaseman (UC Davis)

6/18/2021 1 https://chem.libretexts.org/@go/page/1823
Chemical Shift (Shielding)
The chemical shift in NMR is extremely important, as it gives vital information about the local structure surrounding the
nucleus of interest. For a majority of scientists, the chemical shift is used exlusivley to determine structure, especially in
organic systems. Additional information may be gained by examining the anisotropy of the chemical shift. This section will be
devoted to looking at chemical shift from a mathematical standpoint including a full treatment of the chemical shift tensor and
the relation to the NMR lineshape.

Shielding and Chemical Shift


As electrons orbit the nucleus, the slightly alter the magnetic field that the nucleus experineces, which slightly changes the
difference between the energy levels which gives the resulting spectra. However, these changes are from a select reference and
therefore are relative. The resulting change in the energy levels is on the order of Hz, while the Zeeman interaction is on the
order of MHz. Thus we can develop a scale based on the relative changes in the energy levels.
Beginning with the equation for Zeeman splitting
ΔE = −γℏB0 (1)

the effect of shielding σ, results in


ΔE = −γℏ(1 − σ)B0 (2)

in which the quantitiy (1 − σ)B is known as the effective field experineced by the nucleus
0 Bef f . Relating the change in
energy to a frequency using
E = hν (3)

The value for the shielding is based off the resonant frequency for some reference sample, such as tetrmethylsilane (TMS),
σref
. This can cause confusion as the strength of the magnetic field dictates the exact frequency is field dependent. To make
the shielding constant between magnetic fields it becomes customary to diving by the resonant frequency of the given nucleus.
This gives values on the order of 10 or ppm. This value is known as chemical shift, δ , and is given by
−6

6
δ = 10 (σref − σsample ) (4)

Chemical Shift Tensor


Consider a system consisting of a single spin 1/2 nucleus; I. H
^
CS may then be represented as
^
γh I σ B0
^
H cs = − (5)

where γ is the gyromagnetic ratio, h is Plank's constant, I^ is the spin operator, and σ is the chemical shift tensor. This tensor
may be represented as
σxx σxy σxz
⎛ ⎞

σ = ⎜ σyx σyy σyz ⎟ (6)

⎝ ⎠
σzx σzy σzz

The principle axis system (PAS), denoted by axes XPAS, YPAS, ZPAS,represented in the figure below. The PAS is the x, y, and z
coordinates with respect to the nucleus whereas X, Y, and Z of the rotating frame are defined with B0 along the +Z direction.
The chemical shift tensor describes the electric field surrounding the nucleus with the principle components of the tensor; σ , xx

σyy,σ .zz

5/27/2021 1 https://chem.libretexts.org/@go/page/1824
Figure 1. Graphical representation of the relation of the Principle Axis System (dotted blue lines) to the laboratory frame
(black solid lines). The projection of the PAS onto the laboratory frame is indicated by the black dotted line. The principle
components of the chemical shift tensor are illustrated using the red lines, which the black ellipses represent the electron field
surrounding the nucleus.
In a solid, the sample is not a single spin oriented along B0. Rather, the sample contains orientations sampling all directions in
all 3 dimensions. Each magnetic moment of the system can be related for its PAS to the Z axis by an angle, θ , and its position
in the x y plane given by angle ϕ . The magnetization experienced at the nucleus on the basis of the PAS is given by
P AS
B = (sinθcosϕ, sinθsinϕ, cosθ) (7)
0

.
The chemical shift, ω , is then
cs

2 2 2 2 2
ωcs = −ω0 (σxx si n θcos ϕ + σyy si n θsi n ϕ + σzz cos θ) (8)

,
which reduces to
1
2 2
ωcs = −ω0 σiso − ω0 Δ[3cos θ − 1 + ηsi n θcos2ϕ] (9)
2

.
νiso
−ω0 σiso is the isotropic frequency, ωiso = , routinely observed in liquid spectra.

Chemical Shift Anisotropy (CSA)


If a polycrystalline sample is then placed in a spectrometer, a lineshape similar to these may be observed:

5/27/2021 2 https://chem.libretexts.org/@go/page/1824
Figure 2. Comparison of the static NMR signal with changing Δ and η. The dashed lines are used to show how the principle
components the chemical shielding tensor relate the the lineshape.
This is due to asymmetry of the local electronic environment surrounding the nucleus. If the electronic cloud were symmetric
then σ = σ = σ and Δ and η would be zero leaving only the isotropic peak. Note how broad this pattern is when
xx yy zz

compared to liquid spectra! CSA is not a large factor in liquids becasue rapid molecular tumbling averages the CSA during the
experimental timescale. The above patterns can be deconstructed into the principle components of the chemical shift tensor, as
indicated on the figure. This is especially useful as NMR directly correlates to the immediate local electronic structure
surrounding the nucleus probed. Take for example a compound which has σ = σ , which is said to be axially symmetric
xx yy

and η = 0 . It is immediately obvious that there is symmetry about the nucleus which is reflected in the static lineshape (black
line in above figure).

CSA Conventions
There are three major conventions people use to describe the CSA tensor. They are outlined below and once should always
keep in mind the difference between shielding and chemical shift scales.
IUPAC Convention

Figure 3. CSA tensor as described by the IUPAC convention

5/27/2021 3 https://chem.libretexts.org/@go/page/1824
The IUPAC convention describes the CSA tensor by three values, called principle components of the CSA tensor, δ 11 , δ22 , δ33 .
These values follow the following magnitude
δ11 ⩾ δ22 ⩾ δ33 (10)

.
The average value or these is then the isotropic chemical shift.
δ11 + δ22 + δ33
δiso = (11)
3

Herzfeld Berger Convention

Figure 4. CSA tensor as described by the Herzfeld Berger convention


The Herzfeld Berger convention uses the IUPAC definition of the principle components, but they are represented using a
different notation, the span, Ω, and the skew, κ . They are defined as follows:
δ11 + δ22 + δ33
δiso = (12)
3

Ω = δ11 − δ33 (13)

3(δ22 − δiso )
κ = (14)
Ω

Ω will always be larger than or equal to 0 and κ will range from -1 to 1


Haeberlen Convention

Figure 5. CSA tensor as described by the Haeberlen convention


The Haberlen convention uses different combinations of the principle components to describe the CSA. They are
| δzz − δiso | ⩾ | δxx − δiso | ⩾ | δyy − δiso | (15)

δ11 + δ22 + δ33


δiso = (16)
3

Δ = δzz − δiso (17)


δ = (18)
2

δxx − δyy
η = (19)
δzz − δiso

Conversion Between Conventions


Converting between the standard and Haeberlen convention is often needed to compare values of the CSA. There are two
cases, one in which Δ > 0 and when Δ < 0 . When Δ > 0

5/27/2021 4 https://chem.libretexts.org/@go/page/1824
δ11 = δiso + Δ (20)

Δ(1 − η)
δ22 = δiso − (21)
2

Δ(1 + η)
δ33 = δiso − (22)
2

When Δ > 0
δ33 = δiso + Δ (23)

Δ(1 − η)
δ22 = δiso − (24)
2

Δ(1 + η)
δ11 = δiso − (25)
2

Dynamics and CSA


The CSA is very sensitive to any type of motion occurring in the system. A great example of this is a liquid. Liquids only
exhibit a narrow line at the isotropic chemical shift, indicating that any CSA is averaged. The average timescale of rotation for
a liquid is on the order of picoseconds which is about 1 trillion times larger than the CSA . For solids the case is more
interesting as increasing the temperature can give rise to molecular motion. There are two types of motion isotropic tumbling
anf rotation about a fixed axis. We will consider the implications on the CSA lineshape for each case.
Isotropic Tumbling
Isotropic rotation, or isotropic tumbling motion is occurring if the molecule is rotating in all possible directions. Therefore at
one time a molecules spin may be aligned with the magnetic field, and then at the next time point be oriented perpendicular to
the external magnetic field. Correspondingly, we can begin to envision that the processes of isotropic tumbling closely mirrors
that of chemical exchange at a site in which each orientation of the spin may be represented as a single peak.

Figure 6: The effect of isotropic rotation on a CSA lineshape. The CSA for this site is \(\Delta = 200ppm\), \(\eta=0.5\), and
δiso= 0 . The rotation speeds, τ is given on the left of each spectrum

As the orientation of the spin changes, the peak position will change. Thus as the spins begin to change their orientation faster
than width of the CSA, they will come to a average position, which will be the isotropic chemical shift. Mathematically this
can be described as follows.
Assume there is a single crystal with a single CSA. The rotational jump is of random nature and occurs at a single rate
constant, κ . The jump is from site Ω → Ω and spends an average time at any given site of τ . We also assume there is no

correlation between Ω and Ω . Assuming a Markoff process for the time dependence, then

5/27/2021 5 https://chem.libretexts.org/@go/page/1824
−1
κ =τ (26)

.
The probability of finding a spin at orientation Ω after a time t from an initial position of Ω is 0

d ′ ′
P (Ω0 /Ω, t) = ∫ π(Ω , Ω)P (Ω0 /Ω, t)dΩ (27)
dt

where the initial probability is then


P (Ω0 /Ω, t) = δ(Ω − Ω0 ) (28)

and δ is the Kroenecker Delta function and π(Ω , Ω) is a memory function that keeps the history of Ω

As expected, as the time approaches zero then the probability approaches the initial condition, or

P (Ω0 /Ω, t) = δ(Ω − Ω0 ) (29)

The memory function must then be the transition probability per unit time which is


d
π(Ω , Ω) = [ P (Ω0 /Ω, t)]t=0 (30)
dt

For a given time t, the probability of |Omega rotates to Omega is given by kappat then

′ ′
π(Ω , Ω) = (1 − κt)δ(Ω − Ω) + W (Ω)κt (31)

where W (Ω is the probability of finding the molecule in orientation Ω


then it can be shown
′ ′
π(Ω , Ω) = [W (Ω) − δ(Ω − Ω0 )] (32)

and for a fintie number ,k, sites


π(Ωm , Ωn ) = [W (Ωn ) − δmn ] (33)

The NMR spectrum will be the Fourier Transform of the FID, G(t) given by
1
G(t) = ∫ dΩG(t, Ω (34)
N Ω

where

G(t, Ω) = ∑ Qk (35)

for k sites
Assuming a line is generated for each k site at ω where k

ωk = ω(Ωk ) (36)

the the equation of motion is


d
= iw(Ωk Qk + ∑ π(Ωj , Omegak )Qj (37)
dt
j

Qk = Q(t, Kk ) (38)

The equation of motion can be expressed as a k-dimensional vector, K as


\[\dfrac{d}{dt}K=(k\omega+\pi)K
where π is the previously defined jump matrix and ω is the diagonal matrix with elements
ωkj = δkj ωkj (39)

5/27/2021 6 https://chem.libretexts.org/@go/page/1824
K can be solved to give
K(t) = K(0)exp[(iω + π)t] (40)

K(0) = W = (W (Ω1 ), W (Ω2 ), W (Ω3 ), W (Ωn )) (41)

Then
G(t, Ω) = W exp[(iω + π)t] (42)

and the FTed spectrum is


−1
I = (ω, Ω)RealW A (43)

A = i(ω − ωE) + π (44)

Rotation about an axis


In contrast to isotropic tumbling motion, molecules may only rotate about a single fixed axis. This will ultimately change the
CSA to a uniaxial pattens, once the rotation is fast enough. Depending on the Δ and η the averaged pattern may even switch
the sign of Δ!

Figure 7: The effect of rotation about a single axis on a CSA lineshape. The CSA for this site is Δ = 200ppm, \(\eta\)=0.5, and
δiso= 0ppm . The rotation speeds, τ is given on the left of each spectrum

Lets explore what happens when this is the case Mathematically. It is easiest to consider the case of symmetry related jumps
where a molecule performs a rotation at rate 1/τ between symmetry equivalent positions, ω . The exchange between sites
jk j

may then be written as

Ag = i ω1 M0 1 (45)

where A is a coupling matrix, g is the magnetization at site j, ω is the RF field strength and
1 M0 is the thermal equilibrium
magnetization. The exchange matrix is a diagonal matrix with elements
1 1
Ajj = i(ω − ωj ) − (46)
T2 − τj

we can the solve for g and forier transform to get the intensity ( an exercise left for the reader) to obtain the equation for the
lineshape

I (ω) = ∫ (g(ω, Ω)dΩ (47)


Ω

References
1. M. Mehring. High Resolution NMR spectroscopy in Solids. Springer Verlag Berlin 1976
2. J. Herzfeld, A. E. Berger, J. Chem. Phys. 1980, 73, 6021
3. J. Mason, Solid State Nucl. Magn. Reson. 1993, 2, 285
4. U. Haeberlen, In Advances in Magnetic Resonance; Suppl. 1; J. S. Waugh, Ed.; Academic Press: New York, 1976.
5. Massiot et al. Magnetic Resonance in Chemistry, 40 pp70-76 (2002)
6. D.E. Wemmer et al. J.American Chemical Society. 1981, 103, 28-33

5/27/2021 7 https://chem.libretexts.org/@go/page/1824
Outside Links
Tensor Calculator http://anorganik.uni-tuebingen.de/kl...ntions/csa/csa
Jump Frequency Spectra Generator http://weblab.mpip-mainz.mpg.de/cgi-...&version=4.5.1

Contributors and Attributions


Derrick C. Kaseman UC Davis

5/27/2021 8 https://chem.libretexts.org/@go/page/1824
Dipolar Coupling
Contributors and Attributions
Name #1 here (if anonymous, you can avoid this) with university affiliation

5/10/2021 1 https://chem.libretexts.org/@go/page/1825
J-Coupling (Scalar)
Page Under Construction

Introduction
A J-coupling is an interaction between nuclei containing spin. J-couplings are also known as scalar couplings. This interaction
is mediated through bonds, in contrast to dipole interactions, which are mediated through space. Typically, we consider the J-
coupling to be a weak interaction, in comparison to the Zeeman interaction. J-couplings are typically used in combination with
chemical shifts to deduce the through-bond connectivity in small molecules and proteins. While typically a liquid state
phenomena, solid-state J-coupling constants are observable. J-coupling values range in 0.1 Hz in organic compounds to kHz in
transition metal complexes. The J-coupling typically reduces in magnitude the more bonds exist between the coupled nuclei.
Furthermore, J-couplings may be either homonuclear (i.e. between hydrogens with different chemical shifts) or heteronuclear
(i.e. between hydrogen and carbon).

Pascals Triangle
Nuclei in different chemical environments up to nine bonds (though 1-4 are typically readily measurable) is away can
influence one another’s effective local magnetic field by carrying spin orientational information through bonding
electrons.This effect is most prominent among chemically equivalent nuclei, giving rise to the N+1 rule for equivalent protons.
A proton with N protons on contiguous carbon atoms splits into N+1 peaks with intensity pattern. The splitting pattern for A
when it is coupled to a number of X nuclides would follow the relation represented by Pascal’s triangle. To be more specific, I
will take AX2 for example. AX2 represent a spin system that contains three nuclei, two of which have the same chemical shift
and one of which is different (e.g. ClCH2CHCl2). Here A is CH proton and X are CH2 protons. According to table 4 and figure
12, CH proton will be split into 1:1 doublet, while CH2 will give a fine structure of 1:2:1 triplet. The spacing between peaks is
defined as coupling constant J, which can be used to describe the degree of coupling.
Table 4. Pascal's triangle according to AX configurations

6/29/2021 1 https://chem.libretexts.org/@go/page/1826
Figure 12.NMR spectrum of nucleus A in an AX2 spin system.

Pascals Trinagle Construction


The Pascal’s triangle is a graphical device used to predict the ratio of heights of lines in a split NMR peak. To construct the
Pascal’s triangle, use the following procedure.
Step 1: Draw a short, vertical line and write number one next to it.

Step 2: Draw two vertical lines underneath it symmetrically.

Step 3: Connect each of them to the line above using broken lines.

6/29/2021 2 https://chem.libretexts.org/@go/page/1826
Step 4: Each of the two lines is connected to the single line above, which carries the number one. Therefore, write number one
next to each line.

Step 5: Draw three vertical lines symmetrically underneath the two lines.

Step 6: Connect each of them to the nearest line(s) above.

Step 7: Each terminal line is connected to one line above, which carries the number one. Therefore, write number one next to
each of them. The internal line is connected to two lines above, each carrying the number one. 1 + 1 = 2; write number two
next to the internal line.

Continue the process as far down as necessary.

6/29/2021 3 https://chem.libretexts.org/@go/page/1826
see also (n+1) Rule, coupling constant

Strong Coupling and the Roof Effect


In the high field approximation, there may be strong coupling between homonuclear spins. Strong coupling refers to the
scenario when the chemical shift difference (in Hz) is on the order of the J-coupling. As chemical shift scales with the
magnetic field, the chemical shift differences are therefore smaller at low magnetic fields, makes this scenario more likely at
low magnetic fields. First we must define the strong coupling parameter ,θ
2πJ
tan2θ = (1)
ωA − ωB

where ω is in radians and J is in Hz.


[insert tan 2theta graph]
Then the J-coupling line intensities become for the doublet (A1,A2)
IA1 ∝ 1 − sin2θ (2)

IA2 ∝ 1 + sin2θ (3)

Just like with the usual J-couplings, having multiple J-couplings will lead to a complex splitting pattern, where each intensity
is modified by the angle as such.

Expansion of the J-Coupling Hamiltonian


Typically, the J-coupling is given in the weak coupling limit. The weak coupling limit is defined by having ν − ν >> J ,
I s

which is typically valid at high magnetic fields of several tesla. However, here we will derive the J-coupling Hamiltonian for a
2 spin system without any assumptions about the coupling limit.
First we begin with the J-coupling Hamiltonian which constains the pertenent interactions; Zeeman and Scalar. Then the
Hamiltonian will be
ωI IZ + ωS SZ + 2πJ(I ⋅ S)

The factor 2π in front of the J-coupling keeps all units in angular frequency, as J is measurable in Hz. Remember that for a 2
spin system, each nucleus has an α and β state. Thus the potential wave functions in matrix form are given by the α and β
states. In Matrix form this would be represented for a 2 spin system as
|αα > |αβ >
[ ] (4)
|βα > |ββ >

Now the α and beta states also have matrix representations as


1
|α >= [ ] (5)
0

6/29/2021 4 https://chem.libretexts.org/@go/page/1826
0
|β >= [ ] (6)
1

and the combined states are given by cross product of the 2 matricies so
1
⎡ ⎤

1 ⎢0⎥ 1
|αα >= [ ]×[ ] =⎢ ⎥ (7)
0 0 ⎢0⎥

⎣ ⎦
0

0
⎡ ⎤

0 ⎢0⎥ 0
|ββ >= [ ]×[ ] =⎢ ⎥ (8)
1 1 ⎢0⎥

⎣ ⎦
1

0
⎡ ⎤

1 ⎢1⎥ 0
|αβ >= [ ]×[ ] =⎢ ⎥ (9)
0 1 ⎢0⎥

⎣ ⎦
0

0
⎡ ⎤

0 ⎢0⎥ 1
|βα >= [ ]×[ ] =⎢ ⎥ (10)
1 0 ⎢1⎥

⎣ ⎦
0

Then without further proof (at the moment) the eigenfunctions of the 2 spin system are
ψ1 = |αα > (11)

ψ4 = cosθ|αβ > +sinθ|βα > (12)

ψ3 = cosθ|βα > −sinθ|αβ > (13)

ψ4 = |ββ > (14)

Remember that according to Schroedinger's Equation,


^
H ψ = Eψ (15)

thus if we apply the Hamiltonian Operator on each of the eigenstates, we should be able to deduce the energy level. Now
looking at the Hamiltonian, there are I and S operators which have the following matrix representations
1 1 0
Iz = Sz = [ ] (16)
2 0 −1

1 0 1
Ix = Sx = [ ] (17)
2 1 0

i 0 −1
Iy = Sy = [ ] (18)
2 1 0

However since we are considering a 2 spin system, 2x2 matrices won't cut it. Rather, we need to transform these matrices into
the product basis by calculating the direct products of the wavefunctions. To do this we recognize that we can obtain 2N,
wavefunctions where N is the total number of spins. To do this

ψk = | m1 > ×| m2 >. . . ×| mN > (19)

Thus for \psi_1


ωI IZ + ωS SZ + 2πJ(I ⋅ S)

Limits of J-Couplings
6/29/2021 5 https://chem.libretexts.org/@go/page/1826
There exists certain limits in which the J-coupling can be calculated. We begin our discussion investigating 13C-labeled
methanol (13C-MeOH). Lets assume that we are in the high-temperature regime, such that the OH proton is uncoupled to the
methyl spins during the experimental timescale. The J-coupling between the 3 methyl protons and the 13C nucleus is
1
JHC=140.5 Hz.

High Field Limit


At high magnetic fields, the line width is governed by the magnetic field inhomogeneity. That is when a superconducting
magnet is build, the field is not completely uniform over the sample volume. Therefore, there is a distribution of chemical
shifts for a given site. Since the field inhomogeneity is typically small, (parts per billion or ppb), this leads to a broader peak
than a perfectly homogenous field. This is the reason in which shimming is performed prior to sample collection. Shimming
makes the field more homogenous. If we assume that the magnetic field inhomogeneity (Binhomo) of the external magnetic
field (B0), is 25 ppb over the sample volume after shimming, then we can calculate the line width Δν of a single peak.
Δν = B0 Binhomo

Assuming B0 is 11.74T (500 MHz spectrometer), then Δν is 2.1 Hz (~0.3 ppm). Of course there are geometrical constraints
about the magnetic field inhomogenity, such as the fraction of sample that is in the inhomogenous region, but for now we
assume that the equal portions of the sample are in the inhomogenous field. Thusly, J-couplings larger than 3.45 Hz could not
be resolved. Of course this is ridiculous for a 500 MHz instrument with good shimming. A much more reasonable

Figure x: Magnitude of J-coupling that is un-resolvable J-couplings as a function of magnetic field, assuming a Binhomo=1.5
ppb. The regions of couplings that are resolvable are marked. Note that this assumes 1H
value would be not resolving J-couplings less than 0.2 H, which would be a field inhomogeneity of 0.10 ppb. Currently, the
largest sustained magnetic fields are on the order of 100T (pulsed, National High Magnetic Field Lab) and ~30T (static).
Therefore, the high field limit (\B_{HF}) can be described as
BH F = J/(B0 Δν γ)

or
2
BH F = J/(B Binhomo γ)
0

where γ is the gyromagnetic ratio in Mhz/T and Δν is in ppm.


Thus, for the same field inhomogenity, the resolvable J-couplings are smaller for small γ nuclei.

6/29/2021 6 https://chem.libretexts.org/@go/page/1826
Figure x. Resolvable J-couplings for 1H (black) and 13C (blue) for Δν of 1.0 ppb (solid) and 2.0 ppb (dashed)

High Field Weak Coupling Limit


When the magnetic field is less than BHF, the heteronuclear J-couplings can be resolved. This leads to the well-known (2I+1)
multiplet splitting. For spin I=1/2 (i.e. 1H, 13C, 31P) this leads to the well known pascal triangle. However, for coupling to
spins I>1/2 (i.e. 14N, 2H, 7Li, 51V) the splitting patterns become more complex. As the J<<ν we are in the weak coupling
limit.

Low Field Strong Coupling Limit


2
J
B
LowF ield
1→2
= First order Perturbation
2Δν (γI − γS )

−−−−−−−−−−−−
2
3J
LowF ield
B =√
2→3 2
4Δν (γI − γS )

1
(1 + – )J
√2
LowF ield
B =
3→exact
(γI − γS )

Ultra Low Field Strong Coupling Limit


4J
Ultra−LowF ield
B =
exact→2
(7 γI − γS )

Ultra-Low Field Weak Coupling Limit


−−−−−−−−−

Ultra−LowF ield
2Δν J
B =√
2→1 2
(γI − γS )

Zero Field Weak Coupling Limit


2Δν J
ZeroF ield
B =
1→0
(γI + γS )

Outside Links
This is not meant for references used for constructing the module, but as secondary and unvetted information available at
other site
Link to outside sources. Wikipedia entries should probably be referenced here.

Problems
1) Calculate the multiplicity for CH3 for both C and H.
2) Calculate the multiplicity for HCl for both H and Cl.

6/29/2021 7 https://chem.libretexts.org/@go/page/1826
Contributors and Attributions
Derrick Kaseman,
Gamini Gunawardena from the OChemPal site (Utah Valley University)

6/29/2021 8 https://chem.libretexts.org/@go/page/1826
Pascal’s Triangle Construction
The Pascal’s triangle is a graphical device used to predict the ratio of heights of lines in a split NMR peak. To construct the
Pascal’s triangle, use the following procedure.
Step 1: Draw a short, vertical line and write number one next to it.

Step 2: Draw two vertical lines underneath it symmetrically.

Step 3: Connect each of them to the line above using broken lines.

Step 4: Each of the two lines is connected to the single line above, which carries the number one. Therefore, write number one
next to each line.

Step 5: Draw three vertical lines symmetrically underneath the two lines.

Step 6: Connect each of them to the nearest line(s) above.

5/29/2021 1 https://chem.libretexts.org/@go/page/206431
Step 7: Each terminal line is connected to one line above, which carries the number one. Therefore, write number one next to
each of them. The internal line is connected to two lines above, each carrying the number one. 1 + 1 = 2; write number two
next to the internal line.

Continue the process as far down as necessary.

see also (n+1) Rule, coupling constant

Contributors and Attributions


Gamini Gunawardena from the OChemPal site (Utah Valley University)

5/29/2021 2 https://chem.libretexts.org/@go/page/206431
Quadrupolar Coupling
1
Nuclei with spin I > comprise more that 2/3 of the NMR active nuclei. These nuclei exhibit a quadrupolar moment which
2
couples to the electric field gradient resulting in extensive peak broadening. This page is dedicated to understanding the
origins of the quadrupole moment and the effects on the NMR line shape. It will be developed through a heavy mathematical
treatment, however, the illustrations and captions will provide a pictorial representation of the mathematical treatment.

The Quadrupole Moment


Understanding a Quadrupole
To fully understand the quadrupole interaction we must first establish what a quadrupole is. Quite simply, a quadrupole can be
thought of as two dipoles. Unlike a dipole however, the quadrupole will not couple to a symmetric field as the forces and
subsequent torques on the quarupole will cancel.

A Quadrupole
If there is an non-symmetric field there will be a force on the quadrupole, i.e., an electric field gradient. We can then define the
quadrupole moment as the tendency of the quadrupole to rotate about an axis. Due to the 3D nature of a quadrupole it may be
described by a second rank tensor Q where

Qxx Qxy Qxz


⎡ ⎤

Q = ⎢ Qyx Qyy Qyz ⎥ (1)

⎣ ⎦
Qzx Qzy Qzz

The quadrupole can then couple to an Electric Field Gradient (EFG) The electric field gradient is denoted a V and is also
described by a second rank tensor.

Vxx Vxy Vxz


⎡ ⎤

V = ⎢ Vyx Vyy Vyz ⎥ (2)

⎣ ⎦
Vzx Vzy Vzz

EFGs are generated in solids and liquids by th electrons in the sample.

Quadrupolar Nuclei
Within the nucleus of an atom, the protons, and subsequent charge of the nucleus, can be distributed symmetrically or
asymmetrically. If the charge distribution is symmetric, the spin,I, of the nucleus is 1/2 and the interaction of the nucleus with
electric field gradients is direction independent. However, if the charge distribution is asymmetric I>1/2, and the electric field
gradient can interact with the nucleus and exhibit a torque on the nucleus. These nuclei are known as quadrupolar nuclei. It is
worth mentioning that the electric field gradient is generated by the electrons present in the sample. Consequently, these nuclei
exhibit a quadrupole moment, Q.

Figure 1: Charge distribution is spherical and non-spherical nuclei


While calculations of the quadrupole moment for a given nuclei are beyond the scope of this page, the moments for the nuclei
have been calculated and a few examples are listed below.

7/4/2021 1 https://chem.libretexts.org/@go/page/1827
Nucleus Spin Q (barns) X 103
2H 1 2.86
6Li 1 0.83
7Li 3/2 -40.6
10B 3 84.7
17O 5/2 -25.7
87Rb 3/2 127.1

Q can be considered a friction coefficient between rotations of the electric field of the molecule. The larger the Q value, the
more strongly the asymmetric nucleus will interact with a non-uniform electric field gradient. This leads to a nuclear spin
reorientation in the nucleus. The exception is cubic symmetry (Td or Oh) where the electric field gradient is symmetic resulting
in no net effect on the non-spherical nucleus.

Spin Energy Levels


As the spin of a quadrupole nucleus is larger than I=1/2 we develop multiple energy levels (2I+1) and therefore multiple
transitions are expected. In half-integer spin systems (e.g. 3/2, 5/2, 7/2) There is still a -1/2 to 1/2 transition knowns as the
central transition. The other transitions are known as satellite transitions. A good example is listed below:
Boron-11 (11B) is a classic example of the spectral changes caused by the quadrupole moment. With a nuclear spin of I = 3

and Δm = ±1 transitions being allowed, 4 states are produced: m = , , - , - . However, only 3 transitions are possible:
3

2
1

2
1

2
3

3 1
↔ (3)
2 2

Satellite Transition
1 1
↔ − (4)
2 2

Central Transition
1 3
− ↔ − (5)
2 2

Satellite Transition
While it may be expected that these transitions happen at the same energy levels, we we see in a latter section that this is not
the case!

Derivation of the Hamiltonian


If we consider the picture below, We can describe the electrostatic interaction between the electron, which has a non-spherical
charge distribution, and the protons of the nucleus through a coulombic interaction such that
Z 2
e
U = −∑ (6)
→ →
p=1 | rp − re |

where Z is the atomic number, and rp and re are the proton and electron distances.
(insert picture)
Most likely, the PAS of the rp and re will not coincide with the lab frame and instead should be expressed by the principle axis
system of angles θ , ϕ , θ as denoted in the figure above. Converting the reference frame to a frame with its origin at the
p p e

center of the nucleus we can obtain


∞ l l
1 1 r< (l)∗ (l)
= 4π ∑ ∑ Ym (θp , ϕp )Ym (θe , ϕe ) (7)
→ → 2l + 1 rl+1
| rp − re | l=0 m=−l >

7/4/2021 2 https://chem.libretexts.org/@go/page/1827
where r< is the smaller value between rp or re and r> is the larger value, and Y denotes the spherical harmonics. In the case that
the electrons do not penetrate the nucleus, re>rp and
Z ∞ l l
1 rp (p)∗ (p
2
U = −4π e ∑∑ ∑ Ym (θp , ϕp )Ym (θe , ϕe ) (8)
2l + 1 re l + 1
p=1 l=0 m=−l

From thie above equation, we will derive the Hamiltonian for the quadrupolar interaction, HQ. In order to do this we first start
by showing the symmetry relation
(l) m (l)
Ym = (−1 ) Y−m ∗ (9)

Then,
l l
(l)∗ (l) m (l)∗ (l) (l) (l)
∑ Ym (θp , ϕp )Ym (θe , ϕe ) = ∑ (−1 ) Y−m (θp , ϕp )Ym (θe , ϕe ) = Y (θp , ϕp ) ⋅ Y (θe , ϕe ) (10)

m=−l m=−l

The dot product infers that we can then separate the potential energy equation into two pieces, one for the nuclear energy
denoted as Q(l) and one for the electronic energy as k(l). These expressions are given as
Z −−−−−−

(l) l (l)
Q = e∑√ rp Y (θp , ϕp ) (11)
2l + 1
p=1

−−−−−−
4π 1
(l) (l)
k = −e√ Y (θp , ϕp ) (12)
2l + 1 rl+1
e

The total potential energy is then


(l) (l)
U = ∑Q ⋅k (13)

for l=0
2
Ze
U =− (14)
re

which is the expected coulombic interaction between the electron and the nucleus. The next higher order (l=1) corresponds to
the interaction between a nucleus electric dipole moment and the electric field generated by the electrons. Since rp is an odd
operator, the expectation value of Q is zero andthe potential energy, U, is also zero. The next higher interaction is l=2
corresponding to electric quadrupole interaction.
(2) (2) (2)
HQ = U =Q ⋅k (15)

In order to fully expand this Hamiltonian, we want to relate Q and K to the molecular parameters. Q will be related to only the
operators involving the nuclear spin such that
(2) (2)
Q = AT (16)

where T is the spherical tensor. K will be related to the EFG. Lets begin our discussion by expanding Q. Consider the effect of
Q on the spins such that
−−

(2) 4π (2)
2
⟨I , mI=I | Q |I , mI=I ⟩ = e√ ∑ rp ⟨I , I | Y (θp , ϕp )|I , I ⟩ (17)
0 0
5
p

1
2 2
= e⟨I , I | ∑(3 zp − rp )|I , I ⟩ (18)
2
p

In which su m 2
p (3 zp
2
− rp ) is the quadrupolar moment. Therefore this term reduces to
1
= eQ (19)
2

7/4/2021 3 https://chem.libretexts.org/@go/page/1827
Using the irreducible tensors,
(2)
= A⟨I , I | T |I , I ⟩ (20)
0

1 2
2
= A⟨I , I | – (3 I0 − I )|I , I ⟩ (21)
√6

A
2
= (3 I − I (I + 1)) (22)

√6

A
= I (2I + 1) (23)

√6

We can then solve for A




3 eQ
A =√ (24)
2 I (2I − 1)

which yeilds


(2)
3 eQ (2)
Q =√ T (25)
2 I (2I − 1)

Adopting a similar approach for k


2 2
d −e 3 ri rj − r δij
( ) = −e = Vri rj (26)
dri drj p r5

Then
−−

4π 1 (2)
2
k = −e√ Y (θp , ϕp ) (27)
0 3 0
5 re

2 2
−e 3 Ze − re
= (28)
5
2 re

1
Vzz (29)
2

We can expand k(2) and find




(2) 1 2
k =∓ √ (Vzx ± i Vyz ) (30)
±1
2 3



(2) 2
k = df rac14 √ (Vxx − Vyy ± 2i Vxy ) (31)
±2
3

Substituting the equations in for k and Q we would be able to obtain HQ in an arbitary frame. However, it is much more
convient to express HQ in terms of the principle axis system of the EFG. Therefore, we choose the following condition
Vi,j = 0 (32)

unless i=j. Resulting in


(2)
k =0 (33)
±1

and we obtain our expression for the HQ as


eQ 1 1 1
2 2 2 2
HQ = [ I (Vxx − Vyy ) + (3 I −I )Vzz + I (Vxx − Vyy ) (34)
+1 0 −1
I (2I − 1) 4 4 4

eQ 2
2 2 2
= [(3 I −I )Vzz + (I +I )(Vxx − Vyy ) (35)
0 +1 −1
4I (2I − 1)

7/4/2021 4 https://chem.libretexts.org/@go/page/1827
We can further simplify this expression by converting the electronic operators (Vij) by defining
Vzz = eq (36)

and
Vxx − Vyy = ηeq (37)

Yielding
2
e Q 2
2 2 2
HQ = [(3 I −I ) + (I +I )η (38)
0 +1 −1
4I (2I − 1)

Furthermore,
2 2 2 2
I +I = Ix − Iy (39)
+1 −1

and
2 2 2 2
I = Ix + Iy + Iz (40)

Then
2
e Q
P AS 2 2 2 2 2
H = [2 Iz − Ix − Iy + η(Ix − Iy )] (41)
Q
4I (2I − 1)

η −1 0 0 Ix
2 ⎡ ⎤⎛ ⎞
e Q
= (Ix , Iy , Iz ) ⎢ 0 −η − 1 0 ⎥ ⎜ Iy ⎟ (42)
4I (2I − 1)
⎣ ⎦⎝ ⎠
0 0 2 Iz

Then HQ reduces to
2
e Q
P AS ⃗  ⃗ 
H = I ⋅Q⋅I (43)
Q
4I (2I − 1)

Expansion of the Hamiltonian


Classical Expansion
For those who wish to not delve into the complex treatment of the quadrupolar hamiltonian we can treat the Hamiltonian semi-
classically and derive an expression for the quadrupolar Hamiltonian. The interaction of a quadrupole with a field gradient in
an arbitrary frame (The PAS of the of the electric field gradient) may be described by
eQ
^ ^ ^
HQ = I ⋅V ⋅I (44)
2I (2I − 1)ℏ

We can actually re-write this expression to account for the description of the EFG that was given in the Cartesian components
if we change spin operators to their Cartesian analogues. This gives the result
eQ 3 2
^ ^ ^ ^ ^ ^
HQ = ∑ Vαβ [ (I α I β + I β I α ) − δαβ I ] (45)
6I (2I − 1)ℏ 2
α,β=x,y,z

The EFG is traceless and can be described using an asymmetry parameter define as
P AS P AS
Vxx − Vyy
ηQ = (46)
P AS
Vzz

and the magnitude of the EFG will then be given as


P AS
eq = Vzz (47)

We must at this point recognize that we are in the reference frame of the electric field gradient. The HQ in the reference frame
of the EFG is then

7/4/2021 5 https://chem.libretexts.org/@go/page/1827
2
e Qq 2P AS
^2 2P AS 2P AS
^ ^ ^ ^
HQ = [3 I zz − I + η(I xx − I yy )] (48)
4I (2I − 1)ℏ

For the remainder of the discussions we can simplify the constant term to χ,
2
e qQ
χ = (49)

This is known as the quadrupole coupling constant and is the accpeted term in the NMR literature. Readers must be wary
however as there are several different definitions floating in the literature. The expansion of the Hamiltonian is done using
perturbation theory. It has been shown that experimental spectra may be exactly calculated using the first and second
perturbations to the Hamiltonian. Using this information, the Hamiltonian may be expanded in terms of polar coordinates and
raising and lowering operators. This is done to transform the Hamiltonian from the PAS of the EFG into the laboratory frame
χ 1 2 2
^ 2 ^ ^
HQ = (3 cos θ − 1)(3 I z − I ) (50)
4I (2I − 1) 2

3
^ ^ ^ ^ ^ ^
+ sin θ cos θ [I z (I + + I − ) + (I + + I − )I z ] (51)
2

3 2 2
2 ^ ^
+ sin θ(I + + I − ) (52)
4

2πχ 1 2 2
2 ^ ^
+ηQ [ cos 2ϕ[(1 − cos θ)(3 I z − I ) (53)
4I (2I − 1)h 2

2 2
2 ^ ^
+(cos θ + 1)(I + + I − )] (54)

1
^ ^ ^ ^
+ sin θ[cos θ cos 2ϕ − i sin 2ϕ)(I + I z + I z I + ) (55)
2

^ ^ ^ ^
+(cos θ cos 2ϕ + sin 2ϕ)(I − I z + I z I − )] (56)

2 2
^ ^
+(i/4) sin 2ϕ cos θ(I + − I − ) (57)

.
From which we can get first and second order corrections to the energy levels,

(1)
χ 1
2 2 2
Em = (I (I + 1) − 3 m ) [ (3 cos θ − 1) − η cos 2ϕ(cos θ − 1)] (58)
4I (2I − 1) 2

2
(2) χ m 1
2 2
Em =− − (I (I + 1) − 3 m )(3 + η ) (59)
Q
4I (2I − 1)ω0 5

1 2 2 2
2
+ (8I (I + 1 − 12 m − 3)[(η − 3)(3 cos θ − 1) + 6 ηQ sin θ cos 2ϕ (60)
Q
28

1 2
1 2 4 2
+ (18I (I + 1) − 34 m − 5)[ (18 + η )(35 cos θ − 30 cos θ + 3) (61)
8 140

3 2
1 4
2
+ ηQ sin θ(7 cos θ − 1) cos 2ϕ + η sin θ cos 4ϕ] (62)
Q
7 4

.
An illustrative figure showing how the energy levels change according to the Zeeman, the first, and the second order
quadrupolar interactions is shown below for a spin 3/2 nucleus. Interestingly, he first order approximation does not affect the
central transition, while the second order transition is inversely proportional the Larmor frequency. With increasing field, the
second order effect on the central transition decreases.

7/4/2021 6 https://chem.libretexts.org/@go/page/1827
Figure 6. Splitting of the energy levels of a quadrupolar nucleus. The diagram progresses from the no magnetic field to the
Zeeman splitting to the first order quadrupole perturbation, and finally the second order quadrupole perturbation. The energy
level values correspond to above equations.

Quadrupolar Lineshapes
The signal exhibited by the quadrupolar nucleus exhibits a very characteristic powder pattern. Looking at the energy levels, it
is easy to see there are multiple transitions which can occur. For a spin 1 nucleus, the transitions are from -1 to 0 and 0 to 1.
These two transitions manifest themselves as a double horned powder pattern, each horn representative of a transition. The
difference in intensities is due to the alignment of the crystallites with respect to the magnetic field. If the crystallite is aligned
with the B0 field, then after application of a pulse, it will lie in the x-y plane and consequently contribute fully to the signal.
π

If the crystallite is oriented with some angle relative to the B0 axis and a pulse is applied, it will not process as long and
consequently will not be detected for as long and give less of a signal.
The frequency at which the transitions will occur is given by the quadrupolar frequency defined as
3 2m − 1
2
ωQ (θ) = ω0 − ( ) χ(3 cos θ − 1) (63)
8 I (2I − 1)

from this it is easy to see why two horns are observed. m is either 1 (red) or -1 (blue) which changes the sign of the
quadrupolar perturbation to the Larmor frequency.

7/4/2021 7 https://chem.libretexts.org/@go/page/1827
Figure 7. The double horned quadrupolar splitting pattern of deuterium. The blue line is the 1 to 0 transition and the red line
corresponds to the 0 to 1 transition. The decrease in intensity is due to the orientation of the nuclei after the pulse is applied.
For nuclei, such as 87Rb, which have multiple transitions, the powder patterns are more complex. Similar to the CSA the
lineshape is dependent on the magnitude of χ and η , as shown in the figure below.

Figure 8. The effect of Cq and η on the shape of the static lineshape for 87Rb quadrupolar nucleus.
Typically, the satellite transitions are not observed in quadrupolar spectra. The frequency for a symmetric transition, such as
-1/2 to 1/2 in quadrupolar nuclei may be represented by the sum of the 0th 2nd and 4th rank Legendre polynomials or
mathematically as

νm,−m = ∑ (α, β, γ)Cl (I , m)Pl cos θMA (64)

l=0,2,4

where
2
C0 (I , m) = 2m[I (I + 1) − 3 m ] (65)

2
C2 (I , m) = 2m[8I (I + 1) − 12 m − 3] (66)

2
C4 (I , m) = 2m[18I (I + 1) − 34 m − 5] (67)

7/4/2021 8 https://chem.libretexts.org/@go/page/1827
1 2
P2 (cos θ) = (3 cos θ − 1) (68)
2

1 4 2
P4 (cos θ) = (35 cos θ − 30 cos θ + 3) (69)
8

The second and fourth rank interactions also have associated frequencies. These are given below

Q 1 χ 2
3
v = [ ] [I (I + 1) − ] F2 (θ, ϕ) (70)
2
192νL I (2I − 1)h 4

Q 1 χ 3
2
v = [ ] [I (I + 1) − ] F4 (θ, ϕ) (71)
4
3360νL I (2I − 1)h 4

35
2 4 2 2 2
F2 (θ, ϕ) = (3 − ηQ cos 2ϕ) sin θ − 5(18 + η − 9 ηQ cos 2ϕ) sin θ + 18 + η (72)
Q Q
4

2 1
2 2 2
F4 (θ, ϕ = 2(3 ηQ cos 2ϕ − η − 3) sin θ+ (22 η − 90 ηQ cos 2ϕ + 120 (73)
Q Q
3 21

.
These equations are more useful mathematically and become important in the discussion of pulse sequences of quadrupolar
nuclei.

Magnetic Field Effects and the Center of Gravity


The nature of the quadrupolar interaction is heavily influenced by the magnetic field. Below is a figure that shows the field
dependence of a Ga resonance in β-Ga2O3. The reader should take note of two things. First, Note how the spectrum narrows as
the field is increased. This shows the effect of the central transition is inversely proportional to ω . Second, note the shift in the 0

center of gravity of the peak. As the field increases, the peak shifts progressivley upfield, although the isotropic peak position,
denoted by the dotted line is constant.

Quadrupole Moments: Effects in NMR Spectra


As mentioned earlier, a quadrupolar nucleus is efficiently relaxed by a non-uniform electric field that is a product of the solute
molecules interaction with the dipolar solvent. This relaxation is dependent on the interaction of the electric field gradient at
the nucleus. When the nucleus is in a molecule that is surrounded by a non-spherical electron density distribution, it creates a
gradient. The field gradient, q, describes the electron charge cloud’s deviation from spherical symmetry. The value of q is
found to equal zero if the groups around the quadrupolar nucleus have a cubic symmetry, such as in the Td point group.
However, if a non-cublic molecule has a threefold or higher symmetry axis, the deviation from spherical symmetry is
expressed as a magnitude of q. The two parameters, q, the field gradient, and η, the asymmetric parameter, become necessary
only if the molecule's point group's highest symmetry axis is a threefold symmetry or less. Depending on the molecule, certain
cancellations can take place leading the asymmetric parameter, η, to equal zero. This is caused by a combination of very
specific bond angles and charge distribution in the molecule being analyzed. Ultimately, the effectiveness of the relaxation is
dependent on the magnitude of the electric field gradient, q.

7/4/2021 9 https://chem.libretexts.org/@go/page/1827
Linewidth broadening in the NMR spectrum is consequential of the rapid nuclear quadrupole relaxation of the quadrupole
nucleus. Consider an analogous situation: chemical exchange. It is known that when the nuclei’s spin state rapidly changes it
causes broadening in the spectrum. Similarly, the nuclear quadrupole relaxation rates of a quadrupolar nucelus corresponds to
an intermediate rate of chemical relaxation.The apparent broadening effect also influences the spectra of the other nuclei
attached to the quadrupolar nucleus, including protons. In some cases, the rapid nucleur quadrupole relaxation times (T1) can
cause extensive homogenous broadening (consequential of readily relaxing nuclei, seen in Figure 2) rendering the proton
signal of the quadrupolar nucleus completely unobservable in the 1H NMR spectrum. T1 is determined by two factors: the
electric quadrupole moment (Q) and the presence of the electric field gradient (q) across the nucleus.
A common approach to resolving quadrupolar effects on the spectra of solution state NMR is elevating temperatures while
collecting NMR data. The molecular reorientational correlation times are then shorter than the normal time scale, so the
homogenous broadening of the line can be reduced. Unfortunately, the temperature required to create this motional tapering is
unfeasibly high for many samples that would deem this technique necessary.

References
1. Smith, J.A.S., Nuclear Quadrupole Resonance Spectroscopy: General Principles. J Chem. Ed. 1971, 48, (1), 39.
2. Drago, Russell S. "Quadrupole Moments" Physical Methods for Chemists. Ft. Worth: Saunders College Pub., 1992.
3. Pyykkö, P., Spectroscopic Nuclear Quadrupole Moments. Mol. Phys. 2001, 99, (19), 1617.
4. Gerothanassis, I.P., Kalodimos, C.G., NMR Shielding and the Periodic Table. J. Chem. Ed. 1996, 73, 801. DOI:
10.1021/ed073p801
5. Laidler, K.J., Meiser, J.H., Sanctuary, B.C. "Magnetic Interaction Leading to Spectra Consequences" Physical Chemistry.
New York: Houghton Mifflin Company., 2003.
6. Harris, R.K., Mann, B.E., NMR and the Periodic Table, Acadademic Press, New York, 1978.
7. ie.lbl.gov/toipdf/mometbl.pdf

Contributors and Attributions


Derrick C. Kaseman (UC Davis) and Megan McKenney (UCDavis)

7/4/2021 10 https://chem.libretexts.org/@go/page/1827
Quantum Mechanic Treatment
Introduction
NMR employs strong magnetic field which interact with the intrinsic spin of the nucleus to form degenerate energy levels. Treating
this phenomenon classically is mathematically tedious and/or impossible. Here we outline the fundamentals of NMR in quantum
mechanical terms.

Zeeman Interaction
If we begin with the historical Schrödinger equation

H |ψ⟩ = E|ψ⟩ (1)

we can immediately realize that by defining the Hamiltonian of the system will result with an corresponding energy level of the
wave function.

The Hamiltonian for a magnetic vector (μ⃗  in J/T)in a magnetic field (also a vector), (B⃗  in T) is given by the expression
⃗ 
H = −μ⃗ ⋅ B (2)

We then see that the Hamiltonian has units of energy (J). Therefore operating the Hamiltonian on a wavefinction will result in an
energy eigenvalue! We can further refine the Hamiltonian by relating the magnetic moment of a nucleus to the particles angular
momentum ari\sing from the intrinsic spin, I, of the nucleus thus
⃗ 
μ⃗ = −γ I (3)

where gamma is the ratio of the angular momentum to the magnetic moment. Then the Hamiltonian may be rewritten as
⃗  ⃗ 
H = −γ I ⋅ B (4)

Assuming that the magnetic field lies along the Z-axis, then
^ ^ ^ ^
Hz = −γ(Ix i + Iy j + Iz k) ⋅ (B0 k) = −γ B0 Iz = −ω0 Iz (5)

where ω is the Larmor frequency. Therefore if the |ψ⟩ is an eigenfunction of H then the |±⟩ spin states are eigenfunctions of the
0

total and z-component of angular momentum. We can describe this is three ways in a DC magnetic field:
Operator:
ω0
Hz |+⟩ = −ω0 Iz |+⟩ = − |+⟩ (6)
2

ω0
Hz |−⟩ = −ω0 Iz |−⟩ = − |−⟩ (7)
2

Matrix:
Hz = (8)

Energy Level
Insert picture
These representations rely on the fact that spin states are eigenfunctions of the angular momentum and therefore have the following
properties, with I and m defining the state of the system.
2
I |I , m⟩ = I (I + 1)|I , m⟩ (9)

Iz |I , m⟩ = m|I , m⟩ (10)

−−−−−−−−−−−−−−−−−− −
1
I+1 |I , m⟩ = −√ (I (I + 1) − m(m + 1)) |I , m+1 ⟩ (11)
2

−−−−−−−−−−−−−−−−−− −
1
I−1 |I , m⟩ = √ (I (I + 1) − m(m − 1)) |I , m−1 ⟩ (12)
2

5/28/2021 1 https://chem.libretexts.org/@go/page/24649
where
1
Ix = – (I+1 − I−1 ) (13)
√2

i
Iy = (I+1 + I−1 ) (14)

√2

Note that in contrast to the harmonic oscillator, the |I , m⟩ spin basis is orthogonal and complete. This means for the spin I , there are
only 2I + 1 spin states and the the overlap integral ⟨I , m|I , m ⟩ = δ
′ ′
δmm holds. This formalism holds equally well for I > 1/2

II

particles as well.

Effects of Radio Frequency


While it is relatively straightforward to understand the Zeeman interaction, understanding the effects of radio frequency pulses on
the Hamiltonian needs careful consideration. To examine this we will look at a modification to the Stern-Gerlach experiment made
by Rabi in the 1930's.
\since no one has bothered to discuss the Stern Gerlach experiment in any detail on the chemwiki a brief glimpse into the initial
experiment is necessary. Essentially a beam of silver atoms was shot through a magnetic field and instead of a \single line of atoms
or \single spot, depending on the dispersion of the atoms, two spots were observed. This of course was due to the the fact that
Zeeman splitting had occurred due to silvers spin=1/2 nature. A figure is provided below for reference.

Rabi modified this experiment by adding a coil of wire which generated a magnetic field perpendicular to the applied magnetic field
from the stern gerlach experiment. By selecting only 1 of the nuclei either +/- 1/2 and running these particles through the radio
frequency coil Rabi noted a characteristic nutation of the magnetization shown below.
insert figure
We can see then that the number of particles appearing on the screen is characterisitc of the frequency. This behavior is denoted as
the Rabi or nutation frequency. This experiment serves as a good bridge into the NMR experiment. First, we must calculate the
appropriate Hamiltonian for the sitation. We have two magnetic fields, one is the large external magnetic field which splits the
degenerate ground states into energy levels, and the second is the osciallting magnetic field from the coil. As such the magnetic field
experienced by the spins is
⃗  ^ ^
B = B0 k + 2 B1 cos(ω0 t) i (15)

Here we define the coil to be along the x direction and the external magnetic field about the z direction. \since the field is oscillating
in time, a \co\sine function is the appropriate choice to model the behavior of this magnetic field.
We may now express the Hamiltonian in eith operator or matrix form as
⃗  ⃗ 
H = −γ I ⋅ B = −γ B0 Iz − 2γ B1 cos(ωo t)Ix = −ω0 Iz − 2 ω1 cos(ω0 t)Ix (16)

insert matrix here.


We now see that our Hamiltonian become time dependent. In order to solve the observation made by Rabi we must use the time
depedent Schrodinger equation.
d
|ψ(t)⟩ = −iH (t)|ψ(t)⟩ (17)
dt

5/28/2021 2 https://chem.libretexts.org/@go/page/24649
where H(t) is in frequency units and |ψ (t=0)>=|+>. Assuming that the Hamiltonian is time independent we can use a time dependent
unity operator U^ defined as
^
|ψ(t)⟩ = U (t)|ψ(0) > (18)

^ −iH t
U (t) = e (19)

that transforms the initial wavefunction |ψ (0)>to a wavefunction at a given time |ψ (t) \rangle.
insert figure
This issue with the case above is that we assumed the Hamiltonian is time-independent, which is clearly not the our case. Therefore,
we need to solve this problem by transforming the time-dependent Schrodinger equation into a reference frame that renders the
Hamiltonian time independent.
Consider the development of the time evolution of |Q(t) \rangle under the following conditions
|ψ(t)⟩ = R(t)|Q(t)⟩ (20)

+
Q(t)⟩ = R (t)|ψ(t)⟩ (21)

+ +
R(t)R (t) = R (t)R(t) = 1 (22)

As you can see R commutes and the adjoint give the unity operator. We can then rewrite the Schrodinger equation as
d d d d
|ψ(t)⟩ = R(t)|Q(t)⟩ = [ R(t)]|Q(t)⟩ + R(t) |Q(t)⟩ (23)
dt dt dt Dt

= −iH (t)|ψ(t)⟩ = −iH (t)R(t)|Q(t)⟩ (24)

Mow if we multiple both sides of the equation by the adjoint then we can rearrange to obtain a new form of the Schrodinger
equations
d d
+ +
[R (t) R(t)]|Q(t)⟩ + |Q(t)⟩ = −i R (t)H (t)R(t)|Q(t)⟩ (25)
dt dt

d
|Q(t)⟩ = −i Hint (t)|Q(t)⟩ (26)
dt

where
d
+ +
Hint = R (t)H (t)R(t) − i R (t) R(t) (27)
dt

To apply this to our situation consider the case where


−iHz t iω0 Iz t
R(t) = e =e (28)

+ −iω0 Iz t
R =e (29)

Then we can develop the following Identities


i ω0 t
⎡ ⎤
2
⎢e 0 ⎥
R(t) = ⎢ ⎥ (30)
⎢ −i ω0 t ⎥

⎣ ⎦
0 e 2

−i ω0 t
⎡ ⎤
2
† ⎢e 0 ⎥
R (t) = ⎢ ⎥ (31)
⎢ i ω0 t ⎥

⎣ ⎦
0 e 2

in which we can easily take derivitives of leading to

5/28/2021 3 https://chem.libretexts.org/@go/page/24649
i ω0 t
⎡ ⎤
d
⎢ e 2 0 ⎥
d ⎢ dt ⎥
R(t) = ⎢ ⎥ (32)
dt ⎢ −i ω0 t ⎥
⎢ ⎥
d
⎣ 0 e 2 ⎦
dt

i ω0 t
⎡ ⎤
iω0
⎢ e 2 0 ⎥
d ⎢ 2 ⎥
R(t) = ⎢ ⎥ (33)
dt ⎢ −i ω0 t ⎥
⎢ ⎥
−iω0
⎣ 0 e 2 ⎦
2

i ω0 t
1 ⎡ ⎤
⎡ 0 ⎤
d 2
⎢e 0

R(t) = i ω0 ⎢ 2 ⎥⋅⎢ ⎥ (34)
⎢ ⎥ ⎢
dt 1 −i ω0 t ⎥
⎣ 0 ⎦
⎣ ⎦
2 0 e 2

d iω0 tIz
R(t) = i ω0 Iz ⋅ e (35)
dt

We can also calculate


−i ω0 t i ω0 t
⎡ ⎤ 1 ⎡ ⎤
⎡ 0 ⎤
2 2
† ⎢e 0 ⎥ 2 ⎢e 0 ⎥
R (t)Ix R(t) = ⎢ ⎥⋅⎢

⎥⋅⎢
⎥ ⎥ (36)
⎢ i ω0 t ⎥ 1 ⎢ −i ω0 t ⎥
⎣ 0 ⎦
⎣ ⎦ ⎣ ⎦
0 e 2 2 0 e 2

−i ω0 t
⎡ ⎤
1
0 e 2
⎢ ⎥
† ⎢ 2 ⎥
R (t)Ix R(t) = ⎢ ⎥ (37)
⎢ i ω0 t ⎥
⎢ ⎥
1
⎣ e 2 0 ⎦
2

1 −i
⎡ 0 ⎤ ⎡ 0 ⎤
† 2 2
R (t)Ix R(t) = cos ωt ⎢ ⎥ + sin ωt ⎢ ⎥ = cos ω tI + sin ω tI
0 x 0 y (38)
⎢ ⎥ ⎢ ⎥
1 i
⎣ 0 ⎦ ⎣ 0 ⎦
2 2

Similarly we can show (left as an exercise for the reader)



R (t)Ix R(t) = cos ω0 tIy − sin ω0 tIx (39)

We now have all the pieces we need to solve the Hamiltonian. We will show this in multiple ways

Matrix Approach
i ω0 t
−i ω0 t ⎡ ⎤
⎡ ⎤ iω0 iω0
e 2 0 ⎡ 0 ⎤
2 ⎢ ⎥
d ⎢e 0 ⎥ ⎢ 2 ⎥ 2
† ⎢ ⎥
R (t) R(t) = ⎢ ⎥⋅⎢ ⎥ =⎢ (40)

dt ⎢ i ω0 t ⎥ ⎢ −i ω0 t ⎥ −iω0
⎢ ⎥ 0
−iω0 ⎣ ⎦
⎣ ⎦
0 e 2 ⎣ 0 e 2 ⎦ 2
2

−i ω0 t i ω0 t
⎡ ⎤ −ω0 ⎡ ⎤
⎡ −ω1 cos ω0 t ⎤
2 2
† ⎢e 0⎥ 2 ⎢e 0 ⎥
R (t)H (t)R(t) = ⎢ ⎥⋅⎢

⎥⋅⎢
⎥ ⎥ (41)
⎢ i ω0 t ⎥ ω0 ⎢ −i ω0 t ⎥
⎣ −ω1 cos ω0 t ⎦
⎣ ⎦ 2 ⎣ ⎦
0 e 2 0 e 2

−ω0
−iω0 t
⎡ −ω1 cos ω0 te ⎤
2
=⎢



ω0
iω0 t
⎣ −ω1 cos ω0 te ⎦
2

Then

5/28/2021 4 https://chem.libretexts.org/@go/page/24649
0 1/2 0 1/2 0 −i/2
Hint = −ω1 [ ] − ω1 cos 2 ω0 t [ ] − ω1 sin 2 ω0 t [ ] (42)
1/2 0 1/2 0 i/2 0

Hint = ω1 Ix − ω1 cos 2 ω0 tIx − ω1 sin 2 ω0 tIy (43)

Operator Approach
We now wish to solve the Hamiltonian u\sing the I operators we have defined previously.
d
† −iω0 Iz iω0 Iz
R (d) R(t) = e ⋅ i ω0 Iz ⋅ e = i ω0 Iz (44)
dt

† −iω0 Iz iω0 Iz
R (t)H (t)R(t) = e (−ω0 Iz − 2 ω1 Ix )e (45)

iω0 tIz −iωo tIz


= −ω0 Iz − 2 ω1 cos ω0 tIx E Ix e (46)

2
= −ω0 Iz − 2 ω1 cos ω0 tIx − 2 ω1 cos ω0 t sin ω0 tIy (47)

= −ω0 Iz − ω1 Ix − ω1 cos 2 ω0 tIx − ω1 sin 2 ω0 tIy (48)

Hint (t) = −ω1 Ix − ω1 cos 2 ω0 tIx − ω1 sin 2 ω0 tIy (49)

This is the exact expression we obtained u\sing the matrix approach!

Rotating Wave Approximation


2
ω
It can be shown that the time dependent part of the Hamiltonian modifies the first term in the above equation by (Need to prove 1

ω0

this). Standard NMR spectrometers typically output ω on the order of kHz, whereas the larmor frequency is on the order of MHz
1

and we can neglect the time department part of the Hamiltonian. This is known as the Rotating Wave Approximation. The
Hamiltonian becomes
1
0
2
Hint = −ω1 Ix = −ω1 [ ] (50)
1
0
2

We now need to find the describe the initial state of the system
1 0 0 0

|Q(0) >= R (t = 0) >= [ ]( ) =( ) (51)
0 1 1 1

Now we if we diagonalize the matrix then we can find the state of the system at any time t. The diagonalized Hamiltonian is then
ω1 t ω1 t
cos i sin
2 2
U (t) = [ ] (52)
ω1 t ω1 t
i sin cos
2 2

We can now evaluate |Q(t) \rangle as \Q(t) \rangle=U(t)|Q(0)


ω1 t ω1 t ω1 t
cos i sin 0 i sin
2 2 2
|Q(t)⟩ = [ ]( ) =[ ] (53)
ω1 t ω1 t ω1 t
i sin cos 1 cos
2 2 2

Finally recognizing that |Ψ(t)⟩=R(t)|Q(t). There for


iω t
iω t 0
0 ω1 t ω1 t
⎡e ⎤ i sin / ⎡ i sin e 2 ⎤
2
0 2 2
|Ψ(t)⟩ = [ ] =⎢ ⎥ (54)
−iω t −iω t
0 ω1 t 0
⎣ ⎦ cos ⎣ ω1 t

0 e 2
2 cos e 2

iω t −iω t
ω1 t 0
ω1 t 0

= i sin e 2
|−⟩ + cos e 2
|+ > (55)
2 2

= c− (t)|−⟩ + c+ (t)|+ > (56)

The probability that Rabi obsereved is given by multiplying the complex congugates which are calculated as
1 1
P|−⟩ = − cos ω1 t (57)
2 2

1 1
P|+> = + cos ω1 t (58)
2 2

5/28/2021 5 https://chem.libretexts.org/@go/page/24649
The Hamiltonian we found to govern Rabi's experiment is the exact Hamiltonian used in the NMR experiment, with two slight
modifications. First, Rabi was able to select an initial state, while in NMR we have a thermal equilibrium population. Secondly we
measure a transverse magnetization instead of |+> and |-\rangle.

Liouville Von Neumann Equation


\since we are dealing with a system that can be described by both a statistical and quantum mechanical aspects, we need a way to
unify these descriptions. The solution is the Liouville-von Neumann equation, which describes how the density operator evolves in
time. More specifically, it describes the time evolution of a mixed state, which is directly applicable to NMR, as we want to know
how the off-diagonal elements of the density matrix evolve after an applied RF. The density matrix is evolving in time, thus we must
derive a differential equation to describe this behavior then integrate it with each step of the NMR experiment. As usual we begin
with the time-dependent Schroedinger Equation.
d
¯
ih Ψ(t) = H (t)Ψ(t) (59)
dt

Lets also look at the thermal equilibrium of the spins. U\sing Boltzmann statistics the initial condition of our spins is described by
E
+

Ne kT

c+ (0)c+ (0) = (60)
Z

E


Ne kT

c− (0)c− (0) = (61)
Z

where
E E
− +
− −
Z = [e kT
+e kT
(62)

and

c+ (0)c (0) = 0 (63)


c− (0)c (0) = 0 (64)
+

Defining
|Ψ(t)⟩ = c+ (t)\+ > +c− (t)|−⟩ (65)

\[\langle \Psi(t)|=c^*_+(t)\langle+| +c^*_-(t)<-|\


we immediately realize that we have a density operator!
∗ ∗ ∗ ∗
ρ(t) = |Ψ(t)⟩ < Ψ(t)| = c+ (t)c+ (t)\+ > ⟨+| + c− (t)c− (t)|−⟩ < −| + c+ (t)c− (t)|+ >< −| + c− (t)c+ (t)|−⟩⟨+| (66)

Then the time evolution is


d d d
ρ(t) = |Ψ(t)⟩ < Ψ(t)| + |Ψ(t)⟩ < Ψ(t)| = −iH (t)|Ψ(t)⟩ < Ψ(t)| + i|Ψ(t)⟩ < Ψ(t)|H (t) (67)
dt dt dt

which gives the Liouville-von Neumann Equation as


d
ρ(t) = i[H (t), ρ(t)] (68)
dt

Lets now look at how this evolves in a simples NMR experiment.


Initially we can define the density matrix at thermal equilibrium as
E+
∗ ∗ −
c+ (0)c+ (0) c+ (0)c− (0) 1 ⎡e kT 0 ⎤
ρ(0) = [ ] = (69)
∗ ∗ E−
c− (0)c+ (0) c− (0)c− (0) Z ⎣ − ⎦
0 e kT

High Temperature Approximation


We can further refine the inital density operator by substituting in the equations for the energy levels E+ and E-.

h̄ω0
E( ±) = ∓ (70)
2

5/28/2021 6 https://chem.libretexts.org/@go/page/24649
but \since we know the kT>>ω then we can use a high temperature approximation of
0

±

0 h̄ω0
e 2kT =1± (71)
2kT

ans
Z =2 (72)

resulting in a new density operator as


¯
1 hω0
⎡ + 0 ⎤ ¯
2 2kT
N N hω0
ρ(0) ⎢ ⎥ = 1+ Iz (73)
¯
⎣ 1 hω0
⎦ 2 2kT
0 −
2 2kT

Normally the magnetization constant is taken to be a scaling factor and the unity operator is invariant to applied fields which firther
reduces the density matrix to
ρ(0) = Iz (74)

NMR Observables
In a simple NMR experiment, the magnetization along the externally applied magnetix field is knocked into the x-y plane u\sing a rf
pulse. As the spins recover along the external field direction (z direction) they precess near their Larmor frequencies through a coil
which in turn produces an EMF. In order for an EMF to be generated the preces\sing magnetization must be perpendicular to the
coil. According the Kirchoff
d
EM F = −μ0 ηA Mx (75)
dt

as is immediatley evident the magnetizaion is time dependent and our problem is essentially reduced to
Mx (t) =< Tx >=< Ψ(t)| Ix |Ψ(t)⟩ (76)

which is the expectation value of Ix. If instead we consider a generic operator <O> then

|Ψ(t)⟩ = ∑ cn (t)|n > (77)


< Ψ(t)| = ∑ cn (t) < n| (78)


< O >=< Ψ(t)|O|Ψ(t)⟩ = ∑ cn (t)cm (t) < m|O|n > (79)

n,m

and recalling that



< n|ρ(t)|m >= cn (t)cm (t) (80)

results in

< O > ∑ < n|ρ(t)|m >< m|O|n >= T r[ρ(t)O] (81)

n,m

Now for Mx
Mx =< Ix >= T R[ Ix ρ(t) (82)

ρ(t) = Iz (83)

H (t) = −ω0 Iz − 2 ω1 cos ω0 tIx (84)

Then all we need to do is calculate ρ(t) and <Ix>!

Matrix Approach
Like we did in the above treatment for the roatating wave, we need to move ρ(t) into a different representation.

σ(t) = R (t)ρ(t)R(t) (85)

5/28/2021 7 https://chem.libretexts.org/@go/page/24649

ρ(t) = R(t)σ(t)R (t) (86)

d
σ(t) = −iH [ Hint , σ(t)] (87)
dt

and remembering that


iω0 tIz
R(t) = e (88)

σ(0) = ρ(0) (89)

We get
ω1 t ω1 t
cos i sin
2 2
U (t) = [ ] (90)
ω1 t ω1 t
i sin cos
2 2

ω1 t ω1 t
cos −i sin
† 2 2
U (t) = [ ] (91)
ω1 t ω1 t
−i sin cos
2 2

which gives

σ(t) = U (t)σ(0)U (t) (92)

ω1 t ω1 t 1 ω1 t ω1 t
cos i sin 0 cos −i sin 1 cos ω1 t −i sin ω1 t
2 2 2 2 2
=[ ][ ][ ] = [ ] (93)
ω1 t ω1 t 1 ω1 t ω1 t
i sin cos 0 −i sin cos 2 −i sin ω1 t cos ω1 t
2 2 2 2 2

Now we can use σ(t) to find ρ(t).


iω t −iω t
0 0


1⎡e 2 0 ⎤ cos ω1 t −i sin ω1 t ⎡e 2 0 ⎤
ρ(t) = R(t)σ(t)R (t) = −iω t
[ ] iω t
(94)
2⎣ 0
⎦ −i sin ω1 t cos ω1 t ⎣
0

0 e 2
0 e 2

iω0 t
cos ω1 t −i sin ω1 te
=[ ]
−iω0 t
−i sin ω1 te cos ω0 t

Now lets look at the observables Ix and Iy.


iω0 t
0 1/2 cos ω1 t −i sin ω1 te
< Ix >= T rIx ρ(t) = T r[ ][ ] (95)
−iω0 t
1/2 0 −i sin ω1 te cos ω1 t

sinω1 tsinω0 t
= (96)
2

Then <Iy> is
sinω1 tcosω0 t
< Iy >= (97)
2

And <Iz> is
cosω1 t
< Iz >= (98)
2

Operator Approach
Alternatively we can use the operator approach, which is much less mathemtically intensive.
σ(0) = ρ(0) = Iz (99)

Hint = −ω1 Ix (100)

−iHint t iHint t
σ(t) = e σ(0)e = cosω1 tIz + sinω1 tIy (101)


ρ(t) = R(t)σ(t)R (t) = cosω1 tIz + sinω1 tcosω0 tIy + sinω1 tsinω0 tIx (102)

Using

5/28/2021 8 https://chem.libretexts.org/@go/page/24649
1
T r[ In Im ] = δnm (103)
2

we obtain the same expectation values as before.

5/28/2021 9 https://chem.libretexts.org/@go/page/24649
Relaxation
Relaxation in NMR is a fundamental concept which describes the coherence loss of the magnetization in the x-y plane and the
recovery of relaxation along the z-axis. One can easily imagine that in the absence of any other effects, the magnetization in
the x-y plane will recover along the z-axis as the external magnetic field forces the spins to align with it. This is known as T1
relaxation. The loss of coherence of the magnetization in the x-y plane is due to spins "forgetting" their orientation with
respect to the bulk magnetization. This is known as T2 relaxation. There are many factors that contribute to the relaxation
processes. This page will examine the different types of relaxation at a basic level. For those interested in a deeper discussion
please see article on T1 and T2 relaxation.

T1 Relaxation
T1 relaxation, also known as spin lattice or longitudinal relaxation is the time constant used to describe when ~63% of the
magnetization has recovered to equilibrium. The T1 of a given spin is dictated by field fluctuations (both magnetic and
electric) that occur in the sample. Consequently, T1 measurements can tell us important information regarding inter and intra
molecular dynamics of the system. Several factor may cause this alternating field: molecular motion, J-Coupling, Dipolar
Coupling, Chemical Shift Anisotropy, and quadrupole-phonon interactions.
We can also look at relaxation from the energy level standpoint. Initially, we have a Boltzmann distribution of spins in the (For
I=1/2) in two energy levels, with the lower energy level slightly more populated than the higher energy level. After a pulse
these spin populations are inverted and now the higher energy level has more spins in it. Eventually, these spins will go back to
their lower energy state due to relaxation. The timescale on which this occurs is T1.

Figure 1: Energy level diagram of T1 relaxation. Initially, the spins are at equilibrium with the red spins in the +1/2 energy
level and the purple spins in the -1/2 energy level. Note that the energy level distribution is grossly exaggerated. After a 90o
pulse (center figure), the energy level spin populations switch, where red spins are now in the -1/2 state. After some time T1,
~60% of the spins have relaxed back to their equilibrium energy levels. After 5T1 (left figure) all of the spins have relaxed
back to their equilibrium levels.
From the Bloch Equations, we know that magnetization is along the Z-axis is
−t
Mz = M0 (1 − exp ) (1)
T1

Figure 2: magnetization is along the Z-axis

T1 Relaxation
ρ

T1 is the relaxation of the magnetization during a spin lock. This becomes particularly important in the cross-polarization.
ρ

Derrick Kaseman 6/21/2021 1 https://chem.libretexts.org/@go/page/1829


T2 Relaxation
T2 relaxation is also known as spin-spin or transverse relaxation. T2 relaxation involves energy transfer between interacting
spins via dipole and exchange interactions. Spin-spin relaxation energy is transferred to a neighboring nucleus. The time
constant for this process is called the spin-spin relaxation time (T2). The relaxation rate is proportional to the concentration of
paramagnetic ions in the sample. This mechanism is largely temperature independent. T2 values are generally much less
dependent on field strength, B, than T1 values.
In the process of relaxation, the component of magnetization in xy-plane will become zero as M0 returns to z-axis. Time
constant T2 is used here to describe spin-spin relaxation with the following function. This process involves energy changing
between the spin-active and their adjacent nuclei.
−t
Mxy = Mxy0 exp (2)
T2

T2* Relaxation
In an ideal NMR spectrometer, the external magnetic field is completely homogeneous. However, all magnets have small
inhomegneities in them. Consequently, nuclei experience different magnetic fields which changes the precessional frequency
of each nucleus. The change in Larmor frequency results in dphasing in the transverse plane and is known as T2*

Figure 3: Effect of magnetic field inhomogeneity on the observed T2. When there is no inhomogeneity, Mxy decays at a slower
rate T2 (blue). Usually there is inhomogeneity in the field, which causes the decay of Mxy to be faster with the apparent time
constant T2* (red).

Magnetic Field Dependence


So far, we have only described what the relaxation parameters are, but have largely omitted the equations that give us the
results. All elaxation measurements are based on the idea of a correlation time, tc. The correlation time is based on the
environment that the spin is in. For T relaxation, the correlation time must be on the order of the larmor frequency of the
1

nucleus, while the correlation time for the for T must be on the order of the inverse of the linewidth to make effects. The
2

correlation time are random fluctuations in the magnetic field that couple to the nucleus. The equations that govern T and T 1 2

are given below with no additional proof..


1 tc 2tc
=C ( + )
2 2 2 2
T1 1 + w tc 1 + 4w tc

1 2 2
1
= 3C tc ( ) +
T2 π 2

where w is the Larmor frequency, and C is a constant that contains temperature and frequency independent terms and defined
as....

References
1. Levitt, M. H., Spin Dynamics, Wiley
2. Bloembergen, E.M. Purcell, R.V. Pound "Relaxation Effects in Nuclear Magnetic Resonance Absorption" Physical Review
1948, 73, 679-746

Derrick Kaseman 6/21/2021 2 https://chem.libretexts.org/@go/page/1829


3. www.uthgsbsmedphys.org/GS02-0032/documents/RelaxationMechanisms.pdf
4. http://chem.ch.huji.ac.il/nmr/techniques/other/t1t2/t1t2.html

Problems
1. Derive why only 63% of spins have recovered to equilibrium after T1
2. Explain the differences between T1, T1/(/rho/),T2, and T2*.

Contributors and Attributions


Derrick Kaseman

Derrick Kaseman 6/21/2021 3 https://chem.libretexts.org/@go/page/1829


NMR: Kinetics
Nuclear magnetic resonance (NMR) is an analytical technique used in chemistry to help identify chemical compounds, obtain
information on the geometry and orientation of molecules, as well as to study chemical equilibrium of species undergoing
physical changes of composition, among many others. Capitalizing on the ability to manipulate the magnetization through
different pulse programs in NMR, allows for the study and understanding of the kinetics of a system. The exchange rates
between two sites can be evaluated through dynamic nuclear magnetic resonance experiments (DNMR). 17O is a common,
NMR active nucleus that is used in the study of kinetics.

Introduction
NMR uses radio frequency radiation to change the direction of nuclear spins that have been placed in a static magnetic field,
and measures the change of magnetization as a function of time. Since its discovery, NMR has gone through many
advancements that have enabled it to become a very useful analytical technique. The Fourier Transform NMR has enabled
more complicated studies through the ability to create pulse programs that can manipulate the spectra, like saturate one species
magnetization so no peak is produced. These pulse programs can also be used to tip the spin of certain nuclei, while keeping
others along the z-axis. This is useful for many applications, including being able to quench signals, change the direction
(positive or negative) of the signal, and track relaxation, to name a few examples. Using different pulse programs allows for
the study of exchange rates between species. This is done by monitoring the changes in the environment of the NMR active
nuclei as a result exchange between the sites. Because of the exchange, spins (magnetization) will be transferred, leading to
changes in the bulk magnetization at both sites. Any NMR active nuclei can be used to study exchange rates, such as 13C, 1H,
17O, but 17O kinetic studies are often performed. This is done because 17O enriched water can be used as one of the exchange

sites, normally the bulk solvent site.

17O NMR for Kinetics Studies


Background and Equations
Oxygen seventeen nuclei have a spin state of 5/2, making them susceptible to nuclear magnetic resonance. This isotope of
oxygen is only 0.0373% naturally abundant, but using isotopically labeled oxygen compounds can result in useful information.
Studying these nuclei in the presence of a magnetic field will provide information about the structure and environment of the
oxygens in the molecule. Using dynamic NMR or DNMR, 17O NMR experiments can be performed to understand chemical
reactivity and kinetics of compounds. DNMR studies the effect of a chemical exchange between two sites that have either a
different chemical shift or coupling constant. These studies are done by obtaining NMR spectra over time and analyzing the
increase and/or decrease of the signals. Unlike other methods that are used to study kinetics, NMR studies can acquire
information about the effects of the exchange on the molecules.
To utilize NMR spectra to establish kinetic information, the Bloch equations must be adapted to include terms that take into
account relaxation as a result of chemical reactivity. While investigating exchange reaction of 17O water between two sites, the
bulk water and water bound to a metal, it is assumed that the kinetics are 1st order, such that:

duM → ←
=− k M uM + k W uW (NMR.1)
dt

duW ← →
=− k W uW + k M uM (NMR.2)
dt

Where k⃗  & k⃗  represent the rate of exchange between the bulk water and the bound water. These two sites can be said to be
M W

coupled because the isotopically enriched oxygen is exchanging between the metal site and bulk water site. As exchange
occurs, the magnetization of the 17O metal ensemble and 17O water ensemble will change, not only due to magnetization
relaxation, but also due to the exchange. The exchange rate terms can be added into the Bloch equations to take into account
the relaxation. With the addition of this term, the equations are known as the Bloch-McConnell equations. Since there are two

5/28/2021 NMR.1 https://chem.libretexts.org/@go/page/1830


sites and three Bloch equations per site, there is a total of six equations for the change in magnetization of the system.
Equations 1-3 are for the metal site, while Equations 4-6 are for the bulk water site.

Bloch-McConnell Equations for Metal Site (Equations 1-3)


duM uM → ←
= vM (ωrf − ωo ) − − k M uM + k W uW (NMR.3)
dt T2M

dvM vM → ←
= −uM (ωrf − ωo ) − − k M vM + k W vW (NMR.4)
dt T2M

dmzM (mzM − mo ) → ←
= vM ω1 − − k M mzM + k W mzW (NMR.5)
dt T1M

Bloch-McConnell Equations for Bulk Water Site (Equations 4-6)


uW ← →
df racduW dt = vW (ωrf − ωo ) − − k W uW + k M uM (NMR.6)
T2W

vW ← →
df racdvW dt = −uW (ωrf − ωo ) − − k W vW + k M vM (NMR.7)
T2W

(mzM − mo ) ← →
df racdmzW dt = vW ω1 − − k W mzW + k M mzM (NMR.8)
T1W

To analyze the NMR spectra, which is obtained by measuring the magnetization in the x-y plane, requires an equation that
explains the magnetization change in the x-y plane as a function of time. In the rotating frame, the total magnetization in the x-
y plane is comprised of two components the “real” and “imaginary” parts. Therefore, the total magnetization in the x-y plane
can be expressed m = u + iv , or u = m − iv . Taking the derivative of this equation with respect to time leads to
xy xt

dmxy

dt
=
du
+i
dt
. Using the previous relationships, the Bloch equations for the two sites can be simplified and rearranged to
dv

dt

give the magnetization in the x-y place as a function of time. Invoking the law of detailed balance, which states that the
exchange rate of the metal site times the amount of 17O at this site is equal to the exchange rate of the bulk water site times the
amount of 17O at this site, will eliminate one of the rate coefficients, simplifying the equations even further gives Equation 7.
Equation 7:
¯
d [ mxy,W  mxy,M ] = − [ i L ¯ ¯
+ R + k ] [ mxy,W  mxy,M ] (NMR.9)

where,
¯
L = [ iW (ωrf − ωo ) 0 0 iM (ωrf − ωo ) ] (NMR.10)

1 1
R̄ = [ 0 0 ] (NMR.11)
T2W T2M

→ ← → ←
k̄ = [ ] (NMR.12)
k − k  − k k

¯
L is the difference in the chemical shifts of the two sites signals, R ¯
is the relaxation in magnetization at each site without
exchange, and k ¯
is the rate coefficients for the exchange. Taking the derivative of the equation, the magnetization of the bulk
water signal and the magnetization of the metal site as a function of time results in Equation 8.
Equation 8:
o
mxy,W (t) m
¯ ¯ ¯ xy,W
−[ iL +R+k ]t
[ ] =e [ ] (NMR.13)
o
mxy,M (t) m
xy,M

The m is the initial magnetization along the x-y plane before relaxation.
o
xy

Using the Bloch-McConnell equation, the width and intensities of the peaks in the spectra become a function of the chemical
exchange. By studying the change in the two peaks, the rate coefficients can be determined, which can be used to calculate
other thermodynamic properties like entropy and enthalpy.

Experiments

5/28/2021 NMR.2 https://chem.libretexts.org/@go/page/1830


T2 Studies
The most common way of studying chemical kinetics in NMR has been through the bandshape technique, which studies the
change in the signals of the spectra as a result of exchange kinetics. Before any exchange occurs, two sharp signals are present,
one for each of the two 17O sites. As the exchange rate speeds up, the two peaks will begin to broaden and overlap. At an
extremely fast exchange rate, the peaks will coalesce and be centered at the weighted average of the Larmor frequency of the
bulk water and the bound water. This occurs because the two exchange sites will have two distinct peaks at slow exchange
rates because the NMR active exchange species will be at each site for long enough time that detection of two separate sites
will occur. As the exchange rate increases, the nucleus will be exchanging so quickly that the detection of the nucleus on either
site becomes averaged out, creating one signal in the spectra.
Figure 1: Coalescing Peaks
This figure shows the two peaks moving closer together and then coalescing as the exchange rate increases.

The figure above shows peaks from a two-site exchange. The top spectrum depicts a slow exchange rate. As they move down,
the rate constants are getting increasingly larger until the peaks coalesce.
The width of the band at half height, or the full width at half max (FWHM) is used to track the rate coefficients of the
exchange because it is proportional to both T2 relaxation and the rate coefficient. During exchange rate experiments, Equation
9 can be used to calculate the exchange rate (k), which can then be used to determine other activation parameters such as
Gibb's free energy, entropy, and enthalpy.
Equation 9:
2
πδv
k = (NMR.14)
∗ o
2(ω −ω )

Swift and Connick used bandshape experiments to study the exchange of water from the bulk to a paramagnetic metal. They
developed an equation that relates the band width of the signal to the mole fraction of each water site, the bulk P and the W

metal P . Equation 10 is the simplified Swift-Connick equation for the change in the paramagnetic water signal. ( Δw is the
M

change in the full width measured at half height)


Equation 10:
1 Δw
(Δw) = (NMR.15)
2
PM 1 Δw
[1 + ]+
˙ 2
kM T 2M kM

By obtaining the NMR spectra and measuring the signal width, the exchange rate can be calculated. It can be seen that the
peak broadness is a function of the T relaxation, or the transverse relaxation. It can be evaluated from the free-induction
2M

decay (FID). The FID is the time-domain signal of each frequency component. Each component results in a sine wave, which
are then added together for the FID signal. At time zero, all the components are aligned, and over time they spread out and
combine in a deconstructive manner, resulting in a decay of the total FID signal. This is an exponential decay and can be used
to calculate the transverse relaxation with Equation 11.
Equation 11:

5/28/2021 NMR.3 https://chem.libretexts.org/@go/page/1830


−t

T
Mxy (t) = Mo e 2
(NMR.16)

The FID is obtained by using a simple pulse directed along the x-axis to tip the magnetization into the y-axis so it is
π

processing in the x-y plane. The magnetization in the x-y plane is detected over time as the different frequency components
process at different rates. This will result in the FID so T can be found.
2M

T1 Studies
A second method of observing the magnetization exchange between bulk solvent and the metal is by studying the
\T_{1}\)relaxation. This method is completed by observing the intensity of one site's signal, while the other signal is saturated
by the applied radio frequency pulse program. As the saturated site's spin is transferred to the second exchange site, the second
site's magnetization intensity increases as result of the additional spin, while the saturated site's magnetization decreases. From
the Bloch equations, the relaxation of magnetization in the z-axis is proportional to . Studying the T relaxation tracks the
1

T1M
1

magnetization change between the metal and bulk site through an inversion-recovery NMR experiment. This experiment is
done using a two-pulse pulse sequence. A 180 shape pulse with a radio frequency close to the bulk solvent's Larmor
o

frequency is directed at the sample. Since a shape pulse is a selective pulse, it will only flip the magnetization of the bulk
solvent. The solvent’s magnetization, having been flipped 180 degrees, will be aligned opposed to the applied magnetic field.
Once probed, the magnetization will begin to relax along the z-axis until it reaches it equilibrium position. After some time (t),
a 90 square pulse is applied. A square pulse is not selective and excites a broader range of frequencies, resulting in both the
o

metal and solvent magnetizations to be tipped into the x-y plane to be detected. At time zero, the magnetization signal for the
solvent will be large and negative because it will tipped into the x-y plane in the negative direction.

Figure 2: Magnetization as a result of the pulse program


The bulk solvent magnetization is tipped to be against the static magnetic field, while the metal site magnetization remains inline with the static
magnetic field. Then the short 900 pulse is applied to tip both site's magnetizations into the x-y plane for detection. The bulk solvent magnetization
will result in a negative signal, and the metal site magnetization will be positive.

As exchange occurs, the magnetization along the z-axis will become positive because of natural relaxation, and because the
17
O from the metal will be positively in the z-direction, helping speed up the relaxation. This can be tracked by varying the
time between the 180 and 90 pulse, since the direction (positive or negative) and intensity of the peaks will change.
o o

Figure 3: Magnetization along z-axis


The magnetization of the bulk water site, directed against the static magnetic field, will begin to relax back towards the initial condition (all
magnetization direct with the static magnetic field). The relaxation will cause the the bulk water site's negative signal to decrease in size and then
become positive as the oxygen on the metal site exchange.

5/28/2021 NMR.4 https://chem.libretexts.org/@go/page/1830


The above image shows that at t=0 the solvent peak is large and negative, since no magnetization will have had time to relax.
Between t=1 and t=2, the magnetization along the z-axis has inverted back, and will produce a positive peak.
The exponential line from the first negative peak to the last positive peak results in Equation 12.
Equation 12:
t

Mz (t) = Mo (1 − e T
1 ) (NMR.17)

Figure 4: T1 equation representation


As the magnetization relaxes along the z-axis, it relaxes exponentially with Equation 12

T1 can then be used to calculate the exchange rate through the Bloch-McConnell equations. A plot of the intensity of
magnetization as a function of time would look like:
Figure 5: Plot of intensity versus time for the bulk solvent site magnetization

These are just two pulse sequences that can be used to study kinetics through NMR. More intricate pulse sequences can be
used to perform kinetic studies on more complex systems.

NMR Kinetic Studies


NMR studies have been carried out to understand the kinetics water exchange with different compounds as a function of
temperature and pressure. There also have been experiments that track the affects of pH on the exchange rates of chemical
systems. Below is a list of some journal articles containing dynamic NMR experiments with 17O. A quick search on DNMR
experiments will produce many journal articles that have been published over the years.
Phillips, Brian L., Susan Neugebauer Crawford, and William H. Casey. "Rate of water exchange between Al(C2O4)(H2O)4+(aq)
complexes and aqueous solutions determined by 17O-NMR spectroscopy." Geochimica et Cosmochimica Acta 61.23 (1997): 4965-
4973.
Szab, Zolt N., Ingmar Grenthe. "On the mechanism of oxygen exchange between Uranyl(VI) Oxygen and water in strongly
alkaline solution as studied by 17O NMR magnetization transfer." Inorganic Chemistry 49.11 (2010): 4928-4933.

References
1. Sandström, J. (1982). Dynamic NMR spectroscopy . London: Academic Press.
2. H. M. McConnell, Journal of Chemical Physics 1958, 28, 430.
3. T. J. Swift, G. M. Anderson, R. E. Connick, M. Yoshimine, Journal of Chemical Physics 1964, 41, 2553.

5/28/2021 NMR.5 https://chem.libretexts.org/@go/page/1830


4. T. J. Swift, R. E. Connick, Journal of Chemical Physics 1962, 37, 307.

Outside Links
http://en.Wikipedia.org/wiki/Nuclear_magnetic_resonance
Pulse Sequence Library: http://nmrwiki.org/wiki/index.php?title=Special:PulseSequenceDatabase
FID: http://en.Wikipedia.org/wiki/Free_induction_decay

Problems
1. Describe the difference between the T1 and T2 relaxation.
2. Why are the Bloch-McConnell equations needed for kinetic NMR studies?
3. Describe what would happen in an inversion-recovery experiment if the signals from the solvent and metal site occur at
similar frequencies?
4. What problems can arise from using the Swift-Connick equation and the bandshape technique?
5. Draw the pulse sequence of the inversion-recovery experiment. (A schematic of the frequency versus time)

Solutions
1. The T1 relaxation is relaxation of the magnetization in the z-axis and is known as longitudinal relaxation. The T2 relaxation
is relaxation in the x and y axis, or xy plane. It is known as transverse relaxation.
2. The Bloch-McConnell equations are the extension to the Bloch equation. They include an extra term that takes into account
exchange between two species. Without this relationship, kinetic studies would not be able to be studied using NMR.
3. If the signals are too close together, the 180 degree shape pulse would flip both magnetizations, not separating the two to
allow for analysis. More intricate pulse sequences can be used with more pulses to obtain the magnetizations in opposite
directions along the xy plane.
4. NMR signals are often small since its a very insensitive technique. Peaks can not only be easily lost in the noise, but the
bandwidth may be extremely difficult to determine, making the Swift-Connick equation hard to use.
5. Pulse sequence drawing for the inversion-recovery experiment

Contributors
Adele Panasci

5/28/2021 NMR.6 https://chem.libretexts.org/@go/page/1830


Nuclear Overhauser Effect
Page Under Construction
1. Introduction
2. The Nuclear Overhouser Effect (NOE)
3. Heading #2
4. References
5. Outside Links
6. Problems
7. Contributors and Attributions

Introduction
The Nuclear Overhouser Effect (NOE)
The Nuclear Overhauser Effect (NOE) uses a second radio frequency field, usually relatively stronger than the first one, in
order to saturate particular resonances. This phenomenon discovered by Albert Overhaouser in 1953.The NOE process can be
summarized as a cross relaxation from one spin state to another spin state. The effect causes the perturbation via dipolar
interactions with further nucleus spins and increases the intensity of other spins. The effect is called the steady-state Nuclear
Overhauser Effect. Since dipolar coupling interacts throughout the space, it is a very useful technique to study the
conformation of molecules.
The significant interaction between the magnetic dipoles of the two spins is required for NOE. Because of the fact that dipolar
interactions drop of very fast with distance, 1H, 1H NOE can only occur with a distance of <5-6 Å (500-600 pm).

Heading #2
Rename to desired sub-topic. You can delete the header for this section and place your own related to the topic. Remember to
hyperlink your module to other modules via the link button on the editor toolbar.

References
1. This is meant for references used for constructing the module. They must be primary and accessible to readers at a library.
2. You need at least two different sources here. Websites are not allowed. DOI links to J. Chem. Ed. are ideal Do not reference
class notes. Also, do not reference textbooks for maximal credit. Using the insert citation button to automatically handle
references is highly suggested (bottom right button on editor toolbar).

Outside Links
This is not meant for references used for constructing the module, but as secondary and unvetted information available at
other site
Link to outside sources. Wikipedia entries should probably be referenced here.

Problems
Be careful not to copy from existing textbooks. Originality is rewarded. Make up some practice problems for the future
readers. Five original with varying difficulty questions (and answers) are ideal.

Contributors and Attributions


Name #1 here (if anonymous, you can avoid this) with university affiliation

6/29/2021 1 https://chem.libretexts.org/@go/page/1831
Solomon Equations
Introduction
The Solomon equations describe the relaxation between two coupled spins, I and S. This page will be dedicated to developing
a theoretical treatment beginning from the Bloch equations. We will investigate the ramifications of the relaxation and the
consequences on the overall relaxation rates.

Theoretical Treatment
We begin by describing a system of two spins, I and S. Each spin has two states; low energy α and high energy β and we can
then come up with a 4 level energy system for these two spins. This is shown in the figure below.

Figure 1. Energy level diagram for a 2 spin system.


A spin can have allowed relaxation induced transitions between the α and β states which occur at a rate (W1IS) as well as
forbidden relaxation induced transitions between like states (W2), which are shown in the figure below. Forbidden transitions
W2 corresponds to both spins flipping, known as a flip-flip transition, while W0 is where each spin flips between states, known
as a flip-flop transition. The superscripts denote the order (zero, single, double) type of transition.

Figure 2. Transition rates of relaxation induced transitions.


We can now define the change in the population of each level in with 1=αα 2=βα, 3= αβ and 4=ββ during a given amount of
time. For example, for energy level 1, the population is initially n10. As time progress, the some of this population will go to
energy levels 2, 3, and 4 at rates, W1I, W1S, W2, respectively. Meanwhile the state will gain population from states 2, 3, and 4
at at rates, W1I, W1S, W2, respectively. It may seem that this would give a net result of 0, but the populations of each state is
given by the Boltzmann distribution and the argument may be easily made that the rate will depend on the number of spins in
the given energy level. Therefore we may now write:
dn1
1 1 2 1 1 2
= −W n1 − W n1 − W n1 + W n3 + W n2 + W n4 (1)
S I S I
dt

(eq. 1)
the rates for the other energy levels may then be written as:
dn2 0 0
1 2 2 1
= +W n1 + W n4 + W n3 − W n2 − W n2 − W n2 (2)
I S S I
dt

(eq. 2)
dn3
1 2 0 2 1 0
= +W n1 + W n4 + W n2 − W n3 − W n3 − W n3 (3)
S I I S
dt

6/21/2021 1 https://chem.libretexts.org/@go/page/61683
(eq. 3)
dn4 2 2
2 2 2 2
= −W n4 − W n4 − W n4 + W n3 + W n2 + W n1 (4)
S I I S
dt

(eq. 4)
We can then calculate the spin magnetization for the I and S spins, along the Z axis. The reason for the Z axis is because we
are concerned with the recovery of the magnetization to it's equilibrium value. Therefore, the net magnetization for spin I will
be the difference in populations between the I transitions and is expressed as

Iz = n1 − n2 + n3 − n4 (5)

(eq. 5)
and spin S will be
Sz = n1 − n3 + n2 − n4 . (6)

(eq. 6)
Further we can define the difference in populations between the 2 I-spin transitions, n1-n2-n3+n4, which is the same for the 2 S-
spin transitions and will be denoted 2IzSz.
2 Iz Sz = n1 − n2 − n3 + n4 . (7)

(eq. 7)
The total population of the system is then given by
T = n1 + n2 + n3 + n4 (eq. 8)
And ni can then be re-written as a combination of Iz, Sz, 2IzSz, and T such that
1
n1 = (T + Iz + Sz + 2 Iz Sz ) (eq. 9)
4

1
n2 = (T − Iz + Sz − 2 Iz Sz ) (eq. 10)
4

1
n3 = (T + Iz − Sz − 2 Iz Sz ) (eq. 11)
4

1
n4 = (T − Iz − Sz + 2 Iz Sz ) (eq. 12)
4

Which can be verified by plugging equations 5-8 into equations 9-12.

Derivation of The Solomon Equations


The Solomon equations are the time-dependent operators Iz, Sz, and 2IzSz expressed in terms of W and the operators. We will
take the following approach to derive each operator:
1) Take the time derivative of each operator.
dni
2) Express the derivative in terms of W by substituting equations 1-4 for .
dt

3) Group the ni terms.


4) Substitute the equations 9-12 for ni and group the like W terms. This will give the final equations.
Each step will be numbered according to this list.
Derivation of Iz
dIz dn1 dn2 dn3 dn4
1) = − + −
dt dt dt dt dt

2) = −W 1
S
n1 − W
1
I
n1 − W
2
n1 + W
1
S
n3 + W
1
I
n2 + W
2
n4

1 2 0 2 1 0
−(+W n1 + W n4 + W n3 − W n2 − W n2 − W n2 )
I S S I

6/21/2021 2 https://chem.libretexts.org/@go/page/61683
1 2 0 2 1 0
+(+W n1 + W n4 + W n2 − W n3 − W n3 − W n3 )
S I I S

2 2 2 2 2 2
−(−W n4 − W n4 − W n4 + W n3 + W n2 + W n1 )
S I I S

dIz
3) = −2 n1 [ W
1
I
+W
2
] + 2 n2 [ W
1
I
+W
0
] − 2 n3 [ W
I
2
+W
0
] + 2 n4 [ W
I
2
+W
2
]
dt

1
4) W 2
[−2 ∗ (T + Iz + Sz + 2 Iz Sz ) + 2
1

4
(T − Iz − Sz + 2 Iz Sz )] = W
2
(−Iz − Sz )
4

1 1
1 1
W [−2 (T + Iz + Sz + 2 Iz Sz ) + 2 (T − Iz + Sz − 2 Iz Sz ) = W (−Iz − 2 Iz Sz )
I 1
4 4

1 1
0 0
W [2 (T − Iz + Sz − 2 Iz Sz ) + −2 (T + Iz − Sz − 2 Iz Sz )] = W (−Iz + Sz )
4 4

1 1
2 2
W [−2 (T + Iz − Sz − 2 Iz Sz ) + 2 (T − Iz − Sz + 2 Iz Sz )] = W (−Iz + 2 Iz Sz )
I I
4 4

dIz
1 2 2 0 2 0 1 2
= −(W +W +W +W )Iz − (W −W )Sz − (W −W )2 Iz Sz
I I I I
dt

Derivation of Sz
Sz = n1 − n3 + n2 − n4 . (8)

dSz n1 n3 n2 n4
1) = − + −
dt dt dt dt dt

dSz
2) = [−W
1
S
n1 − W
I
1
n1 − W
2
n1 + W
1
S
n3 + W
1
I
n2 + W
2
n4 ]
dt

1 2 0 2 1 0
−[+W n1 + W n4 + W n2 − W n3 − W n3 − W n3 ]
S I I S

1 2 0 2 1 0
+[ W n1 + W n4 + W n3 − W n2 − W n2 − W n2 ]
I S S I

2 2 2 2 2 2
−[−W n4 − W n4 − W n4 + W n3 + W n2 + W n1 ]
S I I S

dSz
3) = −2 n1 [ W
S
1
+W
2
] −2 n2 [ W
0
+W
2
S
] +2 n3 [ W
S
1
+W
0
] +2 n4 [ W
2
+W
2
S
]
dt

1
4) 2 W
S
1
[(−T − Iz − Sz − 2 Sz Iz ) + (T + Iz − Sz − 2 Iz Sz )] = W
S
1
(−Sz − 2 Iz Sz )
4

1
2 2
2 W [(−T − Iz − Sz − 2 Sz Iz ) + (T − Iz − Sz + 2 Iz Sz )] = W (Iz + Sz )
4

1
0 0
2 W [(−T + Iz − Sz + 2 Sz Iz ) + (T + Iz − Sz − 2 Iz Sz )] = W (Iz − Sz )
4

1
2 2
2 W 1[(−T − Iz + Sz + 2 Sz Iz ) + (T − Iz − Sz + 2 Iz Sz )] = W (−Sz − 2 Iz Sz )
S S
4

dSz
1 2 0 2 2 1 0 2
= −Sz (W +W +W +W ) + 2 Iz Sz (W −W ) + Iz (W −W )
S S S S
dt

Deriving 2IzSz
2 Iz Sz = n1 − n2 − n3 + n4 . (9)

d2Iz Sz n1 n2 n3 n4
1) = − − +
dt dt dt dt dt

d2Iz Sz
2) = [−W
1
S
n1 − W
1
I
n1 − W
2
n1 + W
1
S
n3 + W
1
I
n2 + W
2
n4 ]
dt

1 2 0 2 1 0
−[ W n1 + W n4 + W n3 − W n2 − W n2 − W n2 ]
I S S I

1 2 0 2 1 0
−[ W n1 + W n4 + W n2 − W n3 − W n3 − W n3 ]
S I I S

2 2 2 2 2 2
+[−W n4 − W n4 − W n4 + W n3 + W n2 + W n1 ]
S I I S

d2Iz Sz
3) = −2 n1 (W
S
1
+W
1
I
) + 2 n2 (W
I
1
+W
S
2
) + 2 n3 (W
S
1
+W
I
2
) − 2 n4 (W
2
S
+W
2
I
) .
dt

6/21/2021 3 https://chem.libretexts.org/@go/page/61683
1 1
4)W S
1
[−2 (T + Iz + Sz + 2 Iz Sz ) + 2 (T + Iz − Sz − 2 Iz Sz )] = −W
S
1
[ Sz + 2 Iz Sz ]
4 4

1 1
1 1
W [−2 (T + Iz + Sz + 2 Iz Sz ) + 2 (T − Iz + Sz − 2 Iz Sz )] = W [ Iz + 2 Iz Sz ]
I I
4 4

1 1
2 2
W [−2 (T − Iz + Sz − 2 Iz Sz ) − 2 (T − Iz − Sz + 2 Iz Sz )] = −W [ Sz − 2 Iz Sz ]
S S
4 4

1 1
2 2
W [−2 (T + Iz − Sz − 2 Iz Sz ) − 2 (T − Iz − Sz + 2 Iz Sz )] = W [ Iz − 2 Iz Sz ]
I I
4 4

d2Iz Sz
2 1 2 1 2 1 2 1
= Sz (W −W ) + Iz (W −W ) − 2 Iz Sz (W +W +W +W )
S S I I S S I I
dt

Relaxation Mechanisms of the Solomon equations


The Solomon equations for operators Iz, Sz, and 2IzSz are given below (in case the reader skips the derivation)
Iz
1 2 2 0 2 0 1 2
= −(W +W +W +W )Iz − (W −W )Sz − (W −W )2 Iz Sz
I I I I
dt

Sz
1 2 0 2 2 1 0 2
= −Sz (W +W +W +W ) + 2 Iz Sz (W −W ) + Iz (W −W )
S S S S
dt

2Iz Sz
2 1 2 1 2 1 2 1
= Sz (W −W ) + Iz (W −W ) − 2 Iz Sz (W +W +W +W )
S S I I S S I I
dt

Now, these equations are not quite right, as the treatment that we gave actually considers only the perturbed spins. Therefore
we may now re-write these equations to account for the the spins that are perturbed by defining the perturbed spins as the
perturbed population+equilibrium population=Iz, while the equilibrium population is I0z. The equations may now be re-written
as
0
Iz − Iz
1 2 2 0 0 2 0 0 1 2
= −(W +W +W +W )(Iz − Iz ) − (W −W )(Sz − Sz ) − (W −W )
I I I I
dt

0
Sz − Sz
0 1 2 0 2 2 1 0 0 2
= −(Sz − Sz )(W +W +W +W ) + 2 Iz Sz (W −W ) + (Iz − Iz )(W −W )
S S S S
dt

2Iz Sz
0 2 1 0 2 1 2 1 2 1
= (Sz − Sz )(W −W ) + (Iz − Iz )(W −W )2 Iz Sz (W +W +W +W )
S S I I S S I I
dt

Self Relaxation
The Solomon equations for a given spin, here described for Iz but applicable for Sz and 2IzSz, show that the rate of relaxation
is tied to Iz, Sz, and 2IzSz. The rate constants associated with Iz corresponds to the self relaxation of the spin magnetization on
its own. This is given by the overall rate
1 2 2 0
W +W +W +W
I I

Note that the rates W2 and W0 correspond to processes cross-relaxation processes that need a spin S to occur. Therefore, the
relaxation rate for Iz is not dependent on any other spins in the system. The summation of the rate constants is denoted by RI.

Cross-Relaxation
The term (W0-W2)Sz is present in the Solomon equations for Iz and corresponds to the the rate at which Sz transfers
magnetization to Iz spin magnetization in a phenomenon is known as cross-relaxation. To properly illustrate the effect, take
= 0 . In this case, relaxation of a spin I is independent of spin S. On the other hand, if W and/orW ≠ 0 but
2 0 2 0
W =W

spin S is not perturbed, (i.e, (S − S ) = 0 ) cross relaxation will not occur either. The cross relaxation term is denoted as σ .
z
0
z IS

2IzSz Relaxation
The relaxation associated with W − W is the transfer of magnetization from IzSz to I, denoted as Δ , which will only occur
1
I I
2
I

if the rate constants are different. Meanwhile the operatore 2IzSz also shows self-relaxation as RIS, denoted by
1 2 1 2
RIS = W +W +W +W
I I S S

References

6/21/2021 4 https://chem.libretexts.org/@go/page/61683
1) I.Solomon. "Relaxation Processing in a System of Two Spins" Phys. Rev. 99, 1955, 559.
2) J. Keeler. "Relaxation" Chapter 8 2004
3) M. Levitt. "Spin Dyanmics"

Contributions

6/21/2021 5 https://chem.libretexts.org/@go/page/61683
Spin-Spin Relaxation
Page Under Construction
1. Introduction
2. T2 and the Bloch Equations
3. T2 Measurments
4. References
5. Outside Links
6. Problems
7. Contributors and Attributions

Introduction
In this section, give a short introduction to your topic to put it in context. The module should be easy to read; please reduce
excess space, crop figures to remove white spaces, full justify all text, with the exception of equations, which should use the
equation command for construction (see FAQ).

T2 and the Bloch Equations


Rename to desired sub-topic. This is where you put the core text of your module. Add any number of headings necessary for
your topic. Try to reduce unnecessary discussion and get to the point in a terse, yet informative, manner possible.

T2 Measurments
Rename to desired sub-topic. You can delete the header for this section and place your own related to the topic. Remember to
hyperlink your module to other modules via the link button on the editor toolbar.

References
1. This is meant for references used for constructing the module. They must be primary and accessible to readers at a library.
2. You need at least two different sources here. Websites are not allowed. DOI links to J. Chem. Ed. are ideal Do not reference
class notes. Also, do not reference textbooks for maximal credit. Using the insert citation button to automatically handle
references is highly suggested (bottom right button on editor toolbar).

Outside Links
This is not meant for references used for constructing the module, but as secondary and unvetted information available at
other site
Link to outside sources. Wikipedia entries should probably be referenced here.

Problems
Be careful not to copy from existing textbooks. Originality is rewarded. Make up some practice problems for the future
readers. Five original with varying difficulty questions (and answers) are ideal.

Contributors and Attributions


Name #1 here (if anonymous, you can avoid this) with university affiliation

6/23/2021 1 https://chem.libretexts.org/@go/page/1832
Spin Lattice Relaxation
Page Under Construction!

Contributors and Attributions


Name #1 here (if anonymous, you can avoid this) with university affiliation

6/13/2021 1 https://chem.libretexts.org/@go/page/1833
Rotations and Irreducible Tensor Operators
Page under construction!

Overview
Irreducible Tensor Operators are extremely valuable to help reduce complex mathematical problems commonly found in NMR. This
page will be devoted to deriving the tensors and showing how to use the tensors in calculations.

Cartesian Rotations
Irreducible Tensor Operators operate on angular momenta, which is a fairly abstract concept to understand. We begin this discussion
by investigating rotations in 2 and 3 dimensions so the reader understands a physical picture before moving into a complex quantum
mechanical treatment.

2D Rotations
Let's begin with the simple example of the hand of a clock. The clock hand is a vector which has a magnitude (the length of the hand)
and a direction (the direction the arrow is pointing). The vector is rotating in a 2D plane (the clock face). We define our cartesian axes
that the y axis lies along 12 and the x axis is along 3 on the clock. Initally, lets say it is 3:00 (t=0) and the vector lies along the x-axis
and rotates counter-clockwise, then a some time, (t=2.2hrs) later the vector points along some arbitrary angle theta from the inital x-
axis. Taking a snapshot of the vector at this time, we can see that vector a can now be described by
^ ^ ^ ^ ^
r ⃗ = a i + b j = r i cos(θ) j + r sin(θ) j (1)

insert picture
We can actually describe this vectors position using any arbitrary axis (time) for x and y as long as x and y remain orthogonal. If for
example the axes are rotated by ϕ then the vector a time=2.2hrs is

r ⃗ = c ^
i + d^
j = r^
i cos(θ − ϕ)^
j + r sin(θ − ϕ)^
j (2)

insert picture
Using the following triognometric identities, we can solve for a ,b, c, and d.
cos(θ − ϕ) = cos θ cos ϕ + sin θ sin ϕ (3)

sin(θ − ϕ) = sin θ cos ϕ − cos θ sin ϕ (4)

the it follows that


a = r cos θ (5)

b = r sin θ (6)

c = r cos(θ − ϕ) = r cos θ cos ϕ + r sin θ sin ϕ = a cos ϕ + b sin ϕ (7)

d = r sin(θ − ϕ) = r sin θ cos ϕ − r cos θ sin ϕ = b cos ϕ − a sin ϕ (8)

From inspection we can see that this can be readily transformed into a matrix
c cos ϕ sin ϕ a
( ) =( )( ) (9)
d − sin ϕ cos ϕ b

In other word we can rotate r into r' by applying a rotation trnasformation R or



r = Rr (10)

where R has the property



RR =1 (11)

leading to

a cos ϕ − sin ϕ c
( ) =( )( ) (12)
b sin ϕ cos ϕ d

5/28/2021 1 https://chem.libretexts.org/@go/page/11906
3D Rotations
In 3 dimensions rotations are slightly more complex. We need to specify another dimension in which the object can be rotated. In order
to do this we need to employ three angles α, β, γ, commonly referred to as the euler angles. α is rotation about the original z-axis,
beta is rotation about y' and γ is rotation about z'. How the angle correspond to rotations are shown below.

insert figure
Our vector, r, may now be described as
^
r ⃗ = a^
i + b^
j + ck (13)

We can now describe rotations R when one of the 3 axes is fixed If Z axis is fixed,
cos α sin α 0
⎛ ⎞
Rz = ⎜ − sin α cos α 0⎟ (14)

⎝ ⎠
0 0 1

When Y axis is fixed


cos β 0 − sin β
⎛ ⎞
Ry = ⎜ 0 1 0 ⎟ (15)

⎝ ⎠
sin β 0 cos β

When X axis is fixed


cos γ sin γ 0
⎛ ⎞
Rz = ⎜ − sin γ cos γ 0⎟ (16)

⎝ ⎠
0 0 1

Then rotation of the original axis system to the new axis system is then
cos α sin α 0 cos β 0 − sin β cos γ sin γ 0
⎛ ⎞⎛ ⎞⎛ ⎞
R = Rx Ry Rz = ⎜ − sin α cos α 0⎟⎜ 0 1 0 ⎟ ⎜ − sin γ cos γ 0⎟ (17)

⎝ ⎠⎝ ⎠⎝ ⎠
0 0 1 sin β 0 cos β 0 0 1

Angular Momentum Rotations


J=1/2
In a similar manner we can rotated angular momentum states. Like we developed a rotation operator, R for Cartesian space, we can
define a new rotation operators which operates on angular momentum states using the Euler relations we made for 3D rotations.
iγJz iβJy iαJx
D(α, β, γ) = e e e (18)

Lets examine what happens when D operates on |J,m>


′ ′ ′ ′
D(α, β, γ) = ∑ | J , m >< J , m |D(α, β, γ)|J, m > (19)
′ ′
J ,m

The operator can only operate on a \single J value reducing the above equation to
′ ′ (J) ′
D(α, β, γ) = ∑ < J, m |D(α, β, γ)|J, m > |J, m = ∑ D (α, β, γ)|J, m > (20)
m′ ,m
′ ′
m m

where
(J) ′
D ′ (α, β, γ) =< J, m |D(α, β, γ)|J, m > (21)
m ,m

which can be simplified to


(J) ′ iγJz iβJy ′ ′ iαJz
D ′ (α, β, γ) = ∑ < J, m | e |J, n >< J, n| e |J, n >< J, n | e |J, m > (22)
m ,m
′ ′
n ,n

It is then realized that


2 2
(iαJz ) (iαm)
iαJz 1αm
e |J, m >= [1 + iα Jz + +. . . ]|J, m >= [1 + iαm + +. . . ]|J, m = e |J, m (23)
2! 2!

5/28/2021 2 https://chem.libretexts.org/@go/page/11906
Then
′ ′
J im γ ′ iβJy imα im γ J imα
D ′
(α, β, γ) = e < J, m | e |J, m > e =e d ′
(β)e (24)
m ,m m ,m

where
J ′ iβJy
d ′ (β) =< J, m | e |J, m > (25)
m ,m

We now must calculate d J


m ,m

(β) for a arbitrary J. Lets examine the case where J=1/2 for rotation operator d 1/2
(β)

1/2 iβJy
d (β) = e (26)

Taking a Taylor expansion gives

iβJy
1 2
1 3
2 3
e = 1 + iβ Jy − β Jy − β j (27)
Y
2! 3!

2 3
β β

1 0 2 0 1 β 2
=( ) (1 − +. . . ) + ( )( − +. . . ) (28)
0 1 2! −1 0 2 3!

We then see that multiplying the Taylor expansion of e iβJy


by the respective matrices, we obtain the Taylor expansions of \cos and \sin
which can be rewritten as:

cos(β/2) sin(β/2)
=( ) (29)
− sin(β/2) cos(β/2)

Then we have obtained the values for d


(1/2)
d = cos(β/2) (30)
1/2,1/2

(1/2)
d = sin(β/2) (31)
1/2,−1/2

J>1/2
Now we must calculate d J
m ,m
′ (β) for J>1/2. To do this we must first uncouple J and m'
J J ′
D ′ (α, β, γ)|J, m >= ∑ D ′ (α, β, γ)|J, m > (32)
m ,m m ,m

m

j2 −j1 −m − −−−−− j1 j2 J
= (−1 ) √ 2J + 1 ∑ m1 , m2 D(α, β, γ)| j1 , m1 > D(α, β, γ)| j2 , m2 > ( ) (33)
m1 m2 −m

j2 −j1 −m − −−−−− ( j1 ) ( j2 ) j1 j2 J
′ ′ ′ ′
= (−1 ) √ 2J + 1 ∑ m1 , m2 , m , m D(α, β, γ ) ′ D(α, β, γ ) ′ ( ) | j1 m ; j2 , m > (34)
1 2 m m1 m m2 1 2
1 2
m1 m2 −m

Taking this result and left multiplying by <Jm'| and uncoupling J and m' to the other j and m states results in the expression for
(J)
D(α, β, γ)
m′ ,m

(J) j −j −m − −−−−− ( j1 ) ( j2 ) j1 j2 J
′ ′
D(α, β, γ ) ′ = (−1 ) 2 1
√ 2J + 1 ∑ m1 , m2 , m , m D(α, β, γ ) ′ D(α, β, γ ) ′ ( ) < J, (35)
m ,m 1 2 m m1 m m2
1 2
m1 m2 −m

′ ′ ′
m || j1 m ; j2 , m >
1 2

j1 j2 J j1 j2 J ( j1 ) ( j2 )
= (2J + 1) ∑ ( )( ) D(α, β, γ ) ′ D(α, β, γ ) ′ (36)
′ ′ ′ m m1 m m2
m1 m2 −m m m −m 1 2

m1 , m2 1 2

(J)
From this result we can calculate any D(α, β, γ) for any J value. This is important since the observables we wish to calculate in
m ,m

NMR are simply rotations of the angular momentum. Looking at the time-dependant Schrodinger equation
d
|ψ(t) >= −iH |ψ(t) > (37)
dt

iH t
|ψ(t) = e |ψ(t) > (38)

Lets now examine what we can do by rotating the angular momentum vector J,m to J m'. Looking at the z component of J we know

< J, m | Jz |J, m >= m δm′ ,m (39)

5/28/2021 3 https://chem.libretexts.org/@go/page/11906
which results in a diagonal matriz for J_z with the digonal elements being the m quantum numbers.
Then J_z may be written as

Jz = ∑ m|J, m >< J, m| (40)

which has the complex conjugate


need to figure this out!
Now we have created the operatore for J_z.. In principle we can create an operator for any combination of |J'm'><J,m|. Now lets see
what happens weh nwe rotate the operator in which the states are weighted by the angular momentum characteristics. We choose to
(1)
label the new operator T (2)
(3, 4) by the following indicie.

k = J +J (41)

The Total Angular Momentum



Q = m −m (42)

The difference in M values


J (43)

The initial and



J (44)

the final angular momentum.



(k) ′
− −−−− ′
J −m
′ J J k ′ ′
T (J, J ) = √ 2k + 1 ∑ (−1 ) ( ) | J , m >< J, m| (45)
(Q) ′

m −m −Q
m,m

To rotate this operator we must left multiply by D(α, β, γ) and right multiple by D(α, β, γ)

. This corresponds the rotation of the
final and initial states.

(k) ′ †
− −−−− ′
J −m
′ J J k (J) (J)† ′
D(α, β, γ)T (J, J )(D(α, β, γ ) = √ 2k + 1 ∑ (−1 ) ( ) D(α, β, γ ) ′
D(α, β, γ )m,n | J , (46)
(Q) ′ n ,m
′ ′
m −m −Q
m, m ,n, n


m >< J, m|

where

j1 j2 J j1 j2 J ′
(J)
J1 J2 † m −m
D (α, β, γ)D (α, β, γ ) = ∑ (2J + 1) ( )( ) (−1 ) D(α, β, γ ) (47)
′ ′ ′ −m′ ,m
m M −m m1 M2 −m
J,m,m
′ 1 2

which simplifies the previous equation to



(k) − −−−− ′ ′ J J k (k)
′ † J −n ′ ′
D(α, β, γ)T (J, J )(D(α, β, γ ) = √ 2k + 1 ∑ (−1 ) ( )D (α, β, γ)| J , n >< J, n| (48)
(Q) ′ qQ

n −n −q
n ,n,q

(k)
which is really T (Q)
(J, J )

with Q replaced by q and m',m with n',n or
(k) ′ † (k) (k) ′
D(α, β, γ)T (J, J )(D(α, β, γ ) = ∑D (α, β, γ)T (J, J ) (49)
(Q) qQ (q)

This is a very unique property of these operators! Rotation of one operator produces another operator that is weighted by
(k) (J)
D
qQ
(α, β, γ) . It is now trivial to create analytical solutions as D mm
′ (α, β, γ) are known.

Examples Using Irreducible Tensor Operators

5/28/2021 4 https://chem.libretexts.org/@go/page/11906
NMR: Introduction
Nuclear Magnetic Resonance (NMR) is a nuceli (Nuclear) specific spectroscopy that has far reaching applications throughout
the physical sciences and industry. NMR uses a large magnet (Magnetic) to probe the intrinsic spin properties of atomic nuclei.
Like all spectroscopies, NMR uses a component of electromagnetic radiation (radio frequency waves) to promote transitions
between nuclear energy levels (Resonance). Most chemists use NMR for structure determination of small molecules.

Introduction
In 1946, NMR was co-discovered by Purcell, Pound and Torrey of Harvard University and Bloch, Hansen and Packard of
Stanford University. The discovery first came about when it was noticed that magnetic nuclei, such as 1H and 31P (read: proton
and Phosphorus 31) were able to absorb radio frequency energy when placed in a magnetic field of a strength that was specific
to the nucleus. Upon absorption, the nuclei begin to resonate and different atoms within a molecule resonated at different
frequencies. This observation allowed a detailed analysis of the structure of a molecule. Since then, NMR has been applied to
solids, liquids and gasses, kinetic and structural studies, resulting in 6 Nobel prizes being awarded in the field of NMR. More
information about the history of NMR can be found in the NMR History page. Here, the fundamental concepts of NMR are
presented.

Spin and Magnetic Properties


The nucleus consists of elementary particles called neutrons and protons, which contain an intrinsic property called spin. Like
electrons, the spin of a nucleus can be described using quantum numbers of I for the spin and m for the spin in a magnetic
field. Atomic nuclei with even numbers of protons and neutrons have zero spin and all the other atoms with odd numbers have
a non-zero spin. Furthermore, all molecules with a non-zero spin have a magnetic moment, μ , given by

μ = γI (NMR.1)

where γ is the gyromagnetic ratio, a proportionality constant between the magnetic dipole moment and the angular
momentum, specific to each nucleus (Table 1).
Table N M R. 1 : The gyromagnetic ratios for several common nuclei
Nuclei Spin Gyromagetic Ratio (MHz/T) Natural Abundance (%)
1 1/2 42.576 99.9985
H
13
C 1/2 10.705 1.07
31
P 1/2 17.235 100
27
Al 5/2 11.103 100
23
Na 3/2 11.262 100
7
Li 3/2 16.546 92.41
29
Si 1/2 -8.465 4.68
17
O 5/2 5.772 0.038
15
N 1/2 -4.361 0.368

The magnetic moment of the nucleus forces the nucleus to behave as a tiny bar magnet. In the absence of an external magnetic
field, each magnet is randomly oriented. During the NMR experiment the sample is placed in an external magnetic field, B , 0

which forces the bar magnets to align with (low energy) or against (high energy) the B . During the NMR experiment, a spin
0

flip of the magnets occurs, requiring an exact quanta of energy. To understand this rather abstract concept it is useful to
consider the NMR experiment using the nuclear energy levels.

7/4/2021 NMR.1 https://chem.libretexts.org/@go/page/1834


Figure N M R. 1 : Application of a magnetic field to a randomly oriented bar magnet. The red arrow denotes magnetic moment
of the nucleus. The application of the external magnetic field aligns the nuclear magnetic moments with or against the field.

Nuclear Energy Levels


As mentioned above, an exact quanta of energy must be used to induce the spin flip or transition. For any m, there are 2m+1
energy levels. For a spin 1/2 nucleus, there are only two energy levels, the low energy level occupied by the spins which
aligned with B and the high energy level occupied by spins aligned against B . Each energy level is given by
0 0

E = −mℏγB0 (NMR.2)

where m is the magnetic quantum number, in this case +/- 1/2. The energy levels for m > 1/2, known as quadrupolar nuclei,
are more complex and information regarding them can be found here.
The energy difference between the energy levels is then

ΔE = ℏγB0 (NMR.3)

where ℏ is Planks constant.


A schematic showing how the energy levels are arranged for a spin=1/2 nucleus is shown below. Note how the strength of the
magnetic field plays a large role in the energy level difference. In the absence of an applied field the nuclear energy levels are
degenerate. The splitting of the degenerate energy level due to the presence of a magnetic field in known as Zeeman Splitting.

Figure N M R. 2 : The splitting of the degenerate nuclear energy levels under an applied magnetic field. The green spheres
represent atomic nuclei which are either aligned with (low energy) or against (high energy) the magnetic field.

Energy Transitions (Spin Flip)


In order for the NMR experiment to work, a spin flip between the energy levels must occur. The energy difference between the
two states corresponds to the energy of the electromagnetic radiation that causes the nuclei to change their energy levels. For
most NMR spectrometers, B is on the order of Tesla (T) while γ is on the order of 10 . Consequently, the electromagnetic
0
7

radiation required is on the order of 100's of MHz and even GHz. The energy of a photon is represented by

E = hν (NMR.4)

and thus the frequency necessary for absorption to occur is represented as:
γB0
ν = (NMR.5)

For the beginner, the NMR experiment measures the resonant frequency that causes a spin flip. For the more advanced NMR
users, the sections on NMR detection and Larmor frequency should be consulted.

7/4/2021 NMR.2 https://chem.libretexts.org/@go/page/1834


Figure N M R. 3 : Absorption of radio frequency radiation to promote a transition between nuclear energy levels, called a spin
flip.

Nuclear Shielding
The power of NMR is based on the concept of nuclear shielding, which allows for structural assignments. Every atom is
surrounded by electrons, which orbit the nucleus. Charged particles moving in a loop will create a magnetic field which is felt
by the nucleus. Therefore the local electronic environment surrounding the nucleus will slightly change the magnetic field
experienced by the nucleus, which in turn will cause slight changes in the energy levels! This is known as shielding. Nuclei
that experinece differnet magnetic fields due to the local electronic interactions are known as inequivalent nuclei. The change
in the energy levels requires a different frequency to excite the spin flip, which as will be seen below, creates a new peak in the
NMR spectrum. The shielding allows for structural determination of molecules.

Figure N M R. 4 : The effect that shielding from electrons has on the splitting of the nuclear energy levels. Electrons impart
their own magnetic field which shields the nucleus from the externally applied magnetic field. This effect is greatly
exaggerated in this illustration.
The shielding of the nucleus allows for chemically inequivalent environments to be determined by Fourier Transforming the
NMR signal. The result is a spectrum, shown below, that consists of a set of peaks in which each peak corresponds to a distinct
chemical environment. The area underneath the peak is directly proportional to the number of nuclei in that chemical
environment. Additional details about the structure manifest themselves in the form of different NMR interactions, each
altering the NMR spectrum in a distinct manner. The x-axis of an NMR spectrum is given in parts per million (ppm) and the
relation to shielding is explained here.

Figure N M R. 5 : 31P spectrum of phosphinic acid. Each peak corresponds to a distinct chemical environment while the area
under the peak is proportional to the number of nuclei in a given environment.

Relaxation
Relaxation refers to the phenomenon of nuclei returning to their thermodynamically stable states after being excited to higher
energy levels. The energy absorbed when a transition from a lower energy level to a high energy level occurs is released when
the opposite happens. This can be a fairly complex process based on different timescales of the relaxation. The two most
common types of relaxation are spin lattice relaxation (T1) and spin spin relaxation (T2). A more complex treatment of
relaxation is given elsewhere.

7/4/2021 NMR.3 https://chem.libretexts.org/@go/page/1834


Figure N M R. 6 : The process of relaxation
To understand relaxation, the entire sample must be considered. By placing rhe nuclei in an external magnetic field, the nuclei
create a bulk magnetization along the z-axis. The spins of the nuclei are also coherent. The NMR signal may be detected as
long as the spins are coherent with one another. The NMR experiment moves the bulk magnetization from the z-axis to the x-y
plane, where it is detected.
Spin-Lattice Relaxation (T ): T1 is the time it takes for the 37% of bulk magnetization to recovery along Z-axis from the
1

x-y plane. The more efficient the relaxation process, the smaller relaxation time (T1) value you will get. In solids, since
motions between molecules are limited, the relaxation time (T1) values are large. Spin-lattice relaxation measurements are
usually carried out by pulse methods.
Spin-Spin Relaxation (T ): T2 is the time it takes for the spins to lose coherence with one another. T2 can either be shorter
2

or equal to T1.

Applications
The two major areas where NMR has proven to be of critical importance is in the fields of medicine and chemistry, with new
applications being developed daily
Nuclear magnetic resonance imaging, better known as magnetic resonance imaging (MRI) is an important medical diagnostic
tool used to study the function and structure of the human body. It provides detailed images of any part of the body, especially
soft tissue, in all possible planes and has been used in the areas of cardiovascular, neurological, musculoskeletal and
oncological imaging. Unlike other alternatives, such as computed tomography (CT), it does not used ionized radiation and
hence is very safe to administer.

Figure N M R. 7 : 1H MRI of a human head showing the soft tissue such as the brain and sinuses. The MRI also clearly shows
the spinal column and skull.
In many laboratories today, chemists use nuclear magnetic resonance to determine structures of important chemical and
biological compounds. In NMR spectra, different peaks give information about different atoms in a molecule according
specific chemical environments and bonding between atoms. The most common isotopes used to detect NMR signals are 1H
and 13C but there are many others, such as 2H, 3He, 15N, 19F, etc., that are also in use.
NMR has also proven to be very useful in other area such as environmental testing, petroleum industry, process control, earth’s
field NMR and magnetometers. Non-destructive testing saves a lot of money for expensive biological samples and can be used
again if more trials need to be run. The petroleum industry uses NMR equipment to measure porosity of different rocks and
permeability of different underground fluids. Magnetometers are used to measure the various magnetic fields that are relevant
to one’s study.

Problems
1. Calculate the magnetic field, B0 that corresponds to a precession frequency of 600 MHz for 1H.
2. What is the field strength (in tesla) needed to generate a 1H frequency of 500 MHz?
3. How do spin-spin relaxation and spin-lattice relaxation differ from each other?

7/4/2021 NMR.4 https://chem.libretexts.org/@go/page/1834


4. The 1H NMR spectrum of toluene shows that it has two peaks because of methyl and aromatic protons recorded at 60 MHz
and 1.41 T. Given this information, what would be the magnetic field at 400 MHz?
5. What is the difference between 13C and 1H NMR?

Solutions
1. B0= 14.1 T.
2. Using the equation used in problem 1 and solving it for B0we get a field strength of 11.74 T.
3. Look under relaxation.
4. Since we know that the NMR frequency is directly proportional to the magnetic strength, we calculate the magnetic field at
400 MHz: B0 = (400 MHz/60MHz) x 1.41 T = 9.40 T
5. Look under applications.

References
Atta-ur-Rahman. Nuclear Magnetic Resonance. New York: Springer-Verlag, 1986.
Freeman, Ray. Magnetic Resonance in Chemistry and Medicine. New York: Oxford University Press, 2003.
Lambert, Joseph B and Eugene P Mazzola. Nuclear Magnetic Resonance Spectroscopy: An Introduction to Princliples,
Applications, and Experimental Methods. Upper Saddle River: Pearson Education, 2004.
Chang, Raymond. Physical Chemistry for the Biosciences. University Science Books, 2005

Contributors and Attributions


Derrick Kaseman (UC Davis) and Revathi Srinivasan Ganesh Iyer (UCD)

7/4/2021 NMR.5 https://chem.libretexts.org/@go/page/1834


CHAPTER OVERVIEW
ROTATIONAL SPECTROSCOPY
Rotational spectroscopy is concerned with the measurement of the energies of transitions between
quantized rotational states of molecules in the gas phase. The spectra of polar molecules can be
measured in absorption or emission by microwave spectroscopy or by far infrared spectroscopy.

MICROWAVE ROTATIONAL SPECTROSCOPY


Microwave rotational spectroscopy uses microwave radiation to measure the energies of rotational
transitions for molecules in the gas phase. It accomplishes this through the interaction of the
electric dipole moment of the molecules with the electromagnetic field of the exciting microwave
photon.

ROTATIONAL SPECTROSCOPY OF DIATOMIC MOLECULES


The rotation of a diatomic molecule can be described by the rigid rotor model. To imagine this model think of a spinning dumbbell.
The dumbbell has two masses set at a fixed distance from one another and spins around its center of mass (COM). This model can be
further simplified using the concept of reduced mass which allows the problem to be treated as a single body system.

ROTATION OF LINEAR MOLECULES


The rotational energy levels of a diatomic molecule in 3D space is given by the quantum mechanical solution to the rotating rigid
rotor.

ROVIBRATIONAL SPECTROSCOPY
In this section, we will learn how the rotational transitions of molecules can accompany the vibrational transitions. It is important to
know how each peak correlates to the molecular processes of molecules. Rovibrational spectra can be analyzed to determine average
bond length.

1 7/7/2021
Microwave Rotational Spectroscopy
Microwave rotational spectroscopy uses microwave radiation to measure the energies of rotational transitions for molecules in
the gas phase. It accomplishes this through the interaction of the electric dipole moment of the molecules with the
electromagnetic field of the exciting microwave photon.

Introduction
To probe the pure rotational transitions for molecules, scientists use microwave rotational spectroscopy. This spectroscopy
utilizes photons in the microwave range to cause transitions between the quantum rotational energy levels of a gas molecule.
The reason why the sample must be in the gas phase is due to intermolecular interactions hindering rotations in the liquid and
solid phases of the molecule. For microwave spectroscopy, molecules can be broken down into 5 categories based on their
shape and the inertia around their 3 orthogonal rotational axes. These 5 categories include diatomic molecules, linear
molecules, spherical tops, symmetric tops and asymmetric tops.

Classical Mechanics
The Hamiltonian solution to the rigid rotor is

H =T (1)

since,
H = T +V (2)

Where T is kinetic energy and V is potential energy. Potential energy, V , is 0 because there is no resistance to the rotation
(similar to a particle in a box model).
Since H = T , we can also say that:
1
2
T = ∑ mi v (3)
i
2

However, we have to determine v in terms of rotation since we are dealing with rotation. Since,
i

v
ω = (4)
r

where ω = angular velocity, we can say that:

vi = ωX ri (5)

Thus we can rewrite the T equation as:


1
T = ∑ mi vi (ωX ri ) (6)
2

Since ω is a scalar constant, we can rewrite the T equation as:


ω ω L
T = ∑ mi (vi X ri ) = ∑ li = ω (7)
2 2 2

where l is the angular momentum of the ith particle, and L is the angular momentum of the entire system. Also, we know
i

from physics that,


L = Iω (8)

where I is the moment of inertia of the rigid body relative to the axis of rotation. We can rewrite the T equation as,
Iω 1 2
T =ω = Iω (9)
2 2

Quantum Mechanics
The internal Hamiltonian, H, is:

6/26/2021 1 https://chem.libretexts.org/@go/page/1839
2 2
i ℏ
H = (10)
2I

and the Schrödinger Equation for rigid rotor is:


2 2
i ℏ
ψ = Eψ (11)
2I

Thus, we get:
2
J(J + 1)h
En = (12)
2
8π I

where J is a rotational quantum number and ℏ is the reduced Planck's constant. However, if we let:
h
B = (13)
2
8π I

where B is a rotational constant, then we can substitute it into the E equation and get:
n

En = J(J + 1)Bh (14)

Considering the transition energy between two energy levels, the difference is a multiple of 2. That is, from J = 0 to J =1 ,
the ΔE 0→1is 2Bh and from J = 1 to J = 2, the ΔE is 4Bh.
1→2

Figure 1: Energy levels and line positions calculated in the rigid rotor approximation. This diagram illustrates how transitions
between the rotational energy levels of molecules map onto the energies at which these transitions are observed during
laboratory experiments. (CC CS-BY 3.0; Nnrw).

Theory
When a gas molecule is irradiated with microwave radiation, a photon can be absorbed through the interaction of the photon’s
electronic field with the electrons in the molecules. For the microwave region this energy absorption is in the range needed to
cause transitions between rotational states of the molecule. However, only molecules with a permanent dipole that changes
upon rotation can be investigated using microwave spectroscopy. This is due to the fact that their must be a charge difference
across the molecule for the oscillating electric field of the photon to impart a torque upon the molecule around an axis that is
perpendicular to this dipole and that passes through the molecules center of mass.
This interaction can be expressed by the transition dipole moment for the transition between two rotational states

Probability of Transition = ∫ ψrot (F )μ


^ψrot (I )dτ (15)

6/26/2021 2 https://chem.libretexts.org/@go/page/1839
Where Ψrot(F) is the complex conjugate of the wave function for the final rotational state, Ψrot(I) is the wave function of the
initial rotational state , and μ is the dipole moment operator with Cartesian coordinates of μx, μy, μz. For this integral to be
nonzero the integrand must be an even function. This is due to the fact that any odd function integrated from negative infinity
to positive infinity, or any other symmetric limits, is always zero.
In addition to the constraints imposed by the transition moment integral, transitions between rotational states are also limited
by the nature of the photon itself. A photon contains one unit of angular momentum, so when it interacts with a molecule it can
only impart one unit of angular momentum to the molecule. This leads to the selection rule that a transition can only occur
between rotational energy levels that are only one quantum rotation level (J) away from another1.

ΔJ = ±1 (16)

The transition moment integral and the selection rule for rotational transitions tell if a transition from one rotational state to
another is allowed. However, what these do not take into account is whether or not the state being transitioned from is actually
populated, meaning that the molecule is in that energy state. This leads to the concept of the Boltzmann distribution of states.
The Boltzmann distribution is a statistical distribution of energy states for an ensemble of molecules based on the temperature
of the sample2.
(−Erot (J)/RT )
nJ e
= (17)
J=n
n0 ∑ e
(−Erot (J)/RT )
J=1

where Erot(J) is the molar energy of the J rotational energy state of the molecule,
R is the gas constant,
T is the temperature of the sample.
n(J) is the number of molecules in the J rotational level, and
n0 is the total number of molecules in the sample.
This distribution of energy states is the main contributing factor for the observed absorption intensity distributions seen in the
microwave spectrum. This distribution makes it so that the absorption peaks that correspond to the transition from the energy
state with the largest population based on the Boltzmann equation will have the largest absorption peak, with the peaks on
either side steadily decreasing.

Degrees of Freedom
A molecule can have three types of degrees of freedom and a total of 3N degrees of freedom, where N equals the number of
atoms in the molecule. These degrees of freedom can be broken down into 3 categories3.
Translational: These are the simplest of the degrees of freedom. These entail the movement of the entire molecule’s center
of mass. This movement can be completely described by three orthogonal vectors and thus contains 3 degrees of freedom.
Rotational: These are rotations around the center of mass of the molecule and like the translational movement they can be
completely described by three orthogonal vectors. This again means that this category contains only 3 degrees of freedom.
However, in the case of a linear molecule only two degrees of freedom are present due to the rotation along the bonds in
the molecule having a negligible inertia.
Vibrational: These are any other types of movement not assigned to rotational or translational movement and thus there
are 3N – 6 degrees of vibrational freedom for a nonlinear molecule and 3N – 5 for a linear molecule. These vibrations
include bending, stretching, wagging and many other aptly named internal movements of a molecule. These various
vibrations arise due to the numerous combinations of different stretches, contractions, and bends that can occur between
the bonds of atoms in the molecule.
Each of these degrees of freedom is able to store energy. However, In the case of rotational and vibrational degrees of freedom,
energy can only be stored in discrete amounts. This is due to the quantized break down of energy levels in a molecule
described by quantum mechanics. In the case of rotations the energy stored is dependent on the rotational inertia of the gas
along with the corresponding quantum number describing the energy level.

Rotational Symmetries
To analyze molecules for rotational spectroscopy, we can break molecules down into 5 categories based on their shapes and
their moments of inertia around their 3 orthogonal rotational axes:4

6/26/2021 3 https://chem.libretexts.org/@go/page/1839
1. Diatomic Molecules
2. Linear Molecules
3. Spherical Tops
4. Symmetrical Tops
5. Asymmetrical Tops

Diatomic Molecules
The rotations of a diatomic molecule can be modeled as a rigid rotor. This rigid rotor model has two masses attached to each
other with a fixed distance between the two masses.

It has an inertia (I) that is equal to the square of the fixed distance between the two masses multiplied by the reduced mass of
the rigid rotor.
2
Ie = μ re (18)

m1 m2
μ = (19)
m1 + m2

Using quantum mechanical calculations it can be shown that the energy levels of the rigid rotator depend on the inertia of the
rigid rotator and the quantum rotational number J2.
E(J) = Be J(J + 1) (20)

h
Be = (21)
2
8 π c Ie

However, this rigid rotor model fails to take into account that bonds do not act like a rod with a fixed distance, but like a
spring. This means that as the angular velocity of the molecule increases so does the distance between the atoms. This leads us
to the nonrigid rotor model in which a centrifugal distortion term (D ) is added to the energy equation to account for this
e

stretching during rotation.


−1 2 2
E(J)(c m ) = Be J(J + 1)– De J (J + 1 ) (22)

This means that for a diatomic molecule the transitional energy between two rotational states equals
′ ′ ′′ ′′ ′2 ′ 2 ′′2 ′ 2
E = Be [ J (J + 1) − J (J + 1)] − De [ J (J + 1) −J (J + 1) ] (23)

Where J’ is the quantum number of the final rotational energy state and J’’ is the quantum number of the initial rotational
energy state. Using the selection rule of ΔJ = ±1 the spacing between peaks in the microwave absorption spectrum of a
diatomic molecule will equal
′′ ′′3
ER = (2 Be − 4 De ) + (2 Be − 12 De )J − 4 De J (24)

Linear Molecules
Linear molecules behave in the same way as diatomic molecules when it comes to rotations. For this reason they can be
modeled as a non-rigid rotor just like diatomic molecules. This means that linear molecule have the same equation for their
rotational energy levels. The only difference is there are now more masses along the rotor. This means that the inertia is now
the sum of the distance between each mass and the center of mass of the rotor multiplied by the square of the distance between
them2.

6/26/2021 4 https://chem.libretexts.org/@go/page/1839
n

2
Ie = ∑ mj r (25)
ej

j=1

Where mj is the mass of the jth mass on the rotor and rej is the equilibrium distance between the jth mass and the center of mass
of the rotor.

Spherical Tops
Spherical tops are molecules in which all three orthogonal rotations have equal inertia and they are highly symmetrical. This
means that the molecule has no dipole and for this reason spherical tops do not give a microwave rotational spectrum.

Figure 2: Geometrical example of a spherical top


Examples:

Symmetrical Tops
Symmetrical tops are molecules with two rotational axes that have the same inertia and one unique rotational axis with a
different inertia. Symmetrical tops can be divided into two categories based on the relationship between the inertia of the
unique axis and the inertia of the two axes with equivalent inertia. If the unique rotational axis has a greater inertia than the
degenerate axes the molecule is called an oblate symmetrical top. If the unique rotational axis has a lower inertia than the
degenerate axes the molecule is called a prolate symmetrical top. For simplification think of these two categories as either
frisbees for oblate tops or footballs for prolate tops.

Figure 3 : Symmetric Tops: (Left) Geometrical example of an oblate top and (right) a prolate top. Images used with
permission from Wikipedia.com.

6/26/2021 5 https://chem.libretexts.org/@go/page/1839
Figure 4: Examples of symmetric tops. Benzene (oblate) XeF4 (oblate) ClCH3 (prolate) NH3 (prolate)
In the case of linear molecules there is one degenerate rotational axis which in turn has a single rotational constant. With
symmetrical tops now there is one unique axis and two degenerate axes. This means an additional rotational constant is needed
to describe the energy levels of a symmetrical top. In addition to the rotational constant an additional quantum number must be
introduced to describe the rotational energy levels of the symmetric top. These two additions give us the following rotational
energy levels of a prolate and oblate symmetric top
−1 2
E(J,K) (c m ) = Be ∗ J(J + 1) + (Ae − Be )K (26)

Where Be is the rotational constant of the unique axis, Ae is the rotational constant of the degenerate axes, J is the total
rotational angular momentum quantum number and K is the quantum number that represents the portion of the total angular
momentum that lies along the unique rotational axis. This leads to the property that K is always equal to or less than J . Thus
we get the two selection rules for symmetric tops
ΔJ = 0, ±1 (27)

ΔK = 0 (28)

when K ≠ 0
ΔJ = ±1 (29)

ΔK = 0 (30)

when K = 0
However, like the rigid rotor approximation for linear molecules, we must also take into account the elasticity of the bonds in
symmetric tops. Therefore, in a similar manner to the rigid rotor we add a centrifugal coupling term, but this time we have one
for each quantum number and one for the coupling between the two.
−1 2 2 2
E(J,K) (c m ) = Be J(J + 1) − DeJ J (J + 1 ) + (Ae − Be ) ∗ K (31)

4 2
−Dek K − Dejk J(J + 1)K (32)

Asymmetrical Tops
Asymmetrical tops have three orthogonal rotational axes that all have different moments of inertia and most molecules fall into
this category. Unlike linear molecules and symmetric tops these types of molecules do not have a simplified energy equation to
determine the energy levels of the rotations. These types of molecules do not follow a specific pattern and usually have very
complex microwave spectra.

Figure 5: Molecular examples of asymmetrical tops. (left) water and (right) acetone

Additional Rotationally Sensitive Spectroscopies


6/26/2021 6 https://chem.libretexts.org/@go/page/1839
In addition to microwave spectroscopy, IR spectroscopy can also be used to probe rotational transitions in a molecule.
However, in the case of IR spectroscopy the rotational transitions are coupled to the vibrational transitions of the molecule.
One other spectroscopy that can probe the rotational transitions in a molecule is Raman spectroscopy, which uses UV-visible
light scattering to determine energy levels in a molecule. However, a very high sensitivity detector must be used to analyze
rotational energy levels of a molecule.

References
1. Harris, D. C.; Bertolucci, M. D., Symmetry and Spectroscopy. University Press: Oxford, 1978.
2. McQuarrie, D. A.; Simon, J. D., Physical Chemistry: A Molecular Approach. University Science Books: 1997.
3. Shoemaker, D. P.; W., G. C.; W., N. J., Experiments in Physical Chemistry. 8th ed.; McGraw Hill: New York, 2009.
4. Hollas, M. J., Basic Atomic and Molecular Spectroscopy. Royal Society of Chemistry: Cambridge, 2002.

Contributors and Attributions


Nicholas Houghton, UC Davis

6/26/2021 7 https://chem.libretexts.org/@go/page/1839
Rotational Spectroscopy of Diatomic Molecules
The rotation of a diatomic molecule can be described by the rigid rotor model. To imagine this model think of a spinning
dumbbell. The dumbbell has two masses set at a fixed distance from one another and spins around its center of mass (COM).
This model can be further simplified using the concept of reduced mass which allows the problem to be treated as a single
body system.

Introduction
Similar to most quantum mechanical systems our model can be completely described by its wave function. Therefore, when
we attempt to solve for the energy we are lead to the Schrödinger Equation. In the context of the rigid rotor where there is a
natural center (rotation around the COM) the wave functions are best described in spherical coordinates. In addition to having
pure rotational spectra diatomic molecules have rotational spectra associated with their vibrational spectra. The order of
magnitude differs greatly between the two with the rotational transitions having energy proportional to 1-10 cm-1 (microwave
radiation) and the vibrational transitions having energy proportional to 100-3,000 cm-1 (infrared radiation). Rotational
spectroscopy is therefore referred to as microwave spectroscopy.

Rigid Rotor Model


A diatomic molecule consists of two masses bound together. The distance between the masses, or the bond length, (l) can be
considered fixed because the level of vibration in the bond is small compared to the bond length. As the molecule rotates it
does so around its COM (observed in Figure 1:. as the intersection of R and R ) with a frequency of rotation of ν given in
1 2 rot

radians per second.

Figure 1: Rigid Rotor Model of a Diatomic Molecule

Reduced Mass
The system can be simplified using the concept of reduced mass which allows it to be treated as one rotating body. The system
can be entirely described by the fixed distance between the two masses instead of their individual radii of rotation.
Relationships between the radii of rotation and bond length are derived from the COM given by:
M1 R1 = M2 R2 , (1)

where l is the sum of the two radii of rotation:


l = R1 + R2 . (2)

Through simple algebra both radii can be found in terms of their masses and bond length:
M2
R1 = l (3)
M1 + M2

and
M1
R2 = l. (4)
M1 + M2

The kinetic energy of the system, T , is sum of the kinetic energy for each mass:

6/22/2021 1 https://chem.libretexts.org/@go/page/1840
2 2
M1 v + M2 v
1 2
T = , (5)
2

where

v1 = 2π R1 νrot (6)

and

v2 = 2π R2 νrot . (7)

Using the angular velocity,

ω = 2πνrot (8)

the kinetic energy can now be written as:


2 2
M1 R + M2 R
1 2
T = ω. (9)
2

With the moment of inertia,


2 2
I = M1 R + M2 R , (10)
1 2

the kinetic energy can be further simplified:


2

T = . (11)
2

The moment of inertia can be rewritten by plugging in for R and R :


1 2

M1 M2 2
I = l , (12)
M1 + M2

where
M1 M2
(13)
M1 + M2

is the reduced mass, μ . The moment of inertia and the system are now solely defined by a single mass, μ , and a single length,
l:

2
I = μl . (14)

Angular Momentum
Another important concept when dealing with rotating systems is the the angular momentum defined by: L = I ω
Looking back at the kinetic energy:
2 2 2 2
Iω I ω L
T = = = (15)
2 2I 2I

The angular momentum can now be described in terms of the moment of inertia and kinetic energy: L 2
= 2I T .

Setting up the Schrödinger Equation


The wave functions for the rigid rotor model are found from solving the time-independent Schrödinger Equation:
^
H ψ = Eψ (16)

Where the Hamiltonian Operator is:


−ℏ
^ 2
H = ∇ + V (r) (17)

where ∇ is the Laplacian Operator and can be expressed in either Cartesian coordinates:
2

6/22/2021 2 https://chem.libretexts.org/@go/page/1840
2 2 2
2
∂ ∂ ∂
∇ = + + (18)
2 2 2
∂x ∂y ∂z

or in spherical coordinates:
2
2
1 ∂ 2
∂ 1 ∂ ∂ 1 ∂
∇ = (r )+ (sin θ )+ (19)
2 2 2 2
r ∂r ∂r r sin θ ∂θ ∂θ r sin θ ∂ϕ

At this point it is important to incorporate two assumptions:



The distance between the two masses is fixed. This causes the terms in the Laplacian containing to be zero.
∂r
The orientation of the masses is completely described by θ and ϕ and in the absence of electric or magnetic fields the
energy is independent of orientation. This causes the potential energy portion of the Hamiltonian to be zero.
The wave functions ψ(θ, ϕ) are customarily represented by Y (θ, ϕ) and are called spherical harmonics.
The Hamiltonian Operator can now be written:
2 2
−ℏ 1 ∂ ∂ 1 ∂
^ ^ =
H =T [ (sin θ )+ ] (20)
2
2μl sin θ ∂θ ∂θ sin θ ∂ϕ2

with the Angular Momentum Operator being defined:


^ ^
L = 2I T (21)

2
1 ∂ ∂ 1 ∂
^ 2
L = −ℏ [ (sin θ )+ ] (22)
sin θ ∂θ ∂θ sin θ ∂ϕ2

The Schrödinger Equation now expressed:


2 2
−ℏ 1 ∂ ∂ 1 ∂
[ (sin θ )+ ] Y (θ, ϕ) = EY (θ, ϕ) (23)
2I sin θ ∂θ ∂θ sin θ ∂ϕ2

Solving the Schrödinger Equation


The Schrödinger Equation can be solved using separation of variables.
Step 1:
2I E
Let Y (θ, ϕ) = Θ (θ) Φ (ϕ) , and substitute: β = 2
.

Set the Schrödinger Equation equal to zero:


2
sin θ d dΘ 1 d Φ
2
(sin θ ) + β sin θ+ =0 (24)
2
Θ (θ) dθ dθ Φ (ϕ) dϕ

Step 2: Because the terms containing Θ (θ) are equal to the terms containing Φ (ϕ) they must equal the same constant in order
to be defined for all values:
sin θ d dΘ
2 2
(sin θ ) + β sin θ =m (25)
Θ (θ) dθ dθ

2
1 d Φ
2
= −m (26)
2
Φ (ϕ) dϕ

Step 3: Solving for Φ is fairly simple and yields:


1 imϕ
Φ (ϕ) = −−e (27)
√2π

where m = 0, ±1, ±2, . . .


Solving for θ is considerably more complicated but gives the quantized result:

6/22/2021 3 https://chem.libretexts.org/@go/page/1840
β = J(J + 1) (28)

where J is the rotational level with J = 0, 1, 2, . . .


Step 4: The energy is quantized by expressing in terms of β:
2
ℏ β
E = (29)
2I

2

Step 5: Using the rotational constant, B = , the energy is further simplified: E = BJ(J + 1)
2I

Energy of Rotational Transitions


When a molecule is irradiated with photons of light it may absorb the radiation and undergo an energy transition. The energy
of the transition must be equivalent to the energy of the photon of light absorbed given by: E = hν . For a diatomic molecule
the energy difference between rotational levels (J to J+1) is given by:
EJ+1 − EJ = B(J + 1)(J + 2) − BJ(J = 1) = 2B(J + 1) (30)

with J=0, 1, 2,...


Because the difference of energy between rotational levels is in the microwave region (1-10 cm-1) rotational spectroscopy is
commonly called microwave spectroscopy. In spectroscopy it is customary to represent energy in wave numbers (cm-1), in this
~
notation B is written as B . To convert from units of energy to wave numbers simply divide by h and c, where c is the speed of
~ h
light in cm/s (c=2.998e10 cm/s). In wave numbers B = .
8πcI

Figure 2:. Rotational Energy Levels


~
Figure 2: predicts the rotational spectra of a diatomic molecule to have several peaks spaced by 2B. This contrasts vibrational
spectra which have only one fundamental peak for each vibrational mode. From the rotational spectrum of a diatomic
~
molecule the bond length can be determined. Because B is a function of I and therefore a function of l (bond length), so l can
be readily solved for:
−−−− −− −
h
l =√ . (31)
~
2
8 π c Bμ

Selection rules only permit transitions between consecutive rotational levels: ΔJ = J ± 1 , and require the molecule to
contain a permanent dipole moment. Due to the dipole requirement, molecules such as HF and HCl have pure rotational
spectra and molecules such as H2 and N2 are rotationally inactive.

Centrifugal Distortion
As molecules are excited to higher rotational energies they spin at a faster rate. The faster rate of spin increases the centrifugal
force pushing outward on the molecules resulting in a longer average bond length. Looking back, B and l are inversely related.
Therefore the addition of centrifugal distortion at higher rotational levels decreases the spacing between rotational levels. The
correction for the centrifugal distortion may be found through perturbation theory:

6/22/2021 4 https://chem.libretexts.org/@go/page/1840
~ ~ 2 2
EJ = BJ(J + 1) − DJ (J + 1 ) . (32)

Rotation-Vibration Transitions
Rotational transitions are on the order of 1-10 cm-1, while vibrational transitions are on the order of 1000 cm-1. The difference
of magnitude between the energy transitions allow rotational levels to be superimposed within vibrational levels.

Figure 3: Rotation-Vibration Transitions


~ ~ ~
Combining the energy of the rotational levels, E = BJ(J + 1) , with the vibrational levels,
J
~
E v = w (v + 1/2) , yields the
total energy of the respective rotation-vibration levels:
~ ~ ~
E v,J = w (v + 1/2) + BJ(J + 1) (33)

Following the selection rule, ΔJ = J ± 1 , Figure 3. shows all of the allowed transitions for the first three rotational states,
where J" is the initial state and J' is the final state.
~ ~
When the ΔJ = +1 transitions are considered (blue transitions) the initial energy is given by: E 0,J
~
= w(1/2) + BJ(J + 1)
~ ~
and the final energy is given by: E = w(3/2) + B(J + 1)(J + 2) .
~
v,J+1

~ ~
The energy of the transition, ~
Δν = E 1,J+1 − E 0,J , is therefore:
~ ~
~
Δν = w + 2 B(J + 1) (34)

where J"=0, 1, 2,...


~ ~
When the ΔJ = −1 transitions are considered (red transitions) the initial energy is given by: ~
E v,J = w (1/2) + BJ(J + 1)

and the final energy is given by:


~ ~ ~
E v,J−1 = w (3/2) + B(J − 1)(J). (35)

~
The energy of the transition is therefore: ~ ~
Δν = w − 2 B(J) where J"=1, 2, 3,...
The difference in energy between the J+1 transitions and J-1 transitions causes splitting of vibrational spectra into two
branches. The J-1 transitions, shown by the red lines in Figure 3, are lower in energy than the pure vibrational transition and
form the P-branch. The J+1 transitions, shown by the blue lines in Figure 3. are higher in energy than the pure vibrational
transition and form the R-branch. Notice that because the ΔJ = ±0 transition is forbidden there is no spectral line associated
with the pure vibrational transition. Therefore there is a gap between the P-branch and R-branch, known as the q branch.

6/22/2021 5 https://chem.libretexts.org/@go/page/1840
Figure used with permission from Wikipedia.
In the high resolution HCl rotation-vibration spectrum the splitting of the P-branch and R-branch is clearly visible. Due to the
small spacing between rotational levels high resolution spectrophotometers are required to distinguish the rotational
transitions.

Rotation-Vibration Interactions
Recall the Rigid-Rotor assumption that the bond length between two atoms in a diatomic molecule is fixed. However, the
anharmonicity correction for the harmonic oscillator predicts the gaps between energy levels to decrease and the equilibrium
bond length to increase as higher vibrational levels are accessed. Due to the relationship between the rotational constant and
bond length:
~ h
B = (36)
2 2
8 π cμl

The rotational constant is dependent on the vibrational level:

~ ~ 1
~
Bv = B − α (v + ) (37)
2

Where α~
is the anharmonicity correction and v is the vibrational level. As a consequence the spacing between rotational levels
decreases at higher vibrational levels and unequal spacing between rotational levels in rotation-vibration spectra occurs.
Including the rotation-vibration interaction the spectra can be predicted.
For the R-branch
~ ~
E 1,J+1 − E 0,J (38)

3 ~ 1 ~
~ ~ ~
ν = [w ( ) + B1 (J + 1) (J + 2)] − [w ( ) + B0 J (J + 1)] (39)
2 2

~ ~ ~ ~ 2 ~ ~ ~
ν = w + (B1 − B0 ) J + (3 B1 − B0 ) J + 2 B1 (40)

where J=0, 1, 2,...


For the P-branch
~ ~
E 1,J−1 − E 0,J (41)

3 ~ 1 ~
~ ~ ~
ν = [w ( ) + B1 (J − 1) J] − [w ( ) + B0 J (J + 1)] (42)
2 2

~ ~ ~ ~ 2 ~ ~
ν = w + (B1 − B0 ) J − (B1 + B0 ) J (43)

where J=1, 2, 3,...


~ ~
Because B 1 < B0 , as J increases:

6/22/2021 6 https://chem.libretexts.org/@go/page/1840
Spacing in the R-branch decreases.
Spacing in the P-branch increases.

References
1. McQuarrie, Donald A. Quantum Chemistry. New York: University Science Books, 2007.

Problems
1. What is the potential energy of the Rigid-Rotor?
2. Derive the Schrodinger Equation for the Rigid-Rotor.
3. Researchers have been interested in knowing what Godzilla uses as the fuel source for his fire breathing. A recent
breakthrough was made and some residue containing Godzilla's non-combusted fuel was recovered. Studies on the residue
showed that the fuel, Compound G, is a diatomic molecule and has a reduced mass of 1.615x10-27 kg. In addition, a
microwave spectrum of Compound G was obtained and revealed equally spaced lines separated by 4.33 cm-1. Using the
Rigid-Rotor model determine the bond length of Compound G.
4. How would deuterium substitution effect the pure rotational spectrum of HCl?

Contributors and Attributions


Benjamin Strong (Hope)

6/22/2021 7 https://chem.libretexts.org/@go/page/1840
Rotation of Linear Molecules
The rotational energy levels of a diatomic molecule in 3D space is given by the quantum mechanical solution to the rotating
rigid rotor:
2

E = J(J + 1) (1)
2I

where
J is a rotational quantum number ranging from J = 0 to J = ∞ .
I is the moment of inertia and is the rotational equivalent to mass in translation.
The moment of inertia of a molecule is very sensitive to the geometry to the the molecule (see below).

Values of rotational inertia for common shapes of objects.


In general, any three-dimension species (e.g., a molecule) will have three degrees of rotational energy since it can rotation in
the x, y and z axis (i.e., the angular momentum vector can lie in each axis). This energy spacing for rotation in each degree is
given by Equation 1 and is shown geometrically below.

Energy spacing for a rigid rotor (in 3D) as a function of \(J) quantum number.
The lowest energy transition is the J = 0 to J = 1 transition and corresponds to
2 2
ℏ ℏ
EJ=0→J=1 = [1(1 + 1) − 0(0 + 1)] = (2)
2I I

5/28/2021 1 https://chem.libretexts.org/@go/page/77585
Most of the mass of the molecule is in the nuclei, so when calculating the moment of inertia I we can ignore the electrons and
just use the nuclei. But the size of the nuclei is around 10 times smaller than the bond length. This means the moment of
−5

inertia around an axis along the bond is going to be about 10 smaller than the moment of inertia around an axis normal to the
10

bond. Therefore the energy level spacings will be around 10 times bigger along the bond than normal to it.
10

It is common to argue that linear molecules do not rotate perpendicular to the axis of symmetry and often justified in terms of
symmetry of the molecule (see Group Theory). Therefore, out of three possible rotational degrees of freedom for a three
dimensional object, only two are applicable to linear molecules and the third rotation is often ignored. This is calipalized in
terms of justifying how many vibrational degrees of freedom a molecule has:
Non-linear molecules have 3N degrees of freedom in total: 3 are translational and 3 are rotational (all are allowed for non-
linear molecules) so the remaining 3N-6 are vibrational.
In contrast, linear molecules have 3 translational and only 2 rotational, and to keep a total of 3N degrees of freedom, they
have 3N-5 vibrational degrees.
These equations are justified since rotation around the axis along the bond of the molecule requires huge energies (Equation 2)
since I due to the much smaller moment of inertia. To excited these rotations, gamma ray photons are required and is the topic
of high energy physics. Moreover, the temperature at which energy levels above the ground level are important is extremely
high (e.g., it can be higher than the temperature of dissociation of the molecule). Therefore, at temperatures of practical
interest, rotation around the axis of the linear molecule is not important for thermodynamic properties.

Contributors and Attributions


John Rennie, akhmeteli (physics StackExchange)

5/28/2021 2 https://chem.libretexts.org/@go/page/77585
Rovibrational Spectroscopy
In this section, we will learn how the rotational transitions of molecules can accompany the vibrational transitions. It is
important to know how each peak correlates to the molecular processes of molecules. Rovibrational spectra can be analyzed to
determine the average bond length.

Introduction
Each of the normal modes of vibration of heteronuclear diatomic molecules in the gas phase also contains closely-spaced (1-10
cm-1 difference) energy states attributable to rotational transitions that accompany the vibrational transitions. A molecule’s
rotation can be affected by its vibrational transition because there is a change in bond length, so these rotational transitions are
expected to occur. Since vibrational energy states are on the order of 1000 cm-1, the rotational energy states can be
superimposed upon the vibrational energy states.

Selection Rules
Rotational and Vibration transitions (also known as rigid rotor and harmonic oscillator) of molecules help us identify how
molecules interact with each other, their bond length as mentioned in the previous section. In order to know each transition, we
have to consider other terms like wavenumber, force constant, quantum number, etc. There are rotational energy levels
associated with all vibrational levels. From this, vibrational transitions can couple with rotational transitions to give
rovibrational spectra. Rovibrational spectra can be analyzed to determine the average bond length.
We treat the molecule's vibrations as those of a harmonic oscillator (ignoring anharmonicity). The energy of a vibration is
quantized in discrete levels and given by
1
Ev = hν (v + )
2

Where v is the vibrational quantum number and can have integer values 0, 1, 2..., and ν is the frequency of the vibration given
by:


1 k
ν = √
2π μ

Where k is the force constant and μ is the reduced mass of a diatomic molecule with atom masses m1 and m2, given by
m1 m2
μ =
m1 + m2

We treat the molecule's rotations as those of a rigid rotor (ignoring centrifugal distortion). The energy of a rotation is also
quantized in discrete levels given by
2
h
Er = J(J + 1)
8π 2 I

In which I is the moment of inertia, given by


2
I = μr

where μ is the reduced mass from above and r is the equilibrium bond length.
Experimentally, frequencies or wavenumbers are measured rather than energies, and dividing by h or hc gives more commonly
seen term symbols, F(J) using the rotational quantum number J and the rotational constant B in either frequency
Er h
F (J) = = J(J + 1) = BJ(J + 1)
2
h 8π I

or wavenumbers
Er h
F (J) = = J(J + 1) = BJ(J + 1)
2
hc 8 π cI

7/1/2021 1 CC-BY https://chem.libretexts.org/@go/page/1841


It is important to note in which units one is working since the rotational constant is always represented as B, whether in
frequency or wavenumbers.

Vibrational Transition Selection Rules


At room temperature, typically only the lowest energy vibrational state v= 0 is populated, so typically v0 = 0 and ∆v = +1. The
full selection rule is technically that ∆v = ±1, however here we assume energy can only go upwards because of the lack of
population in the upper vibrational states.

Rotational Transition Selection Rules


At room temperature, states with J≠0 can be populated since they represent the fine structure of vibrational states and have
smaller energy differences than successive vibrational levels. Additionally, ∆J = ±1 since a photon contains one quantum of
angular momentum and we abide by the principle of conservation of energy. This is also the selection rule for rotational
transitions.
The transition ∆J = 0 (i.e. J" = 0 and J' = 0), but where v0 = 0 and ∆v = +1, is forbidden and the pure vibrational transition is
not observed in most cases. The rotational selection rule gives rise to an R-branch (when ∆J = +1) and a P-branch (when ∆J =
-1). Each line of the branch is labeled R(J) or P(J), where J represents the value of the lower state.

R-branch
When ΔJ = +1 , i.e. the rotational quantum number in the ground state is one more than the rotational quantum number in the
excited state – R branch (in French, riche or rich). To find the energy of a line of the R-branch:
′ ′
ΔE = h ν0 + hB [J(J + 1) − J (J +1)]

= h ν0 + hB [(J + 1)(J + 2) − J(J + 1)]

= h ν0 + 2hB(J + 1)

P-branch
When ΔJ = −1 , i.e. the rotational quantum number in the ground state is one less than the rotational quantum number in the
excited state – P branch (in French, pauvre or poor). To find the energy of a line of the P-branch:
′ ′
ΔE = h ν0 + hB [J(J + 1) − J (J + 1)]

= h ν0 + hB [J(J − 1) − J(J + 1)]

= h ν0 − 2hBJ

Q-branch
When ΔJ = 0 , i.e. the rotational quantum number in the ground state is the same as the rotational quantum number in the
excited state – Q branch (simple, the letter between P and R). To find the energy of a line of the Q-branch:
′ ′
ΔE = h ν0 + hB[J(J + 1) − J (J + 1)]

= hν0

The Q-branch can be observed in polyatomic molecules and diatomic molecules with electronic angular momentum in the
ground electronic state, e.g. nitric oxide, NO. Most diatomics, such as O2, have a small moment of inertia and thus very small
angular momentum and yield no Q-branch.

7/1/2021 2 CC-BY https://chem.libretexts.org/@go/page/1841


Figure 1: Cartoon depiction of rotational energy levels, J, imposed on vibrational energy levels, v. The transitions between
levels that would result in the P- and R-branches are depicted in purple and red, respectively, in addition to the theoretical Q-
branch line in blue.
As seen in Figure 1, the lines of the P-branch (represented by purple arrows) and R-branch (represented by red arrows) are
separated by specific multiples of B (2B), thus the bond length can be deduced without the need for pure rotational
spectroscopy.

Energy
The total nuclear energy of the combined rotation-vibration terms, S(v, J) , can be written as the sum of the vibrational energy
and the rotational energy

S(v, J) = G(v) + F (J)

Where G(v) represents the energy of the harmonic oscillator, ignoring anharmonic components and S(J) represents the
energy of a rigid rotor, ignoring centrifugal distortion. From this, we can derive
1
S(v, J) = ν0 v + + BJ(J + 1)
2

The relative intensity of the P- and R-branch lines depends on the thermal distribution of electrons; more specifically, they
depend on the population of the lower J state. If we represent the population of the Jth upper level as NJ and the population of
the lower state as N0, we can find the population of the upper state relative to the lower state using the Boltzmann distribution:
NJ −Er /kT
= (2J + 1)e
N0

(2J+1) gives the degeneracy of the Jth upper level arising from the allowed values of M (+J to –J). As J increases, the
J

degeneracy factor increases and the exponential factor decreases until at high J, the exponential factor wins out and NJ/N0
approaches zero at a certain level, Jmax. Thus, when
d NJ
( ) =0
dJ N0

by differentiation, we obtain

7/1/2021 3 CC-BY https://chem.libretexts.org/@go/page/1841


1

kT 2
1
Jmax = ( ) −
2hcB 2

This is the reason that rovibrational spectral lines increase in energy to a maximum as J increases, then decrease to zero as J
continues to increase, as seen in Figure 2 and Figure 3.
From this relationship, we can also deduce that in heavier molecules, B will decrease because the moment of inertia will
increase, and the decrease in the exponential factor is less pronounced. This results in the population distribution shifting to
higher values of J. Similarly, as temperature increases, the population distribution will shift towards higher values of J.

Ideal Spectrum
The spectrum we expect, based on the conditions described above, consists of lines equidistant in energy from one another,
separated by a value of 2B. The relative intensity of the lines is a function of the rotational populations of the ground states,
i.e. the intensity is proportional to the number of molecules that have made the transition. The overall intensity of the lines
depends on the vibrational transition dipole moment.

Figure 2: A cartoon depiction of an ideal rovibrational spectrum.


Between P(1) and R(0) lies the zero gap, where the the first lines of both the P- and R-branch are separated by 4B, assuming
that the rotational constant B is equal for both energy levels. The zero gap is also where we would expect the Q-branch,
depicted as the dotted line, if it is allowed.

Real Spectra
We find that real spectra do not exactly fit the expectations from above.

Figure 3: A cartoon depiction of a real rovibrational spectrum.


As energy increases, the R-branch lines become increasingly similar in energy (i.e., the lines move closer together) and as
energy decreases, the P-branch lines become increasingly dissimilar in energy (i.e. the lines move farther apart). This is
attributable to two phenomena: rotational-vibrational coupling and centrifugal distortion.

Rotational-Vibrational Coupling

7/1/2021 4 CC-BY https://chem.libretexts.org/@go/page/1841


As a diatomic molecule vibrates, its bond length changes. Since the moment of inertia is dependent on the bond length, it too
changes and, in turn, changes the rotational constant B. We assumed above that B of R(0) and B of P(1) were equal, however
they differ because of this phenomenon and B is given by
1
Be = (−αe ν + )
2

Where B is the rotational constant for a rigid rotor and α is the rotational-vibrational coupling constant. The information in
e e

the band can be used to determine B0 and B1 of the two different energy states as well as the rotational-vibrational coupling
constant, which can be found by the method of combination differences.
Combination Differences
Combination differences involves finding the values of B0 and B1 rotational-vibrational coupling constant by measuring the
change for two different transitions sharing a common state.

Figure 4: Part of the method of combination differences depicting P- and R-branch transitions sharing a common lower J state.
To determine B1, we pair transitions sharing a common lower state; here, R(1) and P(1). Note that the vibrational level does
not change. Both branches begin with J = 1, so by finding the difference in energy between the lines, we find B1.
′ ′
ΔER − ΔEP = E(ν = 1, J = J + 1) − E(ν = 1, J = J − 1)

Inserting this information into the equation from above, we obtain


~ ~
= ν [R(J − 1)] − ν [P (J + 1)]

= ω0 + B1 (J + 1)(J + 2) − B0 J(J + 1) − ω0 − B1 (J − 1)J + B0 J(J + 1)

1
= 4B1 (J + )
2

1
If we plot ΔE R − ΔEP against J + , we obtain a straight line with slope 4B1.
2

Figure 5: Part of the method of combination differences depicting P- and R-branch transitions sharing a common upper J state.
Similarly, we can determine B0 by finding wavenumber differences in transitions sharing a common upper state; here, R(0)
and P(2). Both branches terminate at J=1 and differences will only depend on B0.
~ ~
= ν [R(J − 1)] − ν [P (J + 1)]

= ω0 + B1 J(J + 1) − B0 J(J − 1) − ω0 − B1 J(J + 1) + B0 (J + 1)(J + 2)

1
= 4B0 (J+ )
2

7/1/2021 5 CC-BY https://chem.libretexts.org/@go/page/1841


1
As before, if we plot ΔE R − ΔEP vs. (J+ ) , we obtain a straight line with slope 4B0.
2

Following from this, we can obtain the rotational-vibrational coupling constant:

= B1 − B0 = α

Centrifugal Distortion
Similarly to rotational-vibrational coupling, centrifugal distortion is related to the changing bond length of a molecule. A real
molecule does not behave as a rigid rotor that has a rigid rod for a chemical bond, but rather acts as if it has a spring for a
chemical bond. As the rotational velocity of a molecule increases, its bond length increases and its moment of inertia
increases. As the moment of inertia increases, the rotational constant B decreases.
2 2
F (J) = BJ(J + 1) − DJ (J + 1)

Where D is the centrifugal distortion constant and is related to the vibration wavenumber, ω
3
4B
D =
2
ω

When the above factors are accounted for, the actual energy of a rovibrational state is
1 1
2
S(v, J) = ν0 v + + Be J(J + 1) − αe (v + ) J(J + 1) − De [J(J + 1)]
2 2

References
1. Hollas, M. J. Modern spectroscopy. (3rd ed.). Chichester: John Wiley & Sons, 1996.
2. Hollas, M. J. Basic atomic and molecular spectroscopy. Cambridge: The Royal Society of Chemistry, 2002.
3. Herzberg, G. Molecular spectra and molecular structure. (2nd ed.). New York: Prentice-Hall, 1950.
4. Fetterolf, Monty L. Enhanced Intensity Distribution Analysis of the Rotational–Vibrational Spectrum of HCl. J. Chem. Ed.
2007, 84, 1064. DOI: 10.1021/ed084p1062

Problems
Find the reduced mass of D35Cl in kg, if the mass of D-2 is 2.014 amu and the mass of Cl-35 is 34.968 amu.

Answer
2.014amu ∗ 34.968amu
gives 1.807 amu. To convert to kg, multiple by 1.66 x 10-27 kg/amu. Answer: 3.00 x 10-27 kg
2.014amu + 34.968amu

Using information found in problem 1, calculate the rotational constant B (in wavenumbers) of D35Cl given that the average
bond length is 1.2745 Å.

Answer
h
We know that in wavenumbers, B = 2
.
8 π cI

First, we must solve for the moment of inertia, I, using


2 −27 −10 2
I = μr = (3.00 ∗ 10 kg)(1.2745 ∗ 10 m)

= 4.87 x 10-47 kg•m2= I


We can now substitute into the original formula to solve for B. h is Planck's constant, c is the speed of light in m/s and I =
4.87 x 10-47 kg•m2. This will give us the answer in m-1, then we can convert to cm-1. Answer: 5.74 cm-1.

Using the rigid rotor approximation, estimate the bond length in a 12C16O molecule if the energy difference between J=1 and
J=3 were to equal 14,234 cm-1.

Answer

7/1/2021 6 CC-BY https://chem.libretexts.org/@go/page/1841


We use the same formula as above and expand the moment of inertia in order to solve for the average bond length.
h
B =
2 2
8 π cμr

We can deduce the rotational constant B since we know the distance between two energy states and the relationship

F (J) = BJ(J + 1)

The distance between J=1 and J=3 is 10B, so using the fact that B = 14,234 cm-1, B=1423.4 cm-1. We convert this to m-1 so
that it will match up with the units of the speed of light (m/s) and obtain B = 142340 m-1.
Using the reduced mass formula, we find that µ = 1.138 x 10-26 kg. Answer: r = 81 Å.

Contributors and Attributions


Joya Cooley (UC Davis)

7/1/2021 7 CC-BY https://chem.libretexts.org/@go/page/1841


SECTION OVERVIEW
VIBRATIONAL SPECTROSCOPY
Infrared spectroscopy (IR spectroscopy or Vibrational Spectroscopy) is the spectroscopy that deals
with the infrared region of the electromagnetic spectrum, that is light with a longer wavelength and
lower frequency than visible light. It covers a range of techniques, mostly based on absorption
spectroscopy.

VIBRATIONAL MODES
COMBINATION BANDS, OVERTONES AND FERMI RESONANCES
INTRODUCTION TO VIBRATIONS
ISOTOPE EFFECTS IN VIBRATIONAL SPECTROSCOPY
MODE ANALYSIS
NORMAL MODES
NUMBER OF VIBRATIONAL MODES IN A MOLECULE
SYMMETRY ADAPTED LINEAR COMBINATIONS
BACK MATTER
INDEX

INFRARED SPECTROSCOPY
Infrared Spectroscopy is the analysis of infrared light interacting with a molecule. This can be analyzed in three ways by measuring
absorption, emission and reflection. The main use of this technique is in organic and inorganic chemistry. It is used by chemists to
determine functional groups in molecules. IR Spectroscopy measures the vibrations of atoms, and based on this it is possible to
determine the functional groups.

HOW AN FTIR SPECTROMETER OPERATES


IDENTIFYING THE PRESENCE OF PARTICULAR GROUPS
INFRARED: APPLICATION
INFRARED: INTERPRETATION
INFRARED SPECTROSCOPY
INTERPRETING INFRARED SPECTRA

CARBON NITROGEN BONDS


IR10. MORE PRACTICE WITH IR SPECTRA
IR11. APPENDIX: IR TABLE OF ORGANIC COMPOUNDS
IR2. HYDROCARBON SPECTRA
IR3. SUBTLE POINTS OF IR SPECTROSCOPY
IR4. CARBON CARBON MULTIPLE BONDS
IR5. CARBON OXYGEN SINGLE BONDS
IR6. CARBON OXYGEN DOUBLE BONDS
IR8. MORE COMPLICATED IR SPECTRA
IR9. MISLEADING PEAKS
WHAT DOES AN IR SPECTRUM LOOK LIKE?
IR SPECTROSCOPY BACKGROUND
THE FINGERPRINT REGION
BACK MATTER
INDEX

RAMAN SPECTROSCOPY
Raman spectroscopy is a chemical instrumentation technique that exploits molecular vibrations.

RAMAN: APPLICATION
RAMAN: THEORY
RESONANT VS. NONRESONANT RAMAN SPECTROSCOPY

1 7/7/2021
CHAPTER OVERVIEW
VIBRATIONAL MODES

COMBINATION BANDS, OVERTONES AND FERMI RESONANCES


Combination bands, overtones, and Fermi resonances are used to help explain and assign peaks in
vibrational spectra that do not correspond with known fundamental vibrations. Combination bands
and overtones generally have lower intensities than the fundamentals, and Fermi resonance causes
a spilt and shift in intensity of peaks with similar energies and identical symmetries. Hot bands
will also be briefly addressed.

INTRODUCTION TO VIBRATIONS
IR spectroscopy which has become so useful in identification, estimation, and structure
determination of compounds draws its strength from being able to identify the various vibrational
modes of a molecule. A complete description of these vibrational normal modes, their properties and their relationship with the
molecular structure is the subject of this article.

ISOTOPE EFFECTS IN VIBRATIONAL SPECTROSCOPY


This page provides an overview of how an isotope can affect the frequencies of the vibrational modes of a molecule. Isotopic
substitution is a useful technique due to the fact that the normal modes of an isotopically substituted molecule are different than the
normal modes of an unsubstituted molecule, leading to different corresponding vibrational frequencies for the substituted atoms.

MODE ANALYSIS
NORMAL MODES
Normal modes are used to describe the different vibrational motions in molecules. Each mode can be characterized by a different type
of motion and each mode has a certain symmetry associated with it. Group theory is a useful tool in order to determine what
symmetries the normal modes contain and predict if these modes are IR and/or Raman active. Consequently, IR and Raman
spectroscopy is often used for vibrational spectra.

NUMBER OF VIBRATIONAL MODES IN A MOLECULE


All atoms in a molecule are constantly in motion while the entire molecule experiences constant translational and rotational motion. A
diatomic molecule contains only a single motion. Polyatomic molecules have more than one type of vibration, known as normal
modes.

SYMMETRY ADAPTED LINEAR COMBINATIONS


The construction of linear combinations of the basis of atomic movements allows the vibrations belonging to irreducible
representations to be investigated. The wavefunction of these symmetry equivalent orbitals is referred to as Symmetry Adapted Linear
Combinations, or SALCs.

BACK MATTER

INDEX

1 7/7/2021
Combination Bands, Overtones and Fermi Resonances
Combination bands, overtones, and Fermi resonances are used to help explain and assign peaks in vibrational spectra that do
not correspond with known fundamental vibrations. Combination bands and overtones generally have lower intensities than
the fundamentals, and Fermi resonance causes a spilt and shift in intensity of peaks with similar energies and identical
symmetries. Hot bands will also be briefly addressed.

Introduction
Fundamental vibrational frequencies of a molecule corresponds to transition from v=0 to v=1. For a non-linear molecule there
will by 3N-6 (where N is the number of atoms) number vibrations. The same holds true for linear molecules, however the
equations 3N-5 is used, because a linear molecule has one less rotational degrees of freedom. (For a more detailed explanation
see: Normal Modes). Figure 1 shows a diagram for a vibrating diatomic molecule. The levels denoted by vibrational quantum
numbers v represent the potenital energy for the harmonic (quadratic) oscillator. The transition 0 → 1 is fundamental,
transitions 0 → n (n>1) are called overtones, and transitions 1 → n (n>1) are called hot transitions (hot bands).

Figure 1: Potential energy diagram for a vibrating diatomic molecule.

Symmetry Requirements
The symmetry requirement of vibrational transisition is given by the transition moment integral,


μ =∫ ψ μψdτ ≠ 0 (1)

where,
μ = iμx + j μy + kμz (2)

These integrals can be separated into each component: x,y, and z. Because the ground state contains the totally symmetric
representation, the coordinate x, y, or z and ψ* must belong to the same representation so that the direct product will contain
the totally symmetric representation.

Harmonic Oscillator Breakdown


The harmonic oscillator approximation is convenient to use for diatomic molecules with quantized vibrational energy levels
given by the following equation:

−1
1
Ev (c m ) = (v + ) ωe (3)
2

A more accurate description of the vibrational energies is given by the anharmonic oscillator (also called Morse potential) with
energy of
2 3
1 1 1
−1
Ev (c m ) = ωe (v + ) − ωe xe (v + ) + ωe ye (v + ) +. . . (4)
2 2 2

where ωe is the vibrational frequency for the re internuclear separation and ωe >> ωexe >> ωeye. This accounts for the fact that as
the higher vibrational states deviate from the perfectly parabolic shape, the level converge with increasing quantum numbers.

6/8/2021 1 https://chem.libretexts.org/@go/page/1853
It is because of this anharmoniticity that overtones can occur.
While it may seem that the harmonic oscillator and the anharomic oscillator are closely related, this is in fact not the case. The
differences in the wavefunctions lead to a breakdown of selection rules, specifically, Δv=±1 selection rule can not be applied,
and higher order terms must be accounted in the energy calculations.

Figure 2: Pictured above is the HOA (green parabola) superimposed on the anharmonic oscillator (blue curve) on a potential
energy diagram. V(R) is the potential energy of a diatomic molecule and R is the radius between the centers of the two atoms.
Towards the left is compression of the bond, towards the right is extension. Image used with prmission (CC-By-SA-3.0;
Created by Mark Somoza March 26 2006).
There is only a small correction from the ground state to the first excited state for the anharmonic correction, but it becomes
much larger for more highly excited states which are populated as the temperature increases. The deviation from the harmonic
oscillator to the anharmonic oscillator results in expanding the energy function with additional terms and treating these terms
with perturbation theory. The results in the correct vibrational energies and also relaxes the selection rules. A Δv=±1 is still
most predominant, however, weaker overtones with Δv=±2, ±3,… can occur. It should be noted that a Δv=2 transition does not
occur at twice the frequency of the fundamental transition, but at a lower frequency. Overtone transitions are not always
observed, especially in larger molecules, because the transitions become weaker with increasing Δv.

Overtones
Overtones occur when a vibrational mode is excited from v = 0 to v = 2 , which is called the first overtone, or v=0 to v=3,
the second overtone. The fundamental transitions, \(v=±1\0, are the most commonly occurring, and the probability of
overtones rapid decreases as the number of quanta (Δv=±n) increases. Based on the harmonic oscillator approximation, the
energy of the overtone transition would be n times larger than the energy of the fundamental transition frequency, but the
anharmonic oscillator calculations show that the overtones are less than a multiple of the fundamental frequency. This is
demonstrated with the vibrations of the diatomic HCl in the gas phase:
Table 1: HCl vibrational spectrum.
Transition Term ṽ obs [cm-1] ṽ Harmonic [cm-1] ṽ Anharmonic [cm-1]

0 → 1 fundamental 2,885.9 2,885.9 2,885.3

0 → 2 first overtone 5,668.0 5,771.8 5,665.0


0 → 3 second overtone 8,347.0 8,657.7 8,339.0
0 → 4 third overtone 10,923.1 11,543.6 10,907.4
0 → 5 fourth overtone 13,396.5 14,429.5 13,370

We can see from Table 1, that the anharmonic frequencies correspond much better with the observed frequencies, especially as
the vibrational levels increase.

Special case
If one of the symmetries is doubly degenerate in the excited state a recursion formula is required to determine the symmetry of
the vth wave function, given by,
1 v
χv (R) = [χ(R)χv−1 (R) + χ(R )] (5)
2

6/8/2021 2 https://chem.libretexts.org/@go/page/1853
Where χv(R) is the character under the operation R for the vth energy level; χ(R) is the character under R for the degenerate
irreducible representation; χv-1(R) is the character of the (v-1)th energy level; and χ(Rv) is the character of the operation Rv.
This is demonstrated for the D3h point group below.

Combination Bands
Combination bands are observed when more than two or more fundamental vibrations are excited simultaneously. One reason
a combination band might occur is if a fundamental vibration does not occur because of symmetry. This is comparable to
vibronic coupling in electronic transitions in which a fundamental mode can be excited and allowed as a “doubly excited
state.” Combination implies addition of two frequencies, but it also possible to have a difference band where the frequencies
are subtracted.
To determine if two states can be excited simultaneous the transition moment integral must be evaluated with the appropriate
excited state wavefunction. For example, in the transition,
1 2 3 1 2 3
ψ (0) ψ (0) ψ (0) → ψ (2) ψ (0) ψ (1) (6)

the symmetry of the excited state will be the direct product of the irreducible representation for ψ1(2) and ψ3(1).
For example, in the point group C4v, v1 has symmetry e and v3 has symmetry a2. By performing the calculations listed above,
it is determined that ψ1(2) has (a1 + b1 + b2) symmetry:

Γ[ ψes ] = Γ[ ψ1 (2)] ⊗ Γ[ ψ3 (1)] = (a1 + b1 + b2 ) × a2 = a2 + b2 + b1 (7)

A practical use for understanding overtones and combination bands is applied to organic solvents used in spectroscopy. Most
organic liquids have strong overtone and combination bands in the mid-infrared region, therefore, acetone, DMSO, or
acetonitrile should only be used in very narrow spectral regions. Solvents such at CCl4, CS2 and CDCl3 can be used above
1200 cm-1.

Hot Bands
Hot bands are observed when an already excited vibration is further excited. For example an v1 to v1' transition corresponds to
a hot band in its IR spectrum. These transitions are temperature dependent, with lower signal intensity at lower temperature,
and higher signal intensity at higher temperature. This is because at room temperature only the ground state is highly
populated (kT ~ 200 cm-1), based on the Boltzmann distribution. The Maxwell-Boltzmann distribution law states that if
molecules in thermal equilibrium occupy two states of energy εj and εi, the relative populations of molecules occupying these
states will be,
− εj /RT
nj e
−Δε/RT
= = e (8)
− εi /RT
ni e

where, k is the Boltzmann constant and T is the temperature in Kelvin.


In the harmonic oscillator model, hot bands are not easily distinguished from fundamental transitions because the energy levels
are equally spaced. Because the spacing between energy levels in the anharmonic oscillator decrease with increasing
vibrational levels, the hot bands occur at lower frequencies than the fundamentals. Also, the transition moment integrals are
slightly different since the ground state will not necessarily be totally symmetric since it is not in v=0.
1 2 3 1 2 3
ψ (0) ψ (0) ψ (1) → ψ (0) ψ (0) ψ (2) (9)

6/8/2021 3 https://chem.libretexts.org/@go/page/1853
Figure 4: Example of hot bands in a vibrational line spectrum of a diatomic molecule: (A) harmonic frequencies; (B) hot band
transitions; (C) combination of both spectra.

Fermi Resonances
Fermi resonance results in the splitting of two vibrational bands that have nearly the same energy and symmetry in both IR and
Raman spectroscopies. The two bands are usually a fundamental vibration and either an overtone or combination band. The
wavefunctions for the two resonant vibrations mix according to the harmonic oscillator approximation, and the result is a shift
in frequency and a change in intensity in the spectrum. As a result, two strong bands are observed in the spectrum, instead of
the expected strong and weak bands. It is not possible to determine the contribution from each vibration because of the
resulting mixed wave function.
If the symmetry requirements are fulfilled and the energies of the two states are similar, mixing occurs, and the resulting
modes can be described by a linear combination of the two interacting modes. The effect of this interaction is to increase the
splitting between the engery levels. The splitting will be larger if the original energy difference is small and the coupling
energy is large. The mixing of the two states also equalized the intensities of the vibrations which allows a weak overtone or
combination band to show significant intensity from the fundamental with which it has Fermi resonance with.

Figure 5. Example of intensity and frequency shifts due to Fermi resonance. The top bands represent two fundamental
vibrations without Fermi resonance, and the bottom bands show the change in bands as a result.The two energy levels are spilt
such that one increases and the other decreases in energy, known as a “Fermi doublet,” and they move away from each other.
Because the vibrations have nearly the same frequency, the interaction will be affected if one mode undergoes a frequency
shift from deuteration or a solvent effect while the other does not.The molecule most studied for this type of resonance (even
what Fermi himself used to explain this phenomena), is carbon dioxide, CO2. The three fundamental vibrations are v1= 1337
cm-1, v2=667 cm-1, v3=2349 cm-1. The first overtone of v2 is v1 + 2v2 with symmetries σg+ and (σg+ + δg+), respectively, and
frequencies of 1337 cm-1 (v1) and 2(667) = 1334 cm-1 (v2). According to group theory calculations, CO2 belongs to the point
group D∞h and should only have one Raman (symmetric stretching vibrations) and two IR active modes (asymmetric
stretching and bending vibrations). CS2 is an analog to this system.
Another typical example of Fermi resonance is found in the vibrational spectra of aldehydes, where the C-H bond in the CHO
group interacts with the second harmonic level, 2δ(CHO), derived from the fundamental frequency of the deformation
vibration of the CHO group (2*1400 cm-1). The result is a Fermi doublet with branches around 2830 and 2730 cm-1. It is

6/8/2021 4 https://chem.libretexts.org/@go/page/1853
important for Fermi resonance that the vibrations connected with the two interacting levels be localized in the same part of the
molecule.
When bands have non-negligible widths, Fermi resonance perturbation of localized levels cannot be applied. This broadening
can be the result of a number of things, such as, intermolecular interaction, shortened excited state lifetimes, or interaction of
vibrational modes with phonons. In place of perturbation theory, the distribution of interacting vibrational states can be
approximated as a collection of discrete level. The influence from each level can be calculated.
It is useful to understand Fermi resonance because it helps assign and identify peaks within vibrational spectra (ie. IR and
Raman) that may not otherwise be accounted for, however it should not be used lightly when assigning spectra.It is easy to
jump to the conclusion that an unidentifiable band is the result of Fermi resonance, however this explanation may not fully
account for the inconsistency and further characterization may be required for the system being investigated. It is important to
assign spectra before doing the normal mode (coordinate) calculations because doing these calculations beforehand often leads
to incorrect assignments of the peaks in the spectra.

References
1. Atkins, P. W. Molecular Quantum Mechanics. Oxford University Press, 1970.
2. Guillory, William. Introduction to Molecular Structure and Spectroscopy. Boston, MA: Allyn & Bacon, Inc. , 1977.
3. Struve, Walter. Fundamentals of Molecular Spectroscopy. Wiley, 1989.
4. Harris, Daniel, and Michael Bertolucci. Symmetry and Spectroscopy . Oxford University Press, 1978.
5. Cotton, F. Albert. Chemical Applications of Group Theory . 3rd. Wiley, 1990.
6. Horak, M., and Antonin Vitek. Interpretations and Processing of Vibrational Specta. Wiley, 1978.
7. Diem, M. Introduction to Modern Vibrational Spectroscopy. Wiley, 1993.
8. Kondratyuk, P. "Analytical Formulas for Fermi Resonance Interactions in Continuous Distributions of States."
Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 61.4 (2005): 589-93. Web.

Problems
Q1. Given ν1 = 1151 cm-1, ν2 = 1361 cm-1, ν3 = 519 cm-1 for SO2, and the fact that there are 4 overtones and/or combination
bands, predict the vibrational spectra and calculations.
A1.
v [cm-1] Assignment

519 v2

606 ν1 - v2
1151 v1
1361 v3
1871 v2 + v3
2305 2v1
2499 v1 + v3

Q2. What are the two main effects of Fermi resonance?


A2. An overtone band can gain intensity from a nearby fundamental frequency with similiar symmetry. The energy levels of
both bands are shifted away from one another.
Q3. Explain the difference between a combination band and an overtone.
An overtone is the result of Δv>1 from the ground state. A combination band is the result of a 2 fundamental frequencies being
excited simultaneously so that the excitation is allowed by symmetry. The overtone is not subject to a symmetry requirement.
Q4. Why are hot bands temperature dependent?
A4. For a hot band to occur, a state other than the ground state must already be populated, and this requires >200cm-1 to over
come the thermal energy of kT (Boltzmann constant times temperature). The more heat that it put into the system, the more
likely a hot band is to occur, and the stronger the signal it will produce.

6/8/2021 5 https://chem.libretexts.org/@go/page/1853
Q5. Show the calculations for the values in Figure 1 for both the harmonic and anharmonic oscillators.
A5. Equations to use: ṽ = 2885.90v, for harmonic and ṽ = 2990.9v – 52.82v(v+1) for anharmonic.
For v=3, ṽ H = 2885.90(3) = 8657.7 ṽ AH = 2990.9(3) – 52.82(3)(3+1) = 8339.0

Contributors and Attributions


Alexandra Holmes (University of California, Davis)

6/8/2021 6 https://chem.libretexts.org/@go/page/1853
Introduction to Vibrations
IR spectroscopy which has become so useful in identification, estimation, and structure determination of compounds draws its
strength from being able to identify the various vibrational modes of a molecule. A complete description of these vibrational
normal modes, their properties and their relationship with the molecular structure is the subject of this article.

Introduction
We are familiar with resolving a translational vector into its three components along the x-, y-, and z- axes. Similarly a
rotational motion can also be resolved into its components. Likewise the same is true for vibrational motion. The complex
vibration that a molecule is making is really a superposition of a number of much simpler basic vibrations called “normal
modes”. Before we take up any further description of “normal modes” it is necessary to discuss the degrees of freedom.

Degrees of Freedom
Degree of freedom is the number of variables required to describe the motion of a particle completely. For an atom moving in
3-dimensional space, three coordinates are adequate so its degree of freedom is three. Its motion is purely translational. If we
have a molecule made of N atoms (or ions), the degree of freedom becomes 3N, because each atom has 3 degrees of freedom.
Furthermore, since these atoms are bonded together, all motions are not translational; some become rotational, some others
vibration. For non-linear molecules, all rotational motions can be described in terms of rotations around 3 axes, the rotational
degree of freedom is 3 and the remaining 3N-6 degrees of freedom constitute vibrational motion. For a linear molecule
however, rotation around its own axis is no rotation because it leave the molecule unchanged. So there are only 2 rotational
degrees of freedom for any linear molecule leaving 3N-5 degrees of freedom for vibration.

Vibrational modes
1. A normal mode is a molecular vibration where some or all atoms vibrate together with the same frequency in a defined
manner.
2. Normal modes are basic vibrations in terms of which any other vibration is derived by superposing suitable modes in the
required proportion.
3. On the other hand, no normal mode is expressible in terms of any other normal mode. Each one is pure and has no
component of any other normal mode (i.e. they are orthogonal to each other). Mathematically, the integral is
∫ ψA ψB dR = 0 (integration is done over the entire space)
4. The required number of “normal modes” is equal to the vibrational degree of freedom available so the number of modes for
a nonlinear molecule is 3N − 6 and that for a linear molecule is 3N − 5 .
5. Each mode has a definite frequency of vibration. Sometimes 2 or 3 modes may have the same frequency but that does not
change the fact that they are distinct modes; these modes are called degenerate.
6. Sometimes some modes are not IR active but they exist all the same. We shall revert back to the problem of IR activity and
selection rules later.
The number of vibrational normal modes can be determined for any molecule from the formula given above. For a diatomic
molecule, N = 2 so the number of modes is 3 × 2 − 5 = 1 . For a triatomic linear molecule (CO2), it is 3 × 3 − 5 = 4 and
triatomic nonlinear molecule (H2O), it is 3 × 3 − 6 = 3 and so on.

Example 1: Water

1. The Symmetric Stretch (Example shown is an H2O molecule at 3685 cm-1)


2. The Asymmetric Stretch (Example shown is an H2O molecule at 3506 cm-1)

6/29/2021 1 https://chem.libretexts.org/@go/page/77584
3. Bend (Example shown is an H2O molecule at 1885 cm-1)

A linear molecule will have another bend in a different plane that is degenerate or has the same energy. This accounts for the
extra vibrational mode.

Example 2: Carbon Dioxide

Example 3: The Methylene Group

It is important to note that there are many different kinds of bends, but due to the limits of a 2-dimensional surface it is
not possible to show the other ones.

The frequency of these vibrations depend on the inter atomic binding energy which determines the force needed to stretch or
compress a bond. We discuss this problem in the next section. The determination of the nature of the relative displacement of
each atom with respect to each other is more complicated and beyond the scope of this article. However, such motion can be
seen in some common molecules as shown below.

Energetics
For studying the energetics of molecular vibration we take the simplest example, a diatomic heteronuclear molecule AB.
Homonuclear molecules are not IR active so they are not a good example to select. Let the respective masses of atoms A and B
be m and m . So the reduced mass μ is given by:
A B AB

mA mB
μAB = (1)
mA + mB

The equilibrium internuclear distance is denoted by r . However as a result of molecular vibrations, the internuclear distance
eq

is continuously changing; let this distance be called r(t). Let x(t) = r(t) − r . When x is non-zero, a restoring force F
eq

exists which tries to bring the molecule back to x = 0 , that is equilibrium. For small displacements this force can be taken to
be proportional to x.
F = −kx (2)

where k is the force constant.


The negative sign arises from the fact that the force acts in the direction opposite to x . This is indeed a case of Simple
Harmonic Motion where the following well known relations hold.

x(t) = A sin(2πν t) (3)

where

6/29/2021 2 https://chem.libretexts.org/@go/page/77584
−−−−
1 k
ν = √ (4)
2π μAB

The potential energy is given by V =


1

2
kx
2
. The total energy E (Kinetic+Potential) is obtained by solving the Schrödinger
equation:
2 2
h d ψ 1
2
− + kx ψ = Eψ (5)
2 2
8π μAB dx 2

A set of wave functions ψ ) and the corresponding Eigenvalues E are obtained. E = (n + (1/2))hv where n is an integer
n n n

(-1,0,1,2 etc.). The energy is quantized, the levels are equally spaced, the lowest energy is (1/2)hv, and the spacing between
adjacent levels is hv.

Interaction with Electromagnetic Radiation


As show above, the energy difference between adjacent vibrational energy levels is hvvibration. On the other hand, the photon
energy is hvphoton. Energy conservation requires that the first condition for photon absorption be,
Hvvibration = hvphoton or vvibration = vphoton.
Such photons are in IR region of the electromagnetic spectrum. In addition, two more conditions must be met.
1. For absorption of electromagnetic radiation, the dipole moment of the molecule must change with increasing internuclear
separation resulting from the vibration (i.e, dμ/dD ≠ 0 ).
2. The probability of a transition from one state to another is large if one of the state is odd and another even. This is possible
if nfinal – ninitial = +1 (for absorption). At room temperature, modes are predominantly in energy state n = 0, so this
transition is from n = 0 to n = 1, and ΔE = hν .

Applications
Spectroscopy in the IR region can determine the frequency and intensity of absorption. These frequencies are generally
specific for a specific bonds such as c-c, c(double bond)c, c(triple bond)c, c-o, c(double bond)o, etc. So the IR absorption data
is very useful in structure determination. The intensity depends on the concentration of the resposble spec. So it is useful for
quantitative estimation and for identification.

Questions
1. Find the number of vibrational modes for the following molecules: N H , C H , C H , C H , C H (linear).
3 6 6 10 8 4 2 2

2. State which of the following vibrations are IR active: N , C O, C O (stretching), H C l


2 2

3. Calculate the vibrational frequency of C O given the following data: mass of C = 12.01 amu, mass of O = 16 amu, the force
constant k = 1.86 × 10 kg ⋅ s .
3 −2

4. Calculate the vibrational energy in Joules per mole of a normal mode in question 3, in its ground state of n = 0 .
5. Assuming the force constant to be the same for H O and D O . A normal mode for H O is at 3650 cm . Do you expect
2 2 2
−1

the corresponding D O wave number to be higher or lower?


2

Answers
1.) NH3 – 6
C6H6 –30
C10H8 –48
CH4 –9
C2H2 – 7
2.) N2 – IR inactive
C0 – active
C02 (stretching) – inactive
HCl – active

6/29/2021 3 https://chem.libretexts.org/@go/page/77584
3.) mAB = mAxmB/(mA+mB) = 11.395x10-27
v = (1/2pi)(k/mAB).5 = 2143.3 cm-1
4.) Energy of the mode for n = 0
E0 = (1/2)hv = 2.13x10-20J
Energy per mole = 2.13x10-20x6.022x1023 = 12.8KJ/mole
5.) v for D2O will be lower because v is inversely proportional to 1/(m.5), where m is the reduced mass.

Outside Links:
Youtube video of Benzene vibrational modes: www.youtube.com/watch?v=ygjPqI0MIj8

References
1. Atkins, Peter and Julio de Paula. Physical Chemistry for the Life Sciences. 2006. New York, NY: W.H. Freeman and
Company. p. 550-555
2. Chang, Raymond. Physical Chemistry for the Biosciences. USA: University Science Books, 2005
3. Vollhardt, K. Peter C., and N. Schore. Organic chemistry structure and function. New York: W.H. Freeman, 2007.

Contributors and Attributions


Richard Banks, Charles Chu, Gaurav Sinha

6/29/2021 4 https://chem.libretexts.org/@go/page/77584
Isotope Effects in Vibrational Spectroscopy
This page provides an overview of how an isotope can affect the frequencies of the vibrational modes of a molecule. Isotopic
substitution is a useful technique due to the fact that the normal modes of an isotopically substituted molecule are different
than the normal modes of an unsubstituted molecule, leading to different corresponding vibrational frequencies for the
substituted atoms. Vibrational spectroscopy is done in the infrared region of the electromagnetic spectrum, which ranges from
around 10-6 to 10-3 meters. IR and Raman spectroscopy observe the vibrations of molecules, displaying the normal and local
modes of the molecule in the spectra. Isotopes are atoms that share the same number of protons but differ in the number of
neutrons contained in the nucleus, thus giving these atoms different mass numbers. The specific mass of each atom will affect
the reduced mass of the overall molecule, therefore changing the vibrational frequencies of that molecule.

Diatomics
A diatomic molecule, as seen in Figure 1, contains two atoms, which can either be composed of the same or different
elements. It is easier to focus on these types of molecules when analyzing and calculating vibrational frequencies because they
are simpler systems than polyatomic molecules. Whether or not the diatomic consists of the same or different elements, a
diatomic molecule will have only one vibrational frequency. This singular normal mode is because of the diatomic's linear
symmetry, so the only vibration possible occurs along the bond connecting the two atoms.

Figure 1: Diagram of a diatomic molecule with the only possible vibration it can undergo.

Normal Modes
Normal modes describe the possible movements/vibrations of each of the atoms in a system. There are many different types of
vibrations that molecules can undergo, like stretching, bending, wagging, rocking, and twisting, and these types can either be
out of plane, asymmetric, symmetric, or degenerate. Molecules have 3n possible movements due to their 3-dimensionality,
where n is equal to the number of atoms in the molecule. Three movements are subtracted from the total because they are
saved for the displacement of the center of mass, which keeps the distance and angles between the atoms constant. Another 3
movements are subtracted from the total because they are for the rotations about the 3 principle axes. This means that for
nonlinear molecules, there are 3n – 6 normal modes possible. Linear molecules, however, will have 3n – 5 normal modes
because it is not possible for internuclear axis rotation, meaning there is one less possible rotation for the molecule. This
explain why diatomic molecules only have 1 vibrational frequency, because 3(2) – 5 = 1.
Molecular vibrations are often thought of as masses attached by a spring (Figure 2),

Figure 2: Representation of a diatomic molecule as masses attached by a spring separated by an equilibrium distance of req.
and Hook’s law can be applied

F = −kx (1)

where
F is the resulting force,
x is the displacement of the mass from equilibrium (x = r– r eq ) , and
k is the force constant, defined as

6/28/2021 1 https://chem.libretexts.org/@go/page/1854
2
∂ V (r)
k =( ) (2)
2
∂r
req

Figure 3: Depiction of the diatomic when the atoms undergo a vibration and increase their separation by a distance x.
1
in which V (r) = k(r − req ) , which comes from incorporating Hook’s law to the harmonic oscillator. The diatomic
2
molecule is thought of as two masses (m1 and m2) on a spring, they will have a reduced mass, µ, so their vibrations can be
mathematically analyzed.
m1 m2
μ = (3)
m1 + m2

When an atom in a molecule is changed to an isotope, the mass number will be changed, so µ will be affected, but k will not
(mostly). This change in reduced mass will affect the vibrational modes of the molecule, which will affect the vibrational
spectrum.
Vibrational energy levels, ν , are affected by both k and µ, and is given by
e



1 k
νe = √ (4)
2π μ

These vibrational energy levels correspond to the peaks which can be observed in IR and Raman spectra. IR spectra observe
the asymmetric stretches of the molecule, while Raman spectra observe the symmetric stretches.

Effects on Experimental Results


When an atom is replaced by an isotope of larger mass, µ increases, leading to a smaller ν and a downshift (smaller
e

wavenumber) in the spectrum of the molecule. Taking the diatomic molecule HCl, if the hydrogen is replaced by its isotope

deuterium, µ is doubled and therefore ν will be decreased by √2. Deuterium substitution leads to an isotopic ratio of 1.35-
e

1.41 for the frequencies corresponding to the hydrogen/deuterium vibrations. There will also be a decrease by √2 in the band
width and integrated band width for the vibrational spectra of the substituted molecule. Isotopic substitution will affect the
entire molecule (to a certain extent), so it is not only the vibrational modes for the substituted atom that will change, but rather
the vibrational modes of all the atoms of the molecule. The change in frequency for the atoms not directly invovled in the
substitution will not display as large a change, but a downshift can still occur.

Figure 4: Structure of polyaniline.


When polyaniline (Figure 4) is fully deuterated, the vibrational peaks will downshift slightly. The following data was
summarized from Quillard et al.

Type of vibration Nondeuterated (frequency, cm-1) Deuterated (frequency, cm-1)

C-C stretch 1626 1599

C-C stretch 1581 1560


C-H bend – benzenoid ring 1192 876
C-H bend – quinoid ring 1166 856

6/28/2021 2 https://chem.libretexts.org/@go/page/1854
Type of vibration Nondeuterated (frequency, cm-1) Deuterated (frequency, cm-1)

N-H bend 1515 1085

Changing hydrogen to deuterium leads to the largest effect in a vibrational spectrum since the mass is doubled. Other isotopic
substitutions will also lead to a shift in the vibrational energy level, but because the mass change is not as significant, µ will
not change by much, leading to a smaller change in ν . This smaller change in vibrational frequency is seen in the sulfur
e

substitution of sulfur hexafluoride (Figure 5:), from 32S to 34S. The frequencies as reported by Kolomiitsova et al. are shown
below.

Figure 5: Structure of sulfur hexafluoride.


32
Vibration assignment SF6 (frequency, cm-1) 34
SF6 (frequency, cm-1)

ν3 939.3 922.2

ν4 613.0 610.3

These two examples show the consistency of downshifted vibrational frequencies for atoms substituted with an isotope of
higher mass.

Applications
Substituting atoms with isotopes has been shown to be very useful in determining normal mode vibrations of organic
molecules. When analyzing the spectrum of a molecule, isotopic substitution can help determine the vibrational modes specific
atoms contribute to. Those normal modes can be assigned to the peaks observed in the spectrum of the molecule. There are
specific CH3 rocks and torsions, as well as CH bends that can be identified in the spectrum upon deuterium substitution. Other
torsion bands from hydroxyl and amine groups can also be assigned when hydrogen is replaced with deuterium. Experimental
data has also shown that using deuterium substitution can help with symmetry assignments and the identification of metal
hydrides.

Isotopic substitution can also be used to determine the force constants of the
molecule. Calculations can be done using the frequencies of the normal modes in
determining these values, based on both calculated frequencies and experimental
frequencies.
Researchers have also attempted to contribute peak shape changes and splits in peaks of vibrational spectra to naturally
occurring isotopes in molecules. It has been shown, however, that the shape of a peak is not related to the size of the atom, so
substitution to an atom of larger mass will not affect the peak shape in the molecule's spectrum. As previously stated, isotopic
substitution of atoms of higher mass will not have a significant enough effect on the shifts in frequencies for the corresponding
vibrations, so analyzing the frequency shifts of smaller mass isotopes, like deuterium and 13C is necessary.

Figure 6: Spectra of unsubstituted and substituted TCDD depicting the isotopic 13C effects.

6/28/2021 3 https://chem.libretexts.org/@go/page/1854
As depicted in the rough representation of the vibrational spectra of the molecule tetrachlorinated dibenzodioxin (TCDD), the
13
C substituted TCDD spectrum is slightly downshifted compared to the unsubstituted TCDD spectrum. Although the shifts
and split peaks do occur in the spectra of isotopically substituted molecules, not all observed peaks can be attributed to the
isotope. This is because the intensities of the peaks shown are not large enough to relate to the natural abundance of the 13C
isotope, and not all peaks can be accounted for by the substitution.

References
1. Harris, D. C.; Bertolucci, M. D. Symmetry and Spectroscopy; Oxford University Press, Inc.: New York, NY, 1989.
2. Kolomiitsova, T. D.; Kondaurov, V. A.; Sedelkova, E. V.; Shchepkin, D. N. Opt. Spectrosc. 2002, 92, 564. DOI:
10.1134/1.1473589
3. Mielke, Z.; Sobczyk, L. In Isotope Effects in Chemistry and Biology; Kohen, A., Limbach, H.-H., Eds. 2006, p 281.
4. Quillard, S.; Louarn, G.; Buisson, J. P.; Boyer, M.; Lapkowski, M.; Pron, A.; Lefrant, S. Synth. Met. 1997, 84, 805. DOI:
10.1016/S0379-6779(96)04155-0
5. Rauhut, G.; Pulay, P. J. Am. Chem. Soc. 1995, 117, 4167. DOI: 10.1021/ja00119a034
6. Signorell, R.; Kunzmann, M. K. Chem. Phys. Lett. 2003, 371, 260. DOI: 10.1016/S0009-2614(03)00253-7

Problems
1. How many normal modes would be found in CO2? What are the different types of vibrational modes for this molecule?
2. If the diatomic molecule HCl, with 1H and 35Cl were substituted with 37Cl, what change occurs to the reduced mass?
3. For a nondeuterated hydrofluoric acid diatomic, HF, the vibrational frequency of this molecule is found at 845 cm-1. If the
hydrogen atom of this molecule was substituted with deuterium, where would you expect to now find the vibrational
frequency?

Answers
1. 4; symmetric stretch, asymmetric stretch, and two degenerate bends
2. With 37Cl the reduced mass increases from 0.97222 to 0.97368
3. DF would have a calculated band at 597.5 cm-1

6/28/2021 4 https://chem.libretexts.org/@go/page/1854
Mode Analysis
ΓTotal = ΓStretch + ΓBend + ΓTranslation + ΓRotation
and
ΓVibration = ΓStretch + ΓBend
For easier understanding, there are steps immediately followed by examples using the D3h.

Finding Γ Total
First, add all the x, y and z rows on the Character Tables of Symmetry groups. If x, y or z are in () on the far right then only
count them once, otherwise count the row a second time (Keep the column separated). This is called Γx,y,z. Next, move the
molecule with the designated column symbols and if an atom does not move then it is counted. Finally, multiply Γx,y,z and
UMA and it will equal Γtotal.

D3h E 2C2 3C2 σh 2S3 3σv IR Raman

A1' 1 1 1 1 1 1 x2+y2, z2

A2' 1 1 -1 1 1 -1 Rz

E' 2 -1 0 2 -1 0 (x,y) (xy, x2-y2)

A1" 1 1 1 -1 -1 -1

A2" 1 1 -1 -1 -1 1 z

E" 2 -1 0 -2 1 0 (Rx,Ry) (xz, yz)

Γx,y,z 2+1=3 -1+1=0 0-1=-1 2-1=1 -1-1=-2 0+1=1

UMA
(unmoved 6 3 2 4 1 4
atoms)

Γtotal (3)(6)=18 (0)(3)= 0 (-1)(2)=-2 (1)(4)=4 (-2)(1)=-2 (1)(4)=4

D3h molecule
-need to find the irreducible representation of Gamma total.
Take Γtotal multiply by the number in front of the symbols (the order) and multiply by each number inside of the character
table. Add up each row, and divide each row by the total order. For the D3h the order is 12.

add the up and


D3h 1E 2C2 3C2 σh 2S3 3σv
divide by 12

A1' (1x1x18)=18 (2x1x0)=0 (3x1x-2)=-6 (1x1x4)=4 (2x1x-2)=-4 (3x1x4)=12 (24/12)=2

A2' (1x1x18)=18 (2x1x0)=0 (3x-1x-2)=-6 (1x1x4)=4 (2x1x-2)=-4 (3x-1x4)=-12 (0/12)=0

E' (1x2x18)=36 (2x-1x0)=0 (3x0x-2)=0 (1x2x4)=8 (2x-1x-2)=4 (3x0x4)=0 (48/12)=4

A1" (1x1x18)=18 (2x1x0)=0 (3x1x-2)=-6 (1x-1x4)=-4 (2x-1x-2)=4 (3x-1x4)=-12 (0/12)=0

A2" (1x1x18)=18 (2x1x0)=0 (3x-1x-2)=6 (1x-1x4)=-4 (2x-1x-2)=4 (3x1x4)=12 (36/12)=3

E" (1x2x18)=36 (2x-1x0)=0 (3x0x-2)=0 (1x-2x4)=-8 (2x1x-2)=-4 (3x0x4)=0 (24/12)=2

Γtotal= 2A1' + A2' + 4E' + 3A2" + 2E"

Γ Translation
The irreducible form of Γtrans, one needs to look at the second to last column. look at the row that the x, y and z and take the
irreducible representation. For instance the D3h would be Γtrans= E'+A2"

6/20/2021 1 https://chem.libretexts.org/@go/page/1855
Γ Rotation
One can find Γrot the same way as Γtrans. Instead of looking at x,y,z, one would look at the Rx, Ry, Rz. For the D3h. The Γrot =
A2'+E"

Γ Vibration
To find ΓVibration, just take Γtot -Γtrans-Γrot= Γvibration.
D3h example

Γtot 2A1' + A2' + 4E' + 3A2" + 2E"

Γtrans - E' - A2"


Γrot - A2' -E"
Γvibration 2A1' + 3E' + 2A2" + E"

Number of Vibrational Active IR Bands


Only Rx, Ry, Rz, x, y, and z can be ir active. which means only A2', E', A2", and E" can be IR active bands for the D3h. Next
add up the number in front of the irreducible representation and that is how many IR active bonds. For instance for the same
problem there are 3E'+2A2". There are 5 bands, three of them (meaning the E) are two fold degenerate.
Number of Vibrational Active Raman bands
Only x2+y2, z2, xy, xz, yz, x2-y2 can be Raman active. which means only A1', E', and E" can be raman active for the D3h.
Next add up the number in front of the irreducible representation, and that is how many Raman active bonds there are. For
instance for the same problem there are 3E' + E". There are four bands, 4 of which are two fold degenerate.

Finding ΓStretch
Looking at the molecules point group, do each of the symmetry representation and count the number of unmoved bonds.
Γstretch = Γσ =Γrad . next multiply unmoved bonds by the symmetry operations and then the numbers inside the character tables.
Then add the rows up and divide by the order of the point group.

D3h E 2C2 3C2 σh 2S3 3σv IR Raman

A1' 1 1 1 1 1 1 x2+y2, z2

A2' 1 1 -1 1 1 -1 Rz

E' 2 -1 0 2 -1 0 (x,y) (xy, x2-y2)

A1" 1 1 1 -1 -1 -1

A2" 1 1 -1 -1 -1 1 z

E" 2 -1 0 -2 1 0 (Rx,Ry) (xz, yz)

Γx,y,z 2+1=3 -1+1=0 0-1=-1 2-1=1 -1-1=-2 0+1=1

UMB
(unmoved 5 2 1 3 0 3
bonds)

add the up and


D3h 1E 2C2 3C2 σh 2S3 3σv
divide by 12

A1' (1x1x5)=5 (2x1x2)=4 (3x1x1)=3 (1x1x3)=3 (2x1x0)=0 (3x1x3)=9 (24/12)=2

A2' (1x1x5)=5 (2x1x2)=4 (3x-1x1)=-3 (1x1x3)=3 (2x1x0)=0 (3x-1x3)=-9 (0/12)=0

E' (1x2x5)=10 (2x-1x2)=-4 (3x0x1)=0 (1x2x3)=6 (2x-1x0)=0 (3x0x3)=0 (12/12)=1

A1" (1x1x5)=5 (2x1x2)=4 (3x1x1)=3 (1x-1x3)=-3 (2x-1x0=0 (3x-1x3)=-9 (0/12)=0

A2" (1x1x5)=5 (2x1x2)=4 (3x-1x1)=-3 (1x-1x3)=-3 (2x-1x0)=0 (3x1x3)=9 (12/12)=1

E" (1x2x5)=10 (2x-1x2)=-4 (3x0x1)=0 (1x-2x3)=-6 (2x1x0)=0 (3x0x3)=0 (24/12)=0

6/20/2021 2 https://chem.libretexts.org/@go/page/1855
Γstretch = Γσ = Γrad = 2A1' + 1E'+ 1A2"
Γπ= Γtan =

6/20/2021 3 https://chem.libretexts.org/@go/page/1855
Normal Modes
Normal modes are used to describe the different vibrational motions in molecules. Each mode can be characterized by a
different type of motion and each mode has a certain symmetry associated with it. Group theory is a useful tool in order to
determine what symmetries the normal modes contain and predict if these modes are IR and/or Raman active. Consequently,
IR and Raman spectroscopy is often used for vibrational spectra.

Overview of Normal Modes


In general, a normal mode is an independent motion of atoms in a molecule that occurs without causing movement to any of
the other modes. Normal modes, as implied by their name, are orthogonal to each other. In order to discuss the quantum-
mechanical equations that govern molecular vibrations it is convenient to convert Cartesian coordinates into so called normal
coordinates. Vibrations in polyatomic molecules are represented by these normal coordinates.
There exists an important fact about normal coordinates. Each of these coordinates belongs to an irreducible representation of
the point the molecule under investigation. Vibrational wavefunctions associated with vibrational energy levels share this
property as well. The normal coordinates and the vibration wavefunction can be categorized further according to the point
group they belong to. From the character table predictions can be made for which symmetries can exist. The irreducible
representation offers insight into the IR and/or Raman activity of the molecule in question.

Degrees of Freedom
3N where N represents the number of nuclei present in the molecule is the total number of coordinates needed to describe the
location of a molecule in 3D-space. 3N is most often referred to as the total number of degrees of freedom of the molecule
being investigated. The total number of degrees of freedom, can be divided into:
3 coordinates to describe the translational motion around the center of mass; these coordinates are called the translational
degrees of freedom
3 coordinates to describe the rotational motion in non-linear molecules; for linear molecules only 2 coordinates are
required; these coordinates are called the rotational degrees of freedom
the remaining coordinates are used to describe vibrational motion; a non-linear molecule has 3N - 6 vibrational degrees of
freedom whereas a linear molecule has 3N -5 degrees of freedom.
Table 1: Overview of degrees of freedom
Translational degrees of Rotational degrees of Vibrational degrees of
Total Degree of Freedom
freedom freedom freedom

Nonlinear Molecules 3N 3 3 3N -6

Linear Molecules 3N 3 2 3N - 5

Example 1: Ethane vs. Carbon Dioxide


Ethane, C H has eight atoms (\N=8\) and is a nonlinear molecule so of the 3N = 24 degrees of freedom, three are
2 6

translational and three are rotational. The remaining 18 degrees of freedom are internal (vibrational). This is consistent
with:
3N − 6 = 3(8) − 6 = 18 (1)

Carbon Dioxide, C O has three atoms (N = 3 and is a linear molecule so of the 3N = 9 degrees of freedom, three are
2

translational and two are rotational. The remaining 4 degrees of freedom are vibrational. This is consistent with:
3N − 5 = 3(3) − 5 = 4 (2)

Mathematical Introduction to Normal Modes


If there is no external field present, the energy of a molecule does not depend on its orientation in space (its translational
degrees of freedom) nor its center of mass (its rotational degrees of freedom). The potential energy of the molecule is therefore
made up of its vibrational degrees of freedom only of 3N − 6 (or 3N − 5 for linear molecules).

6/17/2021 1 https://chem.libretexts.org/@go/page/1856
The difference in potential energy is given by:
ΔV = V (q1 , q2 , q3 , . . . , qn ) − V (0, 0, 0, . . . , 0) (1)

Nvib Nvib 2
1 ∂ V
= ∑∑( ) qi qj (2)
2 ∂ qi ∂ qj
i=1 j=1

Nvib Nvib
1
= ∑ ∑ fij qi qj (3)
2
i=1 j=1

where
q represents the equilibrium displacement and
Nvib the number of vibrational degrees of freedom.
For simplicity, the anharmonic terms are neglected in this equation (consequently there are no higher order terms present). A
theorem of classical mechanics states that the cross terms can be eliminated from the above equation (the details of the
theorem are very complex and will not be discussed in detail). By using matrix algebra a new set of coordinates {Qj} can be
found such that
Nvib
1
2
ΔV = ∑ Fj Q (4)
j
2
j=1

Note that there are no cross terms in this new expression. These new coordinates are called normal coordinates or normal
modes. With these new normal coordinates in hand, the Hamiltonian operator for vibrations can be written as follows:
Nvib 2 2 Nvib
ℏ d 1
^ 2
H vib = − ∑ + ∑ Fj Q (5)
j
2μi dQ2 2
j=1 j j=1

The total wavefunction is a product of the individual wavefunctions and the energy is the sum of independent energies. This
leads to:
Nvib Nvib 2 2 Nvib
−ℏ d 1
^
H ^ 2
vib = ∑ H vib,j = ∑ ( + ∑ Fj Q )
j
(6)
2
2μj dQ 2
j=1 j=1 i j=1

and the wavefunction is then

ψvib = Q1 , Q2 , Q3 . . . , Qvib = ψvib,1 (Q1 )ψvib,2 (Q2 )ψvib,3 (Q3 ), . . . , ψvib,Nvib (QNvib ) (7)

and the total vibrational energy of the molecule is


Nvin
1
Evib = ∑ h νj ( vj + ) (8)
2
j=1

where
vj = 0, 1, 2, 3...

The consequence of the result stated in the above equations is that each vibrational mode can be treated as a harmonic
oscillator approximation. There are N harmonic oscillators corresponding to the total number of vibrational modes present
vib

in the molecule.
In the ground vibrational state the energy of the molecule is equal to (1/2)hνj. The ground state energy is referred to as zero
point energy. A vibration transition in a molecule is induced when it absorbs a quantum of energy according to E = hv. The
first excited state is separated from the ground state by Evib = (3/2)hν since vj = 1, the next energy level separation is (5/2)hν,
etc...
The harmonic oscillator is a good approximation, but it does not take into account that the molecule, once it has absorbed
enough energy to break the vibrating bond, does dissociate. A better approximation is the Morse potential which takes into

6/17/2021 2 https://chem.libretexts.org/@go/page/1856
account anharmonicity. The Morse potential also accounts for bond dissociation as well as energy levels getting closer together
at higher energies.

Pictorial description of normal coordinates using CO


The normal coordinate q is used to follow the path of a normal mode of vibration. As shown in Figure 2 the displacement of
the C atom, denoted by Δro(C), and the displacement of the O atom, denoted by Δro(O), occur at the same frequency. The
displacement of atoms is measured from the equilibrium distance in ground vibrational state, ro.

Figure 2: The Normal coordinate for C O is equation to Δr(C ) + Δr(O)

Description of vibrations
ν = stretching is a change in bond length; note that the number of stretching modes is equal to the number of bonds on the
molecule
δ = bending is a change in bond angle
ρr = rocking is change in angle between a group of atoms
ρw = wagging is change in angle between the plane of a group of atoms
ρt = twisting is change in angle between the planes of two groups of atoms
π= out of plane
In direct correlation with symmetry, subscripts s (symmetric), as (asymmetric) and d (degenerate) are used to further describe
the different modes.
A normal mode corresponding to an asymmetric stretch can be best described by a harmonic oscillator: As one bond
lengthens, the other bond shortens. A normal mode that corresponds can be best described by a Morse potential well: As the
bond length increases the potential energy increases and levels off as the bond length gets further away from the equilibrium.

The use of Symmetry and Group Theory


Symmetry of normal modes
It is important to realize that every normal mode has a certain type of symmetry associated with it. Identifying the point group
of the molecule is therefore an important step. With this in mind it is not surprising that every normal mode forms a basis set
for an irreducible representation of the point group the molecule belongs to. For a molecule such as water, having a structure
of XY2, three normal coordinates can be determined. The two stretching modes are equivalent in symmetry and energy. The
figure below shows the three normal modes for the water molecule:

Figure 3: Three normal modes of water


By convention, with nonlinear molecules, the symmetric stretch is denoted v1 whereas the asymmetric stretch is denoted v2.
Bending motions are v3. With linear molecules, the bending motion is v2 whereas asymmetric stretch is v3.
The water molecule has C2v symmetry and its symmetry elements are E, C2, σ(xz) and σ(yz). In order to determine the
symmetries of the three vibrations and how they each transform, symmetry operations will be performed.
As an example, performing C2 operations using the two normal mode v2 and v3 gives the following transformation:

6/17/2021 3 https://chem.libretexts.org/@go/page/1856
Once all the symmetry operations have been performed in a systematic manner for each modes the symmetry can be assigned
to the normal mode using the character table for C2v:
Table 2: Character table for the C2v point group
C2v E C2 σ (xz) σ (yz)

ν1 1 1 1 1 = a1

ν2 1 1 1 1 = a1
ν3 1 -1 -1 1 = b2

Water has three normal modes that can be grouped together as the reducible representation
Γvib = 2 a1 + b2 . (3)

Determination of normal modes becomes quite complex as the number of atoms in the molecule increases. Nowadays,
computer programs that simulate molecular vibrations can be used to perform these calculations.
The example of [PtCl4]2- shows the increasing complexity. The molecule has five atoms and therefore 15 degrees of freedom,
9 of these are vibrational degrees of freedom. The nine normal modes are exemplified below along with the irreducible
representation the normal mode belongs to (D4h point group).

A1g, b1g and eu are stretching vibrations whereas b2g, a2u, b2u and eu are bending vibrations.

Determining if normal modes are IR and/or Raman active


Transition Moment Integral
A transition from v --> v' is IR active if the transition moment integral contains the totally symmetric irreducible
representation of the point group the molecule belongs to.
The transition moment integral is derived from the one-dimensional harmonic oscillator. Using the definition of dipole
moment the integral is:

If μ, the dipole moment, would be a constant and therefore independent of the vibration, it could be taken outside the integral.
Since v and v' are mutually orthogonal to each other, the integral will equal zero and the transition will not be allowed. In order
for the integral to be nonzero, μ must change during a vibration. This selection rule explains why homonuclear diatomic

6/17/2021 4 https://chem.libretexts.org/@go/page/1856
molecules do not produce an IR spectrum. There is no change in dipole moment resulting in a transition moment integral of
zero and a transition that is forbidden.
For a transition to be Raman active the same rules apply. The transition moment integral must contain the totally symmetric
irreducible representation of the point group. The integral contains the polarizability tensor (usually represented by a square
matrix):

α must be nonzero in order for the transition to be allowed and show Raman scattering.

Character Tables
For a molecule to be IR active the dipole moment has to change during the vibration. For a molecule to be Raman active the
polarizability of the molecule has to change during the vibration. The reducible representation Γvib can also be found by
determining the reducible representation of the 3N degrees of freedom of H2O, Γtot. By applying Group Theory it is
straightforward to find Γx,y,z as well as UMA (number of unmoved atoms). Again, using water as an example with C2v
symmetry where 3N = 9, Γtot can be determined:

C2v E C2 σ (xz) σ (yz)

Τx,y,z 3 -1 1 1

UMA 3 1 1 3
Γtot 9 -1 1 3 =3a1 + a2 + 2b1 + 3b2
Note that Γtot contains nine degrees of freedom consistent with 3N = 9.

Γtot contains Γtranslational, Γrotational as well as Γvibrational. Γtrans can be obtained by finding the irreducible representations
corresponding to x,y and z in the right side of the character table, Γrot by finding the ones corresponding to Rx, Ry and Rz. Γvib
can be obtained by Γtot - Γtrans - Γrot.
Γvib (H2O) = (3a1 + a2 + 2b1+ 3b2) - (a1 + b1 + b2) - (a2 + b1 + b2) = 2a1 + b2
In order to determine which modes are IR active, a simple check of the irreducible representation that corresponds to x,y and z
and a cross check with the reducible representation Γvib is necessary. If they contain the same irreducible representation, the
mode is IR active.
For H2O, z transforms as a1, x as b1 and y as b2. The modes a1 and b2 are IR active since Γvib contains 2a1 + b2.
In order to determine which modes are Raman active, the irreducible representation that corresponds to z2, x2-y2, xy, xz and yz
is used and again cross checked with Γvib. For H2O, z2 and x2-y2 transform as a1, xy as a2, xz as b1 and yz as b2.The modes a1
and b2 are also Raman active since Γvib contains both these modes.
The IR spectrum of H2O does indeed have three bands as predicted by Group Theory. The two symmetric stretches v1 and v2
occur at 3756 and 3657 cm-1 whereas the bending v3 motion occurs at 1595 cm-1.
In order to determine which normal modes are stretching vibrations and which one are bending vibrations, a stretching
analysis can be performed. Then the stretching vibrations can be deducted from the total vibrations in order to obtain the
bending vibrations. A double-headed arrow is drawn between the atom as depicted below:

Then a determination of how the arrows transform under each symmetry operation in C2v symmetry will yield the following
results:

C2v E C2 σ (xz) σ (yz)

Γstretch 2 0 0 2 = a1 + b2

6/17/2021 5 https://chem.libretexts.org/@go/page/1856
Γbend = Γvib - Γstretch = 2a1 + b2 -a1 - b2 = a1
H2O has two stretching vibrations as well as one bending vibration.
This concept can be expanded to complex molecules such as PtCl4-. Four double headed arrows can be drawn between the
atoms of the molecule and determine how these transform in D4h symmetry. Once the irreducible representation for Γstretch has
been worked out, Γbend can be determined by Γbend = Γvib - Γstretch.

Fundamental transition, overtones and hot bands


The transition from v=0 (ground state) -> v=1 (first excited state) is called the fundamental transition. This transition has the
greatest intensity. The transition from v=0 --> v=2 is is referred to as the first overtone, from v=0 --> v=3 is called the second
overtone, etc. Ovetones occur when a mode is excited above the v = 1 level. The harmonic oscillator approximation supports
the prediction that the transition to a second overtone will be twice as energetic as a fundamental transition.
Most molecules are in their zero point energy at room temperature. Therefore, most transitions do originate from the v=0 state.
Some molecules do have a significant population of the v=1 state at room temperature and transitions from this thermally
excited state are called hot bands. Combination bands can occur if more than one vibration is excited by the absorption of a
photon. The overall energy of a combination band is the result of the sum of individual transitions.

References
1. Merlin, J.C., Cornard, J.P., J. Chem. Educ., 2006, 83 (9), p 1383. DOI: 10.1021/ed083p1393
2. McGuinn, C.J., J. Chem. Educ., 1982, 59 (10), p 813. DOI: 10.1021/ed059p813
3. Harris, D.C., Bertolucci, M.D., Symmetry and Spectroscopy: An introduction to Vibrational and Electronic Spectroscopy.
Dover Publocations, Inc., New York, 1989.
4. McQuarrie, D. A., Simon, J.D., Physical Chemistry: A Molecular Approach, University Science Books, Sausalito,
California, 1997; 518-521.
5. Housecroft, C.E., Sharpe, A.G., Inorganic Chemistry. Pearson Education Limited, England, 2008, 107.
6. Atkins, P., dePaula, J., Physical Chemistry, W.H. Freeman and Company, New York, 2002, 520-523.
7. Bishop, D.M., Group Theory and Chemistry, Dover Publications, Inc, New York, 1973, 166.

Problems
1. Chlorophyll a is a green pigment that is found in plants. Its molecular formula is C55H77O5N4Mg. How many degrees of
freedom does this molecule possess? How many vibrational degrees of freedom does it have?
2. CCl4 was commonly used an as organic solvent until its severe carcinogenic properties were discovered. How many
vibrational modes does CCl4 have? Are they IR and/or Raman active?
3. The same vibrational modes in H2O are IR and Raman active. WF6- has IR active modes that are not Raman active and
vice versa. Explain why this is the case.
4. How many IR peaks do you expect from SO3? Estimate where these peaks are positioned in an IR spectrum.
5. Calculate the symmetries of the normal coordinates of planar BF3.

Answers to Problems
1. chlorophyll has 426 degrees of freedom, 420 vibrational modes
2. The point group is Td, Tvib = a1 + e + 2t2, a1 and e are Raman active, t2 is both IR and Raman active
3. For molecules that possess a center of inversion i, modes cannot be simultaneously IR and Raman active
4. Point group is D3h, one would expect three IR active peaks. Asymmetric stretch highest (1391 cm-1), two bending modes
(both around 500 cm-1). The symmetric stretch is IR inactive
5. T3N = A1' + A2' + 3E' + 2 A2" + E" and Tvib= A1' + 2E' + A2"

Contributors and Attributions


Kristin Kowolik, University of California, Davis

6/17/2021 6 https://chem.libretexts.org/@go/page/1856
Number of Vibrational Modes in a Molecule
The Heisenberg uncertainty principle argues that all atoms in a molecule are constantly in motion (otherwise we would know
position and momentum accurately). For molecules, they exhibit three general types of motions: translations (external),
rotations (internal) and vibrations (internal). A diatomic molecule contains only a single motion., while polyatomic molecules
exhibit more complex vibrations, known as normal modes.

Molecular Vibrations
A molecule has translational and rotational motion as a whole while each atom has it's own motion. The vibrational modes can
be IR or Raman active. For a mode to be observed in the IR spectrum, changes must occur in the permanent dipole (i.e. not
diatomic molecules). Diatomic molecules are observed in the Raman spectra but not in the IR spectra. This is due to the fact
that diatomic molecules have one band and no permanent dipole, and therefore one single vibration. An example of this would
be O2 or N2. However, unsymmetric diatomic molecules (i.e. CN) do absorb in the IR spectra. Polyatomic molecules undergo
more complex vibrations that can be summed or resolved into normal modes of vibration.
The normal modes of vibration are: asymmetric, symmetric, wagging, twisting, scissoring, and rocking for polyatomic
molecules.
Symmetric Stretching Asymmetric Stretching Wagging

Twisting Scissoring Rocking

Figure 1 : Six types of Vibrational Modes. Images used with permission (Public Domain; Tiago Becerra Paolini).

Calculate Number of Vibrational Modes


Degree of freedom is the number of variables required to describe the motion of a particle completely. For an atom moving in
3-dimensional space, three coordinates are adequate so its degree of freedom is three. Its motion is purely translational. If we
have a molecule made of N atoms (or ions), the degree of freedom becomes 3N, because each atom has 3 degrees of freedom.
Furthermore, since these atoms are bonded together, all motions are not translational; some become rotational, some others
vibration. For non-linear molecules, all rotational motions can be described in terms of rotations around 3 axes, the rotational
degree of freedom is 3 and the remaining 3N-6 degrees of freedom constitute vibrational motion. For a linear molecule
however, rotation around its own axis is no rotation because it leave the molecule unchanged. So there are only 2 rotational
degrees of freedom for any linear molecule leaving 3N-5 degrees of freedom for vibration.
The degrees of vibrational modes for linear molecules can be calculated using the formula:

3N − 5 (1)

The degrees of freedom for nonlinear molecules can be calculated using the formula:

3N − 6 (2)

n is equal to the number of atoms within the molecule of interest. The following procedure should be followed when trying to
calculate the number of vibrational modes:

7/2/2021 1 https://chem.libretexts.org/@go/page/1857
1. Determine if the molecule is linear or nonlinear (i.e. Draw out molecule using VSEPR). If linear, use Equation 1. If
nonlinear, use Equation 2
2. Calculate how many atoms are in your molecule. This is your N value.
3. Plug in your N value and solve.

Example 1 : Carbon dioxide


How many vibrational modes are there in the linear C O molecule ?
2

Answer
There are a total of 3 atoms in this molecule. It is a linear molecule so we use Equation 1. There are

3(3) − 5 = 4

vibrational modes in C O . 2

Would CO2 and SO2 have a different number for degrees of vibrational freedom? Following the procedure above, it is
clear that CO2 is a linear molecule while SO2 is nonlinear. SO2 contains a lone pair which causes the molecule to be
bent in shape, whereas, CO2 has no lone pairs. It is key to have an understanding of how the molecule is shaped.
Therefore, CO2 has 4 vibrational modes and SO2 has 3 modes of freedom.

Followup (SO2)
Would CO2 and SO2 have a different number for degrees of vibrational freedom? Following the procedure above, it is
clear that CO2 is a linear molecule while SO2 is nonlinear. SO2 contains a lone pair which causes the molecule to be
bent in shape, whereas, CO2 has no lone pairs. It is key to have an understanding of how the molecule is shaped.
Therefore, CO2 has 4 vibrational modes and SO2 has 3 modes of freedom.

Example 2 : Carbon Tetrachloride


How many vibrational modes are there in the tetrahedral C H molecule ? 4

Answer
In this molecule, there are a total of 5 atoms. It is a nonlinear molecule so we use Equation 2. There are

3(5) − 6 = 9

vibrational modes in C H . 4

Example 3 : Buckyballs
How many vibrational modes are there in the nonlinear C 60 molecule ?

Answer
In this molecule, there are a total of 60 carbon atoms. It is a nonlinear molecule so we use Equation 2. There are

3(60) − 6 = 174

vibrational modes in C 0.
6

References
1. Harris, Daniel C., and Michael D. Bertolucci. Symmetry and Spectroscopy: an Introduction to Vibrational and Electronic
Spectroscopy. New York: Dover Publications, 1989. Print.
2. Housecroft, Catherine E., and Alan G. Sharpe. Inorganic Chemistry. Harlow: Pearson Education, 2008. Print.

7/2/2021 2 https://chem.libretexts.org/@go/page/1857
Symmetry Adapted Linear Combinations
The construction of linear combinations of the basis of atomic movements allows the vibrations belonging to irreducible representations to be investigated. The wavefunction of these symmetry
equivalent orbitals is referred to as Symmetry Adapted Linear Combinations, or SALCs. SALCs (Symmetry Adapted Linear Combinations) are the linear combinations of basis sets
composed of the stretching vectors of the molecule. The SALCs of a molecule can help determine binding schemes and symmetries. The procedure used to determine the SALCs of a molecule
is also used to determine the LCAO of a molecule. The LCAO, Linear Combination of Atomic Orbitals, uses the basis set of atomic orbitals instead of stretching vectors. The LCAO of a
molecule provides a detailed description of the molecular orbitals, including the number of nodes and relative energy levels.
Symmetry adapted linear combinations are the sum over all the basis functions:

ϕi = ∑ cij bj (1)

ϕi is the ith SALC function, bj is the jth basis function, and cij is a coefficient which controls how much of bj appears in ϕi . In method two, the projection operator is used to obtain the
coefficients consistent with each irreducible representation.1
The SALCs of a molecule may be constructed in two ways. The first method uses a basis set composed of the irreducible representation of the stretching modes of the molecule. On the other
hand, the second method uses a projection operator on each stretching vector. When determining the irreducible representations of the stretching modes, the reducible representations for all the
vibrational modes must first be determined. Basis vectors are assigned characters and are treated as individual objects. A

Background
In order to understand and construct SALCs, a background in group theory is required. The identification of the point group of the molecule is essential for understanding how the application
of operations affects the molecule. This allows for the determination of the nature of the stretching modes. As a review, let’s first determine the stretching modes of water together. Water has
the point group C2v. Table 1 is the character table for the C2v point group.
Table 1: C2v Character Table
C2v E C2 σv(xz) σv'(yz)

A1 1 1 1 1 z x2, y2, z2

A2 1 1 −1 −1 Rz xy

B1 1 −1 1 −1 x, Ry xz

B2 1 −1 −1 1 y, Rx yz

The first step in determining stretching modes of a molecule is to add the characters contained in the x, y, and z rows to obtain the total reducible representation of the xyz coordinates, ΓXYZ.
ΓXYZ can also be found by applying the symmetry operations to the three vectors (x, y, and z) of the coordinate system of the molecule. The next step involves the investigation of the atoms
that remain unchanged when an operation is applied, ΓUMA. This step refers to the unmoved atoms (UMA). Multiplying ΓXYZ and ΓUMA gives the reducible representation for the molecule
referred to as ΓTOTAL. The ΓTOTAL is the reducible representation for all the modes of the molecule (vibrational, rotational, and translational) and can also be determined by applying the
symmetry operations to each coordinate vector (x, y, and z) on each atom.
Table 2: C2v Reducible Representation for H2O
C2v E C2 σv(xz) σv'(yz)

ΓXYZ 3 -1 1 1

ΓUMA 3 1 1 3

ΓTOTAL 9 -1 1 3

ΓTOTAL is then reduced to later give the stretching modes that are unique to the molecule. First, the reduction formula is applied to decompose the reducible representation:
1 R R
R
ai = ∑(X X C ) (2)
i
h
R

Here, ai is the number of times the irreducible representation will appear in the initial reducible representation. The order of the point group is represented by h; R is an operation of the group;
XR is a character of the operation R in the reducible representation; XiR is a character of the operation R in the irreducible representation, and CR is the number of members in the class to which
R belongs. Applying this formula and subtracting the representations obtained from the basis functions x, y, z, Rx, Ry, and Rz (for the translations and rotations of the molecule) gives the
irreducible representation that corresponds to the vibrational states of the molecule:
ΓVibration = 2a1 + b2
A simple check can be performed to determine that the right number of modes was obtained. For linear molecules (3N-5) gives the correct number of normal modes. For molecules with any
other shape otherwise known as non-linear molecules, the formula is (3N-6). N represents the number of atoms in a molecule. Let’s double check the above water example:
(3N − 6)N = 3 (3)

[3(3) − 6] = 3 (4)

Water should have three vibrational modes. When the irreducible representation was obtained, it was seen that water has two a1 modes and a b2mode for a total of three.
When double checking that you have the correct number of normal modes for other molecules, remember that the irreducible representation E is doubly degenerate and counts as two normal
modes. T is triply degenerate and counts for three normal modes, etc.

Constructing SALCs
Method 1
There are multiple ways of constructing the SALCs of a molecule. The first method uses the known symmetries of the stretching modes of the molecule. To investigate this method, the
construction of the SALCs of water is examined. Water has three vibrational states, 2a1+b2. Two of these vibrations are stretching modes. One is symmetric with the symmetry A1, and the other
is antisymmetric with the symmetry B2. While looking for the SALC of a molecule, one uses vectors represented by bj as the basis set. The vectors demonstrate the irreducible representations
of molecular vibrations.

Figure 1: The stretching modes of H2O.


The SALCs of water can be composed by creating a linear combination of the stretching vectors.

6/17/2021 1 https://chem.libretexts.org/@go/page/1858
ϕ(A1 ) = b1 + b2 (5)

and
ϕ(B1 ) = b1 − b2 (6)

Normalization
The final step in constructing the SALCs of water is to normalize expressions. To normalize the SALC, multiply the entire expression by the normalization constant that is the inverse of the
square root of the sum of the squares of the coefficients within the expression.

ϕi = N ∑ cij bj (7)

1
N = (8)
−−−−−
 n

∑ c2
ij

j=1

1
ϕ(A1 ) = (b1 + b2 ) (9)

√2

and
1
ϕ(B1 ) = (b1 − b2 ) (10)

√2

Normalizing the SALCs ensures that the magnitude of the SALC is unity, and therefore the dot product of any SALC with itself will equal one.

Method 2
The other method for constructing SALCs is the projection operator method. The SALC of a molecule can be constructed in the same manner as the LCAO, Linear Combination of Atomic
Orbitals, however the basis set differs. While looking for the SALCs of a molecule, one uses vectors represented by bj, on the other hand, while looking for the LCAO of a molecule, one uses
atomic orbitals as the basis set. The vectors demonstrate the possible vibration of the molecule. While constructing SALCs, the basis vectors can be treated as individual vectors.

Example 1 : Water
Let’s take a look at how to construct the SALC for water. The first step in constructing the SALC is to label all vectors in the basis set. Below are the bond vectors of water that will be
used as the basis set for the SALCs of the molecule.

Figure 2: Labeled vectors of H2O.


Next, the basis vector, v, is transformed by Tj, the jth symmetry operation of the molecule’s point group. As the vector of the basis set is transformed, record the vector that takes its place.
Water is a member of the point group C2V. The Symmetry elements of the C2V point group are E, C2, σv, σv’.
The ith SALC function, ϕ iis shown below using the vector v=b1.

Figure 3: The transformations of the basis vectors of H2O.


Once the transformations have been determined, the SALC can be constructed by taking the sum of the products of each character of a representation within the point group and the
corresponding transformation. The SALCs functions are the collective transformations of the basis sets represented by ϕ iwhere Xi(j) is the character of the ith irreducible representation
and the jth symmetry operation.

ϕi = ∑ Xi (j)Tj ν (11)

Table 3 : Projection Operator method for C2v


The final step in constructing the SALCs of water is to normalize expressions.

Table 4 : Normalized SALCs of H2O


There are two SALCs for the water molecule, ϕ 1(A1) and ϕ 1(B2). This demonstrates that water has two stretching modes, one is a totally symmetric stretch with the symmetry, A1, and the
other is an antisymmetric stretch with the symmetry B2.

6/17/2021 2 https://chem.libretexts.org/@go/page/1858
Figure 4: The stretching modes of H2O resulting from Method 2.

Interpreting SALCs
Both methods of construction result in the same SALCs. Only irreducible representations corresponding to the symmetries of the stretching modes of the molecule will produce a SALC that is
non-zero. Method 1 only utilized the known symmetries of the vibrational modes. All irreducible representations of the point group were used, but the representations that were not vibrational
modes resulted in SALCs equal to zero.
Therefore, with the SALCs of a molecule given, all the symmetries of the stretching modes are identified. This allows for a clearer understanding of the spectroscopy of the molecule. Even
though vibrational modes can be observed in both infrared and Raman spectroscopy, the SALCs of a molecule cannot identify the magnitude or frequency of the peak in the spectra. The
normalized SALCs can, however, help to determine the relative magnitude of the stretching vectors. The magnitude can be determined by the equation below.
a ⋅ b = |a||b|cosθ (12)

The resulting A1 and B1 symmetries for the above water example are each active in both Raman and IR spectroscopies, according to the C2v character table. If the vibrational mode allows for a
change in the dipole moment, the mode can be observed through infrared spectroscopy. If the vibrational mode allows for a change in the polarization of the molecule, the mode can be
observed through Raman spectroscopy. Both stretching and bending modes are seen in the spectra, however only stretching modes are expressed in the SALCs.

Example 2 : Difluorobenzene
The SALCs of a molecule can also provide insight to the geometry of a molecule. For example, SALCs can aid in determining the differences between para-difluorobenzene and ortho-
difluorobenzene. The SALCs for these two molecules are given below.

Figure 5: para-difluorobenzene
1
ϕ(Ag ) = (b1 + b2 + b3 + b4 ) (13)
2

1
ϕ(B1g ) = (b1 − b2 + b3 − b4 ) (14)
2

1
ϕ(B2u ) = (b1 − b2 − b3 + b4 ) (15)
2

1
ϕ(B3u ) = (b1 + b2 − b3 − b4 ) (16)
2

Figure 6: ortho-difluorobenzene
1
ϕ1 (A1 ) = (b1 + b2 + b3 + b4 ) (17)

√2

1
ϕ2 (A1 ) = (b1 − b2 + b3 − b4 ) (18)

√2

1
ϕ1 (B1 ) = (b1 − b2 − b3 + b4 ) (19)

√2

1
ϕ2 (B1 ) = (b1 + b2 − b3 − b4 ) (20)

√2

From the SALCs, it is seen that para-difluorobenzene has four stretching modes and ortho-difluorobenzene has only two. Therefore, it is no surprise that the vibrational spectroscopy of the
para-difluorobenzene shows more peaks than the ortho-difluorobenzene.

6/17/2021 3 https://chem.libretexts.org/@go/page/1858
Figure 7: FT-IR Raman Spectrum of para-difluorobenzene4

Figure 8: FT-IR Raman Spectrum of ortho-difluorobenzene5

Applications
The SALCs of a molecule can be used to understand the stretching modes and binding schemes of a molecule. More information can also be interpreted when applying the projection operator
used in SALCs on the atomic orbitals of the molecule. This results in the determination of the linear combination of atomic orbitals (LCAO), which gives information on the molecular orbitals
of the molecule. The molecular orbitals (MO) of a molecule are often constructed as LCAOs. Each MO is a solution to the Schrödinger equation and is an eigenfunction of the Hamiltonian
operator. The LCAOs can be determined in the same manner as the SALCs of a molecule, with the use of a projection operator. The difference is that the basis set is no longer stretching
vectors, but instead the atomic orbitals of the molecule. Hydrogen only has s orbitals, but oxygen has s and p orbitals, where the px, py, and pz all transform differently and therefore must be
treated differently.
Once the LCAOs of the molecule have been determined, the expressions can be interpreted into images of the orbitals bonding. If two orbitals are of the same sign in the expression, the
electrons in the orbitals are in phase with each other and are bonding. If two orbitals are of the opposite sign in the expression, the electrons in the orbitals are out of phase with each other and
are antibonding. The image below shows the atomic orbitals' phases (or signs) as red or blue lobes.

Figure 9: Atomic Orbitals of H2O.


Any separation between two antibonding atomic orbitals is a planar node. As the number of nodes increases, so does the level of antibonding. This allows for the LCAO to place the molecular
orbitals in order of increasing energy, which can be used in constructing the molecular orbital (MO) diagram of the molecule. The irreducible representation used to construct the LCAO is used
to describe the MOs. The LCAOs for water are shown below with red dotted lines showing the nodes. Notice the nodes for the px orbitals are in a different plane than the s orbitals of the
hydrogens, so these are degenerate and nonbonding.

6/17/2021 4 https://chem.libretexts.org/@go/page/1858
Figure 10: Molecular Orbital diagram for H2O in order of increasing energy.
Information from the LCAO of water can also be used to analyze and anticipate the adsorption of water onto various surfaces. Evarestov and Bandura used this technique to identify the water
adsorption on Y-doped BaZrO3 and TiO2 (Rutile) respectively.2,3

2
Applying a combination of Methods 1 and 2, the SALCs for CBr2H2 can be determined. The point group of this molecule is C2v, making it similar to the determination of SALCs for
water.

Figure 11: CBr2H2, point group C2v.


However, the central carbon contains more than one type of attached atom; therefore, the stretching analysis must be performed in pieces. First, the C-H stretches are examined, followed
by the C-Br stretches:
Table 5: Irreducible Representations for C-H and C-Br stretches in CBr2H2.
C2V E C2 σV σV’

ΓC-H 2 0 0 2

ΓC-Br 2 0 2 0

Applying the projection operator method to C-Br and C-H stretches individually, the SALCs are obtained in the same fashion as before.
Table 6: SALCs for CBr2H2.
ΓC-Br E C2 σV σV’ SUM

A1Tj(b1) b1 b2 b1 b2 2(b1 + b2)

B1Tj(b1) b1 -b2 b1 -b2 2(b1 - b2)

Table 7: SALCs for CBr2H2.


ΓC-H E C2 σV σV’ SUM

A1Tj(a1) a1 a2 a2 a1 2(a1 + a2)

B2Tj(a1) a1 -a2 -a2 a1 2(a1 - a2)

The results are normalized and the following SALCs are obtained for the C2v molecule CBr2H2:
1
ϕC Br(A1 ) = (b1 + b2 ) (21)

√2

1
ϕC Br(B1 ) = (b1 − b2 ) (22)

√2

1
ϕC H (A1 ) = – (a1 + a2 ) (23)
√2

6/17/2021 5 https://chem.libretexts.org/@go/page/1858
1
ϕC H (B2 ) = – (a1 − a2 ) (24)
√2

4
To obtain the SALCs for PtCl4, the same general method is applied.

Figure 12: PtCl4, point group D4h.


However, even though the point group of the molecule is D4h, the cyclic subgroup C4 may be used (this is a more simplified character table used for spherically symmetrical molecules).
Some manipulation is required in order to use this cyclic subgroup and will be discussed. Below is the C4 cyclic character table.
Table 8: C4 cyclic character table.
C4 E C41 C42 C43

A 1 1 1 1

B 1 -1 1 -1

E1 E2 11 i -i -1 -1 -i i

Notice, there are two rows for E, each singly degenerate. To solve for the characters of E, one must take the sum and difference of the two rows. Then, a reduction can be applied to obtain
the easiest possible characters by dividing each row by a common factor (removing the common factor is not necessary, but it does simplify the problem as well as remove any imaginary
terms):
Sum = [ (1+1) (i-i) (-1-1) (-i+i) ] = (2 0 -2 0) ÷ 2 = E1 (1 0 -1 0)
Difference = [ (1-1) (i+i) (-1+1) (-i-i) ] = (0 2i 0 -2i) ÷ 2i = E2 (0 1 0 -1)
Using the above cyclic group, and the newly obtained characters for E, the projection operator can be applied using Method 2 for the construction of SALCs.
Table 9: SALCs for PtCl4 using Method 2.
C4 E C41 C42 C43 SUM

ATj(b1) b1 b2 b3 b4 b1 + b2 + b3 + b4

BTj(b1) b1 -b2 b3 -b4 b1 - b2 + b3 - b4

E1Tj(b1) b1 0 -b3 0 b1 - b3

E2Tj(b1) 0 b2 0 -b4 b2 - b4

Normalizing the sum as mentioned in Method 1, the following SALCs are obtained for the D4h molecule PtCl4:
1
ϕ(A) = (b1 + b2 + b3 + b4 ) (25)
2

1
ϕ(B) = (b1 − b2 + b3 − b4 ) (26)
2

1
1
ϕ(E ) = (b1 − b3 ) (27)

√2

1
2
ϕ(E ) = (b2 − b4 ) (28)

√2

3
Applying a combination of Methods 1 and 2, the SALCs for the C-H stretches of PF2H3 can be determined. The point group of this molecule is Cs. The central carbon contains more than
one type of attached hydrogen; therefore, the stretching analysis must be performed in pieces. First, the C-HA stretches are examined, followed by the C-HB stretches:

Figure 13: PF2H3, point group Cs.


Table 10: Irreducible Representations for C-Ha and C-Hb stretches in PF2H3.
Cs E σh Irreducible Representation

ΓC-Ha 2 0 ΓC-Ha = A’ + A”

ΓC-Hb 1 1 ΓC-Hb = A’

Applying the projection operator method to C-HA and C-HB stretches individually, the SALCs are obtained in the same fashion as before.
Table 11: SALCs for PF2H3 using Method 2.
ΓC-HB E σh SUM

A’ Tj(b1) b1 b2 b1 + b2

A” Tj(b1) b1 -b2 b1 - b2

A’ Tj(a1) a1 a1 a1 + a1

Table 12: SALCs for PF2H3 using Method 2.

6/17/2021 6 https://chem.libretexts.org/@go/page/1858
ΓC-HA E C2 σV σV’ SUM

A1Tj(a1) a1 a2 a2 a1 2(a1+ a2)

B2Tj(a1) a1 -a2 -a2 a1 2(a1- a2)

The results are normalized and the following SALCs are obtained for the Cs molecule PF2H3:


1 1
ϕ1 A = – (b1 + b2 ) + a1 = (b1 + b2 + a1 ) (29)
√2 sqrt3


1 1
ϕ2 A = – (b1 + b2 ) − a1 = (b1 + b2 − a1 ) (30)
√2 sqrt3

′′
1
ϕA = – (b1 − b2 ) (31)
√2

Problems
1. Construct the SALCs for C-H stretches of ortho-difluorobenzene.
2. Construct the SALCs for ammonia.
3. Draw the nodes for the MOs of BeH2 (determined by used of LCAOs) and rank the MOs in order of increasing energy.

References
1. Willock, David J. Molecular Symmetry. Chichester, U.K: Wiley, 2009.
2. R.A. Evarestov, A.V. Bandura, LCAO calculation of water adsorption on (001) surface of Y-doped BaZrO3, Solid State Ionics, In Press, Corrected Proof, Available online 2 October 2010,
ISSN 0167-2738, DOI: 10.1016/j.ssi.2010.09.012.
(www.sciencedirect.com/science...8ab4e49bd4429e)
3. A. V. Bandura,, D. G. Sykes,, V. Shapovalov,, T. N. Troung,, J. D. Kubicki,, and, R. A. Evarestov. Adsorption of Water on the TiO2 (Rutile) (110) Surface: A Comparison of Periodic and
Embedded Cluster Calculations. The Journal of Physical Chemistry B 2004 108 (23), 7844-7853.
4. "1,4-Difluorobenzene" Sigma Aldrich. www.sigmaaldrich.com/catalog/...0&QS=ON&F=SPEC. 14 March, 2011.
5. "1,2-Difluorobenzene" Sigma Aldrich. www.sigmaaldrich.com/catalog/...AND_KEY&F=SPEC. 14 March, 2011.

Solutions to Practice Problems


1. A1Tj(b1)= \frac{1}{\sqrt{2}} (b_{1} +b_{2})\)
B2Tj(b1)= \frac{1}{\sqrt{2}} (b_{1} -b_{2})\)
A1Tj(b3)= \frac{1}{\sqrt{2}} (b_{3} +b_{4})\)
B2Tj(b3)= \frac{1}{\sqrt{2}} (b_{3} -b_{4})\)
Then add and subtract (for in phase and out of phase) the individual linear combinations found by the projection operator to give the SALCs.
ϕ 1(A1) = \frac{1}{2} (b_{1} +b_{2} +b_{3} +b_{4} )\)
ϕ 2(A1) = \frac{1}{2} (b_{1} +b_{2} -b_{3} -b_{4} )\)
ϕ 1(B2) = \frac{1}{2} (b_{1} -b_{2} +b_{3} -b_{4} )\)
ϕ 2(B2) = \frac{1}{2} (b_{1} -b_{2} -b_{3} +b_{4} )\).
2. ATj(b1)= \frac{1}{\sqrt{3}} (b_{1} +b_{2} +b_{3} )\)
E1Tj(b1)= \frac{1}{\sqrt{6}} (2b_{1} -b_{2} -b_{3} )\)
E2Tj(b1)= \frac{1}{\sqrt{2}} (b_{2} -b_{3} )\)
3.

Contributors and Attributions


Jorie Fields, Michelle Faust

6/17/2021 7 https://chem.libretexts.org/@go/page/1858
CHAPTER OVERVIEW
INFRARED SPECTROSCOPY
Infrared Spectroscopy is the analysis of infrared light interacting with a molecule. This can be
analyzed in three ways by measuring absorption, emission and reflection. The main use of this
technique is in organic and inorganic chemistry. It is used by chemists to determine functional
groups in molecules. IR Spectroscopy measures the vibrations of atoms, and based on this it is
possible to determine the functional groups.

HOW AN FTIR SPECTROMETER OPERATES


FTIR spectrometers (Fourier Transform Infrared Spectrometer) are widely used in organic
synthesis, polymer science, petrochemical engineering, pharmaceutical industry and food analysis.
In addition, since FTIR spectrometers can be hyphenated to chromatography, the mechanism of
chemical reactions and the detection of unstable substances can be investigated with such instruments.

IDENTIFYING THE PRESENCE OF PARTICULAR GROUPS


This page explains how to use an infra-red spectrum to identify the presence of a few simple bonds in organic compounds.

INFRARED: APPLICATION
Infrared spectroscopy, an analytical technique that takes advantage of the vibrational transitions of a molecule, has been of great
significance to scientific researchers in many fields such as protein characterization, nanoscale semiconductor analysis and space
exploration.

INFRARED: INTERPRETATION
Infrared spectroscopy is the study of the interaction of infrared light with matter. The fundamental measurement obtained in infrared
spectroscopy is an infrared spectrum, which is a plot of measured infrared intensity versus wavelength (or frequency) of light.

INFRARED SPECTROSCOPY
Infrared (IR) spectroscopy is one of the most common and widely used spectroscopic techniques employed mainly by inorganic and
organic chemists due to its usefulness in determining structures of compounds and identifying them. Chemical compounds have
different chemical properties due to the presence of different functional groups.

INTERPRETING INFRARED SPECTRA


This chapter will focus on infrared (IR) spectroscopy. The wavelengths found in infrared radiation are a little longer than those found
in visible light. IR spectroscopy is useful for finding out what kinds of bonds are present in a molecule, and knowing what kinds of
bonds are present is a good start towards knowing what the structure could be.

CARBON NITROGEN BONDS


IR10. MORE PRACTICE WITH IR SPECTRA
IR11. APPENDIX: IR TABLE OF ORGANIC COMPOUNDS
IR2. HYDROCARBON SPECTRA
IR3. SUBTLE POINTS OF IR SPECTROSCOPY
IR4. CARBON CARBON MULTIPLE BONDS
IR5. CARBON OXYGEN SINGLE BONDS
IR6. CARBON OXYGEN DOUBLE BONDS
IR8. MORE COMPLICATED IR SPECTRA
IR9. MISLEADING PEAKS
WHAT DOES AN IR SPECTRUM LOOK LIKE?
IR SPECTROSCOPY BACKGROUND
THE FINGERPRINT REGION
The fingerprint region is the region to the right-hand side of the diagram (from about 1500 to 500 cm-1) usually contains a very
complicated series of absorptions. These are mainly due to all manner of bending vibrations within the molecule.

BACK MATTER

INDEX

1 7/7/2021
How an FTIR Spectrometer Operates
FTIR spectrometers (Fourier Transform Infrared Spectrometer) are widely used in organic synthesis, polymer science,
petrochemical engineering, pharmaceutical industry and food analysis. In addition, since FTIR spectrometers can be
hyphenated to chromatography, the mechanism of chemical reactions and the detection of unstable substances can be
investigated with such instruments.

Introduction
The range of Infrared region is 12800 ~ 10 cm-1 and can be divided into near-infrared region (12800 ~ 4000 cm-1), mid-
infrared region (4000 ~ 200 cm-1) and far-infrared region (50 ~ 1000 cm-1). The discovery of infrared light can be dated back
to the 19th century. Since then, scientists have established various ways to utilize infrared light. Infrared absorption
spectroscopy is the method which scientists use to determine the structures of molecules with the molecules’ characteristic
absorption of infrared radiation. Infrared spectrum is molecular vibrational spectrum. When exposed to infrared radiation,
sample molecules selectively absorb radiation of specific wavelengths which causes the change of dipole moment of sample
molecules. Consequently, the vibrational energy levels of sample molecules transfer from ground state to excited state. The
frequency of the absorption peak is determined by the vibrational energy gap. The number of absorption peaks is related to the
number of vibrational freedom of the molecule. The intensity of absorption peaks is related to the change of dipole moment
and the possibility of the transition of energy levels. Therefore, by analyzing the infrared spectrum, one can readily obtain
abundant structure information of a molecule. Most molecules are infrared active except for several homonuclear diatomic
molecules such as O2, N2 and Cl2 due to the zero dipole change in the vibration and rotation of these molecules. What makes
infrared absorption spectroscopy even more useful is the fact that it is capable to analyze all gas, liquid and solid samples. The
common used region for infrared absorption spectroscopy is 4000 ~ 400 cm-1 because the absorption radiation of most organic
compounds and inorganic ions is within this region.
FTIR spectrometers are the third generation infrared spectrometer. FTIR spectrometers have several prominent advantages: (1)
The signal-to-noise ratio of spectrum is significantly higher than the previous generation infrared spectrometers. (2) The
accuracy of wavenumber is high. The error is within the range of ± 0.01 cm-1. (3) The scan time of all frequencies is short
(approximately 1 s). (4) The resolution is extremely high (0.1 ~ 0.005 cm-1). (5) The scan range is wide (1000 ~ 10 cm-1). (6)
The interference from stray light is reduced. Due to these advantages, FTIR Spectrometers have replaced dispersive IR
spectrometers.

Development of IR Spectrometers
Up till FTIR spectrometers, there have been three generations of IR spectrometers.
1. The first generation IR spectrometer was invented in late 1950s. It utilizes prism optical splitting system. The prisms are
made of NaCl. The requirement of the sample’s water content and particle size is extremely strict. Further more, the scan
range is narrow. Additionally, the repeatability is fairly poor. As a result, the first generation IR spectrometer is no longer in
use.
2. The second generation IR spectrometer was introduced to the world in 1960s. It utilizes gratings as the monochrometer.
The performance of the second generation IR spectrometer is much better compared with IR spectrometers with prism
monochrometer, But there are still several prominent weaknesses such as low sensitivity, low scan speed and poor
wavelength accuracy which rendered it out of date after the invention of the third generation IR spectrometer.
3. The invention of the third generation IR spectrometer, Fourier transform infrared spectrometer, marked the abdication of
monochrometer and the prosperity of interferometer. With this replacement, IR spectrometers became exceptionally
powerful. Consequently, various applications of IR spectrometer have been realized.

Dispersive IR Spectrometers
To understand the powerfulness and usefulness of FTIR spectrometer, it is essential to have some background information of
dispersive IR Spectrometer. The basic components of a dispersive IR spectrometer include a radiation source, monochromator,
and detector. The common IR radiation sources are inert solids that are heated electrically to promote thermal emission of
radiation in the infrared region of the electromagnetic spectrum. The monochromator is a device used to disperse or separate a
broad spectrum of IR radiation into individual narrow IR frequencies.

6/23/2021 1 CC-BY https://chem.libretexts.org/@go/page/1844


Generally, dispersive spectrometers have a double-beam design with two equivalent beams from the same source passing
through the sample and reference chambers as independent beams. These reference and sample beams are alternately focused
on the detector by making use of an optical chopper, such as, a sector mirror. One beam will proceed, traveling through the
sample, while the other beam will pass through a reference species for analytical comparison of transmitted photon wavefront
information.
After the incident radiation travels through the sample species, the emitted wavefront of radiation is dispersed by a
monochromator (gratings and slits) into its component frequencies. A combination of prisms or gratings with variable-slit
mechanisms, mirrors, and filters comprise the dispersive system. Narrower slits gives better resolution by distinguishing more
closely spaced frequencies of radiation and wider slits allow more light to reach the detector and provide better system
sensitivity. The emitted wavefront beam (analog spectral output) hits the detector and generates an electrical signal as a
response.
Detectors are devices that convert the analog spectral output into an electrical signal. These electrical signals are further
processed by the computer using mathematical algorithm to arrive at the final spectrum. The detectors used in IR
spectrometers can be classified as either photon/quantum detectors or thermal detectors.
It is the absorption of IR radiation by the sample, producing a change of IR radiation intensity, which gets detected as an off-
null signal (e.g. different from reference signal). This change is translated into the recorder response through the actions of
synchronous motors. Each frequency that passes through the sample is measured individually by the detector which
consequently slows the process of scanning the entire IR region. A block diagram of a classic dispersive IR spectrometer is
shown in Figure 1.

Figure 1. Simplified representation of a dispersive IR spectrometer.

FTIR Spectrometers
The Components of FTIR Spectrometers
A common FTIR spectrometer consists of a source, interferometer, sample compartment, detector, amplifier, A/D convertor,
and a computer. The source generates radiation which passes the sample through the interferometer and reaches the detector.
Then the signal is amplified and converted to digital signal by the amplifier and analog-to-digital converter, respectively.
Eventually, the signal is transferred to a computer in which Fourier transform is carried out. Figure 2 is a block diagram of an
FTIR spectrometer.

Figure 2. Block diagram of an FTIR spectrometer


The major difference between an FTIR spectrometer and a dispersive IR spectrometer is the Michelson interferometer.

Michelson Interferometer
The Michelson interferometer, which is the core of FTIR spectrometers, is used to split one beam of light into two so that the
paths of the two beams are different. Then the Michelson interferometer recombines the two beams and conducts them into the
detector where the difference of the intensity of these two beams are measured as a function of the difference of the paths.
Figure 3 is a schematic of the Michelson Interferometer.

6/23/2021 2 CC-BY https://chem.libretexts.org/@go/page/1844


Figure 3. Schematic of the Michelson interferometer

A typical Michelson interferometer consists of two perpendicular mirrors and a beamsplitter. One of the mirror is a stationary
mirror and another one is a movable mirror. The beamsplitter is designed to transmit half of the light and reflect half of the
light. Subsequently, the transmitted light and the reflected light strike the stationary mirror and the movable mirror,
respectively. When reflected back by the mirrors, two beams of light recombine with each other at the beamsplitter.
If the distances travelled by two beams are the same which means the distances between two mirrors and the beamsplitter are
the same, the situation is defined as zero path difference (ZPD). But imagine if the movable mirror moves away from the
beamsplitter, the light beam which strikes the movable mirror will travel a longer distance than the light beam which strikes
the stationary mirror. The distance which the movable mirror is away from the ZPD is defined as the mirror displacement and
is represented by ∆. It is obvious that the extra distance travelled by the light which strikes the movable mirror is 2∆. The extra
distance is defined as the optical path difference (OPD) and is represented by delta. Therefore,
δ = 2Δ (1)

It is well established that when OPD is the multiples of the wavelength, constructive interference occurs because crests overlap
with crests, troughs with troughs. As a result, a maximum intensity signal is observed by the detector. This situation can be
described by the following equation:
δ = nλ (2)

with n = 0,1,2,3...
In contrast, when OPD is the half wavelength or half wavelength add multiples of wavelength, destructive interference occurs
because crests overlap with troughs. Consequently, a minimum intensity signal is observed by the detector. This situation can
be described by the following equation:
1
δ = (n + )λ (3)
2

with n = 0,1,2,3...
These two situations are two extreme situations. If the OPD is neither n-fold wavelengths nor (n+1/2)-fold wavelengths, the
interference should be between constructive and destructive. So the intensity of the signal should be between maximum and
minimum. Since the mirror moves back and forth, the intensity of the signal increases and decreases which gives rise to a
cosine wave. The plot is defined as an interferogram. When detecting the radiation of a broad band source rather than a single-
wavelength source, a peak at ZPD is found in the interferogram. At the other distance scanned, the signal decays quickly since
the mirror moves back and forth. Figure 4(a) shows an interferogram of a broad band source.

Fourier Transform of Interferogram to Spectrum


The interferogram is a function of time and the values outputted by this function of time are said to make up the time domain.
The time domain is Fourier transformed to get a frequency domain, which is deconvolved to product a spectrum. Figure 4
shows the Fast Fourier transform from an interferogram of polychromatic light to its spectrum.

6/23/2021 3 CC-BY https://chem.libretexts.org/@go/page/1844


Figure 4. (a) Interferogram of a monochromatic light; (b) its spectrum

The Fourier Transform


The first one who found that a spectrum and its interferogram are related via a Fourier transform was Lord Rayleigh. He made
the discover in 1892. But the first one who successfully converted an interferogram to its spectrum was Fellgett who made the
accomplishment after more than half a century. Fast Fourier transform method on which the modern FTIR spectrometer based
was introduced to the world by Cooley and Turkey in 1965. It has been applied widely to analytical methods such as infrared
spectrometry, nuclear magnetic resonance and mass spectrometry due to several prominent advantages which are listed in
Table 1 .

Table 1 . Advantages of Fourier Transform over Continuous-Wave Spectrometry


Fourier transform, named after the French mathematician and physicist Jean Baptiste Joseph Fourier, is a mathematical method
to transform a function into a new function. The following equation is a common form of the Fourier transform with unitary
normalization constants:

1 −iωt
F (ω) = ∫ f (t)e dt (4)
−−
√2π −∞

in which t is time, i is the square root of -1.


The following equation is another form of the Fourier transform(cosine transform) which applies to real, even functions:

1
F (ν ) = ∫ f (t) cos(2πν t)dt (5)
−−
√2π −∞

The following equation shows how f(t) is related to F(v) via a Fourier transform:

1
f (t) = ∫ F (ν ) cos(2πν t)dν (6)
−−
√2π −∞

An Alternative Explanation of the Fourier Transform in FTIR Spectrometers


The math description of the Fourier transform can be tedious and confusing. An alternative explanation of the Fourier
transform in FTIR spectrometers is provided here before we jump into the math description to give you a rough impression
which may help you understand the math description.
The interferogram obtained is a plot of the intensity of signal versus OPD. A Fourier transform can be viewed as the inversion
of the independent variable of a function. Thus, Fourier transform of the interferogram can be viewed as the inversion of OPD.
The unit of OPD is centimeter, so the inversion of OPD has a unit of inverse centimeters, cm-1. Inverse centimeters are also

6/23/2021 4 CC-BY https://chem.libretexts.org/@go/page/1844


known as wavenumbers. After the Fourier transform, a plot of intensity of signal versus wavenumber is produced. Such a plot
is an IR spectrum. Although this explanation is easy to understand, it is not perfectly rigorous.
Simplified Math Description of the Fourier Transform in FTIR
The wave functions of the reflected and transmitted beams may be represented by the general form of:
E1 = rtc Em × cos(ν t − 2πkx) (7)

and
E1 = rtc Em × cos[ν t − 2πk(ν x + Δd)] (8)

where
Δd is the path difference,
r is the reflectance (amplitude) of the beam splitter,
t is the transmittance, and

c is the polarization constant.

The resultant wave function of their superposition at the detector is represented as:

E = E1 + E2 = 2(r × t × c × Em ) × cos(ν t − 2πkx) cos(πkΔd) (9)

where Em,, ν, and k are the amplitude, frequency and wave number of the IR radiation source.
The intensity (I ) detected is the time average of E and is written as
2

2 2 2 2 2 2
I = 4 r t c Em cos (ν t − 2πkx) cos (πkΔd) (10)

Since the time average of the first cosine term is just ½, then
2
I = 2I (k) cos (πkΔd) (11)

and

I (Δd) = I (k)[1 + cos(2πkΔd)] (12)

where I (k) is a constant that depends only upon k and I (Δd) is the interferogram.
From I (Δd) we can get I (k) using Fourier transform as follows:
km

I (Δd) − I (∞) = ∫ I (k) cos(2ΠkΔd)dk (13)


0

Letting Km →∞, we can write


I (k) = ∫ [I (Δd) − I (∞)] cos(2ΠkΔd)dΔd (14)


0

The physically measured information recorded at the detector produces an interferogram, which provides information about a
response change over time within the mirror scan distance. Therefore, the interferogram obtained at the detector is a time
domain spectrum. This procedure involves sampling each position, which can take a long time if the signal is small and the
number of frequencies being sampled is large.
In terms of ordinary frequency, ν , the Fourier transform of this is given by (angular frequency ω = sπν ):

−i2Πν t
f (ν ) = ∫ f (t)e dt (15)
−∞

The inverse Fourier transform is given by:



+i2πν t
f (ν ) = ∫ f (t)e dt (16)
−∞

6/23/2021 5 CC-BY https://chem.libretexts.org/@go/page/1844


The interferogram is transformed into IR absorption spectrum (Figure 5) that is commonly recognizable with absorption
intensity or % transmittance plotted against the wavelength or wavenumber. The ratio of radiant power transmitted by the
sample (I) relative to the radiant power of incident light on the sample (I0) results in quantity of Transmittance, (T).
Absorbance (A) is the logarithm to the base 10 of the reciprocal of the transmittance (T):
1 I
A = log10 = −log10 T = −log10 (17)
T I0

Figure 5. IR spectrum of a sample

Hands-on Operation of an FTIR Spectrometer


Step 1: The first step is sample preparation. The standard method to prepare solid sample for FTIR spectrometer is to use
KBr. About 2 mg of sample and 200 mg KBr are dried and ground. The particle size should be unified and less than two
micrometers. Then, the mixture is squeezed to form transparent pellets which can be measured directly. For liquids with high
boiling point or viscous solution, it can be added in between two NaCl pellets. Then the sample is fixed in the cell by skews
and measured. For volatile liquid sample, it is dissolved in CS2 or CCl4 to form 10% solution. Then the solution is injected
into a liquid cell for measurement. Gas sample needs to be measured in a gas cell with two KBr windows on each side. The
gas cell should first be vacuumed. Then the sample can be introduced to the gas cell for measurement.
Step 2: The second step is getting a background spectrum by collecting an interferogram and its subsequent conversion to
frequency data by inverse Fourier transform. We obtain the background spectrum because the solvent in which we place our
sample will have traces of dissolved gases as well as solvent molecules that contribute information that are not our sample.
The background spectrum will contain information about the species of gases and solvent molecules, which may then be
subtracted away from our sample spectrum in order to gain information about just the sample. Figure 6 shows an example of
an FTIR background spectrum.

Figure 6. Background IR spectrum


The background spectrum also takes into account several other factors related to the instrument performance, which includes
information about the source, interferometer, detector, and the contribution of ambient water (note the two irregular groups of
lines at about 3600 cm–1 and about 1600 cm–1 in Figure 6) and carbon dioxide (note the doublet at 2360 cm–1 and sharp spike
at 667 cm–1 in Figure 6) present in the optical bench.
Step 3: Next, we collect a single-beam spectrum of the sample, which will contain absorption bands from the sample as well
as the background (gaseous or solvent).

6/23/2021 6 CC-BY https://chem.libretexts.org/@go/page/1844


Step 4: The ratio between the single-beam sample spectrum and the single beam background spectrum gives the spectrum of
the sample (Figure 7).

Figure 7. Sample IR spectrum


Step 5: Data analysis is done by assigning the observed absorption frequency bands in the sample spectrum to appropriate
normal modes of vibrations in the molecules.

Portable FTIR Spectrometers


Despite of the powerfulness of traditional FTIR spectrometers, they are not suitable for real-time monitoring or field use. So
various portable FTIR spectrometers have been developed. Below are two examples.
Ahonen et al developed a portable, real-time FTIR spectrometer as a gas analyzer for industrial hygiene use. The instrument
consists of an operational keyboard, a control panel, signal and control processing electronics, an interferometer, a heatable
sample cell and a detector. All the components were packed into a cart. To minimize the size of the instrument, the resolution
of FTIR spectrometer was sacraficed. But it is good enough for the use of industrial hygiene. The correlation coefficient of
hygienic effect between the analyzer and adsorption tubes is about 1 mg/m3.
Korb et al developed a portable FTIR spectrometer which only weighs about 12.5 kg so that it can be held by hand. Moreover,
the energy source of the instrument is battery so that the mobility is significantly enhanced. Besides, the instrument can
function well within the temperature range of 0 to 45 oC and the humidity range of 0 to 100%. Additionally, this instrument
resists vibration. It works well in an operating helicopter. Consequently, this instrument is excellent for the analysis of
radiation from the surface and atmosphere of the Earth. The instrument is also very stable. After a three-year operation, it did
not lose optical alignment. The reduction of size was implemented by a creative design of optical system and accessory
components. Two KBr prisms were used to constitute the interferometer cavity. Optical coatings replaced the mirrors and
beam splitter in the interferometer. The optical path is shortened with a much more compact packaging of components. A
small, low energy consuming interferometer drive was designed. It is also mass balanced to resist vibration. The common He-
Ne tube was replaced by a smaller laser diode.

References
1. P.R. Griffiths, Science, 21, 1983, 297
2. W.D. Perkins, "Fourier Transform-Infrared Spectroscopy”. Part 1. Instrumentation. Topics in Chemical Instrumentation.
Ed. Frank A. Settle, Jr. Journal of Chemical Education, 63:1, January 1986: A5-A10.
3. D.A. Skoog and J.J. Leary. “Principles of Instrumental Analysis, 4th Ed.”, Harcourt Brace Jovanovich. Philadelphia, PA,
1992. Chapter 12.
4. F. Daniels, J.W. Williams, P. Bender, R.A. Alberty, C.D. Cornwell, J. E. Harriman. "Experimental Physical Chemistry, 7th
Ed.”, McGraw-Hill, New York, NY, 1970.
5. J.W. Cooley and J.W. Turkey, Math. Comp., 1965, 19, 9
6. A.G. Marshall, Acc. Chem. Res., 1985, 18, 316
7. A. R. Korb, P. Dybwad, W. Wadsworth, J. W. Salisbury, Applied Optics, 1996, 35, 1679
8. I. Ahonen, H. Riipinen, A. Roos, Analyst, 1996, 121, 1253
9. D. W. Ball, Field Guide to Spectroscopy, SPIE Publication, Bellingham, 2006
10. V. Saptari, Fourier-Transform Spectroscopy Instrumentation Engineering, SPIE Publication, Bellingham, 2003
11. P. R. Griffiths, J. A. de Haseth, Fourier Transform Infrared Spectrometry, Wiley, New York, 1986
12. B. C. Smith, Fundamentals of Fourier Transform Infrared Spectroscopy, CRC press, 1996
13. B. Stuart, Modern Infrared Spectroscopy, Wiley, New York, 1996

6/23/2021 7 CC-BY https://chem.libretexts.org/@go/page/1844


14. A. L. Smith, Applied Infrared Spectroscopy : Fundamentals, Techniques, and Analytical Problem-solving, Wiley, New
York, 1979
15. P. R. Griffiths, Chemical Infrared Fourier Transform Spectroscopy, Wiley, New York, 1975

Contributors and Attributions


Nancy Birkner (UCD), Qian Wang (UCD)

6/23/2021 8 CC-BY https://chem.libretexts.org/@go/page/1844


Identifying the Presence of Particular Groups
This page explains how to use an infra-red spectrum to identify the presence of a few simple bonds in organic compounds.

The infrared spectrum for a simple carboxylic acid: Ethanoic acid


Ethanoic acid has the structure:

You will see that it contains the following bonds:


carbon-oxygen double, C=O
carbon-oxygen single, C-O
oxygen-hydrogen, O-H
carbon-hydrogen, C-H
carbon-carbon single, C-C
The carbon-carbon bond has absorptions which occur over a wide range of wavenumbers in the fingerprint region - that makes
it very difficult to pick out on an infra-red spectrum. The carbon-oxygen single bond also has an absorbtion in the fingerprint
region, varying between 1000 and 1300 cm-1 depending on the molecule it is in. You have to be very wary about picking out a
particular trough as being due to a C-O bond.
The other bonds in ethanoic acid have easily recognized absorptions outside the fingerprint region.
The C-H bond (where the hydrogen is attached to a carbon which is singly-bonded to everything else) absorbs somewhere
in the range from 2853 - 2962 cm-1. Because that bond is present in most organic compounds, that's not terribly useful!
What it means is that you can ignore a trough just under 3000 cm-1, because that is probably just due to C-H bonds.
The carbon-oxygen double bond, C=O, is one of the really useful absorptions, found in the range 1680 - 1750 cm-1. Its
position varies slightly depending on what sort of compound it is in.
The other really useful bond is the O-H bond. This absorbs differently depending on its environment. It is easily recognised
in an acid because it produces a very broad trough in the range 2500 - 3300 cm-1.
The infrared spectrum for ethanoic acid looks like this:

The possible absorption due to the C-O single bond is queried because it lies in the fingerprint region. You couldn't be sure that
this trough wasn't caused by something else.

The infrared spectrum for an alcohol: Ethanol

Jim Clark 6/25/2021 1 https://chem.libretexts.org/@go/page/3734


The O-H bond in an alcohol absorbs at a higher wavenumber than it does in an acid - somewhere between 3230 - 3550 cm-1.
In fact this absorption would be at a higher number still if the alcohol isn't hydrogen bonded - for example, in the gas state. All
the infra-red spectra on this page are from liquids - so that possibility will never apply.

Notice the absorption due to the C-H bonds just under 3000 cm-1, and also the
troughs between 1000 and 1100 cm-1 - one of which will be due to the C-O bond.

The infrared spectrum for an ester: Ethyl ethanoate

This time the O-H absorption is missing completely. Don't confuse it with the C-H trough fractionally less than 3000 cm-1. The
presence of the C=O double bond is seen at about 1740 cm-1.
The C-O single bond is the absorption at about 1240 cm-1. Whether or not you could pick that out would depend on the detail
given by the table of data which you get in your exam, because C-O single bonds vary anywhere between 1000 and 1300 cm-1
depending on what sort of compound they are in. Some tables of data fine it down, so that they will tell you that an absorption
from 1230 - 1250 is the C-O bond in an ethanoate.

The infrared spectrum for a ketone: Propanone

You will find that this is very similar to the infra-red spectrum for ethyl ethanoate, an ester. Again, there is no trough due to the
O-H bond, and again there is a marked absorption at about 1700 cm-1 due to the C=O.
Confusingly, there are also absorptions which look as if they might be due to C-O single bonds - which, of course, aren't
present in propanone. This reinforces the care you have to take in trying to identify any absorptions in the fingerprint region.
Aldehydes will have similar infra-red spectra to ketones.

Jim Clark 6/25/2021 2 https://chem.libretexts.org/@go/page/3734


The infrared spectrum for a hydroxy-acid: 2-hydroxypropanoic acid (lactic acid)

This is interesting because it contains two different sorts of O-H bond - the one in the acid and the simple "alcohol" type in the
chain attached to the -COOH group.
The O-H bond in the acid group absorbs between 2500 and 3300, the one in the chain between 3230 and 3550 cm-1. Taken
together, that gives this immense trough covering the whole range from 2500 to 3550 cm-1. Lost in that trough as well will be
absorptions due to the C-H bonds. Notice also the presence of the strong C=O absorption at about 1730 cm-1.

The infrared spectrum for a primary amine: 1-aminobutane

Primary amines contain the -NH2 group, and so have N-H bonds. These absorb somewhere between 3100 and 3500 cm-1. That
double trough (typical of primary amines) can be seen clearly on the spectrum to the left of the C-H absorptions.

Contributors and Attributions


Jim Clark (Chemguide.co.uk)

Jim Clark 6/25/2021 3 https://chem.libretexts.org/@go/page/3734


Infrared: Application
Infrared spectroscopy, an analytical technique that takes advantage of the vibrational transitions of a molecule, has been of
great significance to scientific researchers in many fields such as protein characterization, nanoscale semiconductor analysis
and space exploration.

Introduction
Infrared spectroscopy is the study of interaction of infrared light with matter, which can be used to identify unknown materials,
examine the quality of a sample or determine the amount of components in a mixture. Infrared light refers to electromagnetic
radiation with wavenumber ranging from 13000 – 10 cm-1 (corresponding wavelength from 0.78 – 1000 μm). Infrared region
is further divided into three subregions: near-infrared (13000 – 4000 cm-1 or 0.78 – 2.5 μm), mid-infrared (4000 – 400 cm-1 or
2.5 – 25 μm) and far-infrared (400 – 10 cm-1 or 25 – 1000 μm). The most commonly used is the middle infrared region, since
molecules can absorb radiations in this region to induce the vibrational excitation of functional groups. Recently, applications
of near infrared spectroscopy have also been developed.
By passing infrared light through a sample and measuring the absorption or transmittance of light at each frequency, an
infrared spectrum is obtained, with peaks corresponding to the frequency of absorbed radiation. Since all groups have their
characteristic vibrational frequencies, information regarding molecular structure can be gained from the spectrum. Infrared
spectroscopy is capable of analyzing samples in almost any phase (liquid, solid, or gas), and can be used alone or in
combination with other instruments following different sampling procedures. Besides fundamental vibrational modes, other
factors such as overtone and combination bands, Fermi resonance, coupling and vibration-rotational bands also appear in the
spectrum. Due to the high information content of its spectrum, infrared spectroscopy has been a very common and useful tool
for structure elucidation and substance identification.

Instrumentation
Most commonly used instruments in infrared spectroscopy are dispersive infrared spectrometer and Fourier transform infrared
spectrometer.

Dispersive infrared spectrometer


Dispersive infrared spectrometer is mainly composed of radiation source, monochromator and detector. For mid-infrared
region, Globar (silicon carbide), Nernst glower (oxides of zirconium, yttrium and erbium) and metallic helices (chromium-
nickel alloy or tungsten) are frequently used as radiation sources. Tungsten-halogen lamps and metallic conductors coated with
ceramic are utilized as sources for near-infrared region. A mercury high-pressure lamp is suitable for far-infrared region.
Monochromator in conjunction with slits, mirrors and filters separates the wavelengths of light emitted. The dispersive
elements within monochromator are prisms or gratings. Gratings have gradually replaced prisms due to their comparatively
low cost and good quality. As shown in Figure 1, radiation passes through both a sample and a reference path. Then the beams
are directed to a diffraction grating (splitter), which disperses the light into component frequencies and directs each
wavelength through a slit to the detector. The detector produces an electrical signal and results in a recorder response.

Figure 1. Schematic illustration of dispersive infrared spectrometer. Figure from Wikipedia.


Two types of detectors are employed in dispersive infrared spectrometer, namely, thermal detectors and photon detectors.
Thermal detectors include thermocouples, thermistors, and pneumatic devices, which measure the heating effect generated by
infrared radiation. Photon detectors are semiconductor-based. Radiation is able to promote electrons in photon detectors from
valence band to conduction band, generating a small current. Photon detectors have faster response and higher sensitivity than
do thermal detectors but are more susceptible to thermal noise.

Fourier transform infrared spectrometer


Dispersive infrared spectrometer has many limitations because it examines component frequencies individually, resulting in
slow speed and low sensitivity. Fourier transform infrared (FTIR) spectrometer is preferred over dispersive spectrometer, since

6/27/2021 Infrared.1 https://chem.libretexts.org/@go/page/1845


it is capable of handling all frequencies simultaneously with high throughput, reducing the time required for analysis. The
radiation sources used in dispersive infrared spectrometer can also be used in FTIR spectrometer.
In contrast with the monochromator in dispersive spectrometer, FTIR spectrometer as shown in Figure 2 employs an
interferometer. The beamsplitter within the interferometer splits the incoming infrared beam into two beams, one of which is
reflected by a fixed mirror, while the other one reflected by a moving mirror perpendicular to the fixed one. The length of path
one beam travels is fixed and that of the other one is changing as the mirror moves, generating an optical path difference
between the two beams. After meeting back at the beamsplitter, the two beams recombine, interfere with each other, and yield
an interferogram. The interferogram produces inference signal as a function of optical path difference. It is converted to a
spectrum of absorbance or transmittance versus wavenumber or frequency by Fourier transform.

Figure 2. Schematic representation of Fourier transform infrared spectrometer. Figure from Wikipedia.
Detectors used in FTIR spectrometers are mainly pyroelectric and photoconductive detectors. The former are constructed of
crystalline materials (such as deuterated triglycine sulfate) whose electric polarization rely on temperature. The change in
temperature leads to change in charge distribution of the detector and electric signal is produced. The latter (such as mercury
cadmium telluride) provide better sensitivity and faster speed than do pyroelectric detectors over a broad spectral range.
However, liquid nitrogen is needed for cooling of photoconductive detectors.

Application
Since different molecules with different combination of atoms produce their unique spectra, infrared spectroscopy can be used
to qualitatively identify substances. In addition, the intensity of the peaks in the spectrum is proportional to the amount of
substance present, enabling its application for quantitative analysis.

Qualitative analysis
For qualitative identification purposes, the spectrum is commonly presented as transmittance versus wavenumber. Functional
groups have their characteristic fundamental vibrations which give rise to absorption at certain frequency range in the
spectrum (Figure 3).

Figure 3.Infrared spectrum of 1-hexanol.


Each band in a spectrum can be attributed to stretching or bending mode of a bond. Almost all the fundamental vibrations
appear in the mid-infrared region. For instance, 4000 – 2500 cm-1 region usually can be assigned to stretching modes of O-H,
N-H or C-H. Triple-bond stretching modes appear in the region of 2500 – 2000 cm-1. C=C and C=O stretching bands fall in
the 2000 – 1500 cm-1 region. Hence, characterization of functional groups in substances according to the frequencies and
intensities of absorption peaks is feasible, and also structures of molecules can be proposed. This method is applicable to
organic molecules, inorganic molecules, polymers, etc. A detailed frequency list of functional groups is shown in Figure 4.
Note that several functional groups may absorb at the same frequency range, and a functional group may have multiple-
characteristic absorption peaks, especially for 1500 – 650 cm-1, which is called the fingerprint region.

Figure 4. Characteristic infrared absorption frequencies.


Bands in the near-infrared region are overtones or combination bands. They are weak in intensity and overlapped, making
them not as useful for qualitative analysis as those in mid-infrared region. However, absorptions in this region are helpful in
exploiting information related to vibrations of molecules containing heavy atoms, molecular skeleton vibrations, molecular
torsions and crystal lattice vibrations.
Besides structural elucidation, another qualitative application of infrared spectroscopy is the identification of a compound with
a reference infrared spectrum. If all the peaks of the unknown match those of the reference, the compound can be identified.
Additional reference spectra are available online at databases such as NIST Chemistry WebBook.

Quantitative analysis

6/27/2021 Infrared.2 https://chem.libretexts.org/@go/page/1845


Absorbance is used for quantitative analysis due to its linear dependence on concentration. Given by Beer-Lambert law,
absorbance is directly proportional to the concentration and pathlength of sample:
A = ϵcl (Infrared.1)

where A is absorbance, ε the molar extinction coefficient or molar absorptivity which is characteristic for a specific substance,
c the concentration and l the pathlength (or the thickness) of sample. The conversion from transmittance to absorbance is given
by
A = − log T (Infrared.2)

where T is transmittance.
For quantitative analysis of liquid samples, usually an isolated peak with high molar absorptivity that appears in the spectrum
of the compound is chosen. A calibration curve of absorbance at the chosen frequency against concentration of the compound
is acquired by measuring the absorbance of a series of standard compound solution with known concentrations. These data are
then graphed to get a linear plot, from which the concentration of the unknown can be calculated after measuring its
absorbance at the same frequency. The number of functional groups can also be calculated in this way, since the molar
absorptivity of the band is proportional to the number of functional groups that are present in the compound.
For solid samples, an internal standard with a constant known amount is added to the unknown sample and the standards. Then
similar procedures as those with liquid samples are carried out except that the calibration curve is a graph of the ratio of
absorbance of analyte to that of the internal standard versus concentration of the analyte.
A multi-component analysis of the mixture is also feasible since different components have different values of molar
absorptivity at the same frequency.
However, infrared spectroscopy may be more susceptible to deviation from Beer's law than is UV-Vis spectroscopy because of
its narrow bands, complex spectra, weak incident beam, low transducer sensitivity and solvent absorption.

Reference
1. Settle, F. A. Handbook of instrumental techniques for analytical chemistry; Prentice Hall PTR: Upper Saddle River, NJ,
1997.
2. Heigl, J. J.; Bell, M.; White, J. U. Anal. Chem. 1947, 19, 293.
3. Baker, A. W. J. Phys. Chem. 1957, 61, 450.
4. Kamariotis, A.; Boyarkin, O. V.; Mercier, S. R.; Beck, R. D.; Bush, M. F.; Williams, E. R.; Rizzo, T. R. J. Am. Chem. Soc.
2006, 128, 905.
5. Stuart, B. Infrared spectroscopy fundamentals and applications; J. Wiley: Chichester, Eng.; Hoboken, N.J., 2004.
6. Günzler, H.; Heise, H. M. IR spectroscopy: an introduction; Wiley-VCH: Weinheim, 2002.
7. Wartewig, S. IR and Raman spectroscopy: fundamental processin; Wiley-VCH: Weinheim, 2003.

Outside Links
NIST Chemistry WebBook: http://webbook.nist.gov/
Wikipedia: http://en.Wikipedia.org/wiki/Infrared_spectroscopy

Contributor
Xixuan Li (UC Davis)

6/27/2021 Infrared.3 https://chem.libretexts.org/@go/page/1845


Infrared: Interpretation
Infrared spectroscopy is the study of the interaction of infrared light with matter. The fundamental measurement obtained in
infrared spectroscopy is an infrared spectrum, which is a plot of measured infrared intensity versus wavelength (or frequency)
of light.

Introduction
In infrared spectroscopy, units called wavenumbers are normally used to denote different types of light. The frequency,
wavelength, and wavenumber are related to each other via the following equation(1):

(1)
These equations show that light waves may be described by their frequency, wavelength or wavenumber. Here, we typically
refer to light waves by their wavenumber, however it will be more convenient to refer to a light wave's frequency or
wavelength. The wavenumber of several different types of light are shown in table 1.

Table 1. The Electromagnetic spectrum showing the wavenumber of several different types of light.
When a molecule absorbs infrared radiation, its chemical bonds vibrate. The bonds can stretch, contract, and bend. This is why
infrared spectroscopy is a type of vibrational spectroscopy. Fortunately, the complex vibrational motion of a molecule can be
broken down into a number of constituent vibrations called normal modes. For example, when a guitar string is plucked, the
string vibrates at its normal mode frequency. Molecules, like guitar strings, vibrate at specfic frequencies so different
molecules vibrate at different frequencies because their structures are different. This is why molecules can be distinguished
using infrared spectroscopy. The first necessary condition for a molecule to absorb infrared light is that the molecule must
have a vibration during which the change in dipole moment with respect to distance is non-zero. This condition can be
summarized in equation(2) form as follows:

(2)
Vibrations that satisfy this equation are said to be infrared active. The H-Cl stretch of hydrogen chloride and the asymmetric
stretch of CO2 are examples of infrared active vibrations. Infrared active vibrations cause the bands seen in an infrared
spectrum.
The second necessary condition for infrared absorbance is that the energy of the light impinging on a molecule must equal a
vibrational energy level difference within the molecule. This condition can be summarized in equation(3) form as follows:

(3)

7/2/2021 Infrared.1 https://chem.libretexts.org/@go/page/1846


If the energy of a photon does not meet the criterion in this equation, it will be transmitted by the sample and if the photon
energy satisfies this equation, that photon will be absorbed by the molecule.(See Infrared: Theory for more detail)
As any other analytical techniques, infrared spectroscopy works well on some samples, and poorly on others. It is important to
know the strengths and weaknesses of infrared spectroscopy so it can be used in the proper way. Some advantages and
disadvantages of infrared spectroscopy are listed in table 2.

Advantages Disadvantages

Solids, Liquids, gases, semi-solids, powders and polymers are all


analyzed
Atoms or monatomic ions do not have infrared spectra
The peak positions, intensities, widths, and shapes all provide useful
Homonuclear diatomic molecules do not posses infrared spectra
information
Complex mixture and aqueous solutions are difficult to analyze using
Fast and easy technique
infrared spectroscopy
Sensitive technique (Micrograms of materials can be detected routinely)
Inexpensive

Table 2. The Advantage and Disadvantage of Infrared Spectroscopy

Origin of Peak Positions, Intensities, and Widths


Peak Positions
The equation(4) gives the frequency of light that a molecule will absorb, and gives the frequency of vibration of the normal
mode excited by that light.

(4)
Only two variables in equation(4) are a chemical bond's force constant and reduced mass. Here, the reduced mass refers to
(M1M2)/(M1+M2) where M1 and M2 are the masses of the two atoms, respectively. These two molecular properties determine
the wavenumber at which a molecule will absorb infrared light. No two chemical substances in the universe have the same
force constants and atomic masses, which is why the infrared spectrum of each chemical substance is unique. To understand
the effect of atomic masses and force constant on the positions of infrared bands, table 3 and 4 are shown as an example,
respectively.
Table 3. An Example of an Mass Effect
Bond C-H Stretch in cm-1

C-1H ~3000
C-2D ~2120

The reduced masses of C-1H and C-2D are different, but their force constants are the same. By simply doubling the mass of the
hydrogen atom, the carbon-hydrogen stretching vibration is reduced by over 800cm-1.
Table 4. An Example of an electronic Effect

Bond C-H Stretch in cm-1

C-H ~3000
H-C=O ~2750

When a hydrogen is attached to a carbon with a C=O bond, the C-H stretch band position decrease to ~2750cm-1. These two
C-H bonds have the same reduced mass but different force constants. The oxygen in the second molecule pulls electron
density away from the C-H bond so it makes weaken and reduce the C-H force constant. This cause the C-H stretching
vibration to be reduced by ~250cm-1.

The Origin of Peak Intensities

7/2/2021 Infrared.2 https://chem.libretexts.org/@go/page/1846


The different vibrations of the different functional groups in the molecule give rise to bands of differing intensity. This is
∂μ
because ∂x
is different for each of these vibrations. For example, the most intense band in the spectrum of octane shown in
Figure 3 is at 2971, 2863 cm-1 and is due to stretching of the C-H bond. One of the weaker bands in the spectrum of octane is
at 726cm-1, and it is due to long-chain methyl rock of the carbon-carbon bonds in octane. The change in dipole moment with
respect to distance for the C-H stretching is greater than that for the C-C rock vibration, which is why the C-H stretching band
is the more intense than C-C rock vibration.
Another factor that determines the peak intensity in infrared spectra is the concentration of molecules in the sample. The
equation(5) that relates concentration to absorbance is Beer's law,

(5)

The absorptivity is the proportionality constant between concentration and absorbance, and is dependent on (¶µ/¶x)2. The
absorptivity is an absolute measure of infrared absorbance intensity for a specific molecule at a specific wavenumber. For pure
sample, concentration is at its maximum, and the peak intensities are true representations of the values of ¶µ/¶x for different
vibrations. However, in a mixture, two peaks may have different intensities because there are molecules present in different
concentration.

The Orgins of Peak Widths


In general, the width of infrared bands for solid and liquid samples is determined by the number of chemical environments
which is related to the strength of intermolecular interactions such as hydrogen bonding. Figure 1. shows hydrogen bond in
water molecules and these water molecules are in different chemical environments. Because the number and strength of
hydrogen bonds differs with chemical environment, the force constant varies and the wavenumber differs at which these
molecules absorb infrared light.

Figure 1. Hydrogen Bonding in water molecules


In any sample where hydrogen bonding occurs, the number and strength of intermolecular interactions varies greatly within
the sample, causing the bands in these samples to be particularly broad. This is illustrated in the spectra of ethanol(Fig7) and
hexanoic acid(Fig11). When intermolecular interactions are weak, the number of chemical environments is small, and narrow
infrared bands are observed.

The Origin of Group Frequencies


An important observation made by early researchers is that many functional group absorb infrared radiation at about the same
wavenumber, regardless of the structure of the rest of the molecule. For example, C-H stretching vibrations usually appear
between 3200 and 2800cm-1 and carbonyl(C=O) stretching vibrations usually appear between 1800 and 1600cm-1. This makes
these bands diagnostic markers for the presence of a functional group in a sample. These types of infrared bands are called
group frequencies because they tell us about the presence or absence of specific functional groups in a sample.

7/2/2021 Infrared.3 https://chem.libretexts.org/@go/page/1846


Figure 2. Group frequency and fingerprint regions of the mid-infrared spectrum
The region of the infrared spectrum from 1200 to 700 cm-1 is called the fingerprint region. This region is notable for the large
number of infrared bands that are found there. Many different vibrations, including C-O, C-C and C-N single bond stretches,
C-H bending vibrations, and some bands due to benzene rings are found in this region. The fingerprint region is often the most
complex and confusing region to interpret, and is usually the last section of a spectrum to be interpreted. However, the utility
of the fingerprint region is that the many bands there provide a fingerprint for a molecule.

Spectral Interpretation by Application of Group Frequencies


Organic Compounds
One of the most common application of infrared spectroscopy is to the identification of organic compounds. The major classes
of organic molecules are shown in this category and also linked on the bottom page for the number of collections of spectral
information regarding organic molecules.
Hydrocarbons
Hydrocarbons compounds contain only C-H and C-C bonds, but there is plenty of information to be obtained from the infrared
spectra arising from C-H stretching and C-H bending.
In alkanes, which have very few bands, each band in the spectrum can be assigned:
C–H stretch from 3000–2850 cm-1
C–H bend or scissoring from 1470-1450 cm-1
C–H rock, methyl from 1370-1350 cm-1
C–H rock, methyl, seen only in long chain alkanes, from 725-720 cm-1
Figure 3. shows the IR spectrum of octane. Since most organic compounds have these features, these C-H vibrations are
usually not noted when interpreting a routine IR spectrum. Note that the change in dipole moment with respect to distance for
the C-H stretching is greater than that for others shown, which is why the C-H stretch band is the more intense.

Figure 3. Infrared Spectrum of Octane


In alkenes compounds, each band in the spectrum can be assigned:

7/2/2021 Infrared.4 https://chem.libretexts.org/@go/page/1846


C=C stretch from 1680-1640 cm-1
=C–H stretch from 3100-3000 cm-1
=C–H bend from 1000-650 cm-1
Figure 4. shows the IR spectrum of 1-octene. As alkanes compounds, these bands are not specific and are generally not noted
because they are present in almost all organic molecules.

Figure 4. Infrared Spectrum of 1-Octene


In alkynes, each band in the spectrum can be assigned:
–C?C– stretch from 2260-2100 cm-1
–C?C–H: C–H stretch from 3330-3270 cm-1
–C?C–H: C–H bend from 700-610 cm-1
The spectrum of 1-hexyne, a terminal alkyne, is shown below.

Figure 5. Infrared Spectrum of 1-Hexyne


In aromatic compounds, each band in the spectrum can be assigned:
C–H stretch from 3100-3000 cm-1
overtones, weak, from 2000-1665 cm-1
C–C stretch (in-ring) from 1600-1585 cm-1
C–C stretch (in-ring) from 1500-1400 cm-1
C–H "oop" from 900-675 cm-1
Note that this is at slightly higher frequency than is the –C–H stretch in alkanes. This is a very useful tool for interpreting IR
spectra. Only alkenes and aromatics show a C–H stretch slightly higher than 3000 cm-1.
Figure 6. shows the spectrum of toluene.

7/2/2021 Infrared.5 https://chem.libretexts.org/@go/page/1846


Figure 6. Infrared Spectrum of Toluene
Functional Groups Containing the C-O Bond
Alcohols have IR absorptions associated with both the O-H and the C-O stretching vibrations.
O–H stretch, hydrogen bonded 3500-3200 cm-1
C–O stretch 1260-1050 cm-1 (s)
Figure 7. shows the spectrum of ethanol. Note the very broad, strong band of the O–H stretch.

Figure 7. Infrared Spectrum of Ethanol


The carbonyl stretching vibration band C=O of saturated aliphatic ketones appears:
C=O stretch - aliphatic ketones 1715 cm-1
- ?, ?-unsaturated ketones 1685-1666 cm-1
Figure 8. shows the spectrum of 2-butanone. This is a saturated ketone, and the C=O band appears at 1715.

Figure 8. Infrared Spectrum of 2-Butanone


If a compound is suspected to be an aldehyde, a peak always appears around 2720 cm-1 which often appears as a shoulder-type
peak just to the right of the alkyl C–H stretches.
H–C=O stretch 2830-2695 cm-1
C=O stretch:

7/2/2021 Infrared.6 https://chem.libretexts.org/@go/page/1846


aliphatic aldehydes 1740-1720 cm-1
alpha, beta-unsaturated aldehydes 1710-1685 cm-1
Figure 9. shows the spectrum of butyraldehyde.

Figure 9. Infrared Spectrum of Butyraldehyde


The carbonyl stretch C=O of esters appears:
C=O stretch
aliphatic from 1750-1735 cm-1
?, ?-unsaturated from 1730-1715 cm-1
C–O stretch from 1300-1000 cm-1
Figure 10. shows the spectrum of ethyl benzoate.

Figure 10. Infrared Spectrum of Ethyl benzoate


The carbonyl stretch C=O of a carboxylic acid appears as an intense band from 1760-1690 cm-1. The exact position of this
broad band depends on whether the carboxylic acid is saturated or unsaturated, dimerized, or has internal hydrogen bonding.
O–H stretch from 3300-2500 cm-1
C=O stretch from 1760-1690 cm-1
C–O stretch from 1320-1210 cm-1
O–H bend from 1440-1395 and 950-910 cm-1
Figure 11. shows the spectrum of hexanoic acid.

7/2/2021 Infrared.7 https://chem.libretexts.org/@go/page/1846


Figure 11. Infrared Spectrum of Hexanoic acid
Organic Nitrogen Compounds
N–O asymmetric stretch from 1550-1475 cm-1
N–O symmetric stretch from 1360-1290 cm-1

Figure 12. Infrared Spectrum of Nitomethane


Organic Compounds Containing Halogens
Alkyl halides are compounds that have a C–X bond, where X is a halogen: bromine, chlorine, fluorene, or iodine.
C–H wag (-CH2X) from 1300-1150 cm-1
C–X stretches (general) from 850-515 cm-1
C–Cl stretch 850-550 cm-1
C–Br stretch 690-515 cm-1
The spectrum of 1-chloro-2-methylpropane are shown below.

Figure 13. Infrared Spectrum of 1-chloro-2-methylpropane


For more Infrared spectra Spectral database of organic molecules is introduced to use free database. Also, the infrared
spectroscopy correlation tableis linked on bottom of page to find other assigned IR peaks.

7/2/2021 Infrared.8 https://chem.libretexts.org/@go/page/1846


Inorganic Compounds
Generally, the infrared bands for inorganic materials are broader, fewer in number and appear at lower wavenumbers than
those observed for organic materials. If an inorganic compound forms covalent bonds within an ion, it can produce a
characteristic infrared spectrum.
Main infrared bands of some common inorganic ions:
CO32- 1450-1410, 880-800cm-1
SO42- 1130-1080, 680-610cm-1
NO3- 1410-1340, 860-800cm-1
PO43- 1100-950cm-1
SiO42- 1100-900cm-1
NH4+ 3335-3030, 1485-1390cm-1
MnO4- 920-890, 850-840cm-1
Diatomic molecules produce one vibration along the chemical bond. Monatomic ligand, where metal s coordinate with atoms
such as halogens, H, N or O, produce characteristic bands. These bands are summarized in below.
Chracteristic infrared bands of diatomic inorganic molecules: M(metal), X(halogen)
M-H stretching 2250-1700cm-1
M-H bending 800-600cm-1
M-X stretching 750-100cm-1
M=O stretching 1010-850cm-1
M=N stretching 1020-875cm-1
The normal modes of vibration of linear and bent triatomic molecules are illustrated and some common linear and bent
triatomic molecules are shown below. Note that some molecules show two bands for ?1because of Fermi resonance.

Characteristic infrared bands(cm-1) of triatomic inorganic molecules:


Linear Molecules OCO HCN NCS- ClCN MgCl2
1388, 1286 3311 2053 714, 784 327
667 712 486, 471 380 249
2349 2049 748 2219 842
Bent Molecules H2O O3 SnCl 2
3675 1135 354
1595 716 120
3756 1089 334

Identification
There are a few general rules that can be used when using a mid-infrared
spectrum for the determination of a molecular structure. The following is a
suggested strategy for spectrum interpretation:2
1. Look first at the high-wavenumber end of the spectrum(>1500cm-1) and
concentrate initially on the major bands
2. For each band, 'short-list' the possibilities by using a correlation table
3. Use the lower-wavenumber end of the spectrum for the confirmation or
elaboration of possible structural elements
4. Do not expect to be able to assign every band in the spectrum
5. Keep 'cross-checking' wherever possible.
6. Exploit negative evidence as well as positive evidence

7/2/2021 Infrared.9 https://chem.libretexts.org/@go/page/1846


7. Band intensities should be treated with some caution. Under certain circumstances, they may vary considerably for the
same group
8. Take care when using small wavenumber changes. If in solution, some bands are very 'solvent-sensitive'
9. Do not forget to subtract slovent bands if possible
Infrared spectroscopy is used to analyze a wide variety of samples, but it cannot solve every chemical analysis problem. When
used in conjunction with other methods such as mass spectroscopy, nuclear magnetic resonance, and elemental analysis,
infrared spectroscopy usually makes possible the positive identification of a sample.

References
1. Infrared Spectral Interpretation by Brian Smith, CRC Press, 1999
2. Infrared Spectroscopy: Fundamentals and Applications by Barbara Atuart, John Wiley&Sons, Ltd., 2004
3. Interpretation of Infrared Spectra, A Practical Approach by John Coates in Encyclopedia of Analytical Chemistry pp.
10815-10837, John Wiley&Sons Ltd, Chichester, 2000

Outside Links
Spectral Database for Organic Compounds SDBS: http://riodb01.ibase.aist.go.jp/sdbs/ (National Institute of Advanced
Industrial Science and Technology, date of access)
Infrared Spectroscopy Correlation Table: en.Wikipedia.org/wiki/Infrared_spectroscopy_correlation_table
FDM Reference Spectra Databases: http://www.fdmspectra.com/index.html
Other Usuful Web Pages:
www.cem.msu.edu/~reusch/Virtu...d/infrared.htm
Fermi resonance : en.Wikipedia.org/wiki/Fermi_resonance

7/2/2021 Infrared.10 https://chem.libretexts.org/@go/page/1846


Infrared Spectroscopy
Infrared (IR) spectroscopy is one of the most common and widely used spectroscopic techniques employed mainly by
inorganic and organic chemists due to its usefulness in determining structures of compounds and identifying them. Chemical
compounds have different chemical properties due to the presence of different functional groups.

Introduction
Infrared (IR) spectroscopy is one of the most common and widely used spectroscopic techniques. Absorbing groups in the
infrared region absorb within a certain wavelength region. The absorption peaks within this region are usually sharper when
compared with absorption peaks from the ultraviolet and visible regions. In this way, IR spectroscopy can be very sensitive to
determination of functional groups within a sample since different functional group absorbs different particular frequency of
IR radiation. Also, each molecule has a characteristic spectrum often referred to as the fingerprint. A molecule can be
identified by comparing its absorption peak to a data bank of spectra. IR spectroscopy is very useful in the identification and
structure analysis of a variety of substances, including both organic and inorganic compounds. It can also be used for both
qualitative and quantitative analysis of complex mixtures of similar compounds.
The use of infrared spectroscopy began in the 1950's by Wilbur Kaye. He had designed a machine that tested the near-infrared
spectrum and provided the theory to describe the results. Karl Norris started using IR Spectroscopy in the analytical world in
the 1960's and as a result IR Spectroscopy became an accepted technique. There have been many advances in the field of IR
Spec, the most notable was the application of Fourier Transformations to this technique thus creating an IR method that had
higher resolution and a decrease in noise. The year this method became accepted in the field was in the late 1960's.4

Absorption Spectroscopy
There are three main processes by which a molecule can absorb radiation. and each of these routes involves an increase of
energy that is proportional to the light absorbed. The first route occurs when absorption of radiation leads to a higher rotational
energy level in a rotational transition. The second route is a vibrational transition which occurs on absorption of quantized
energy. This leads to an increased vibrational energy level. The third route involves electrons of molecules being raised to a
higher electron energy, which is the electronic transition. It’s important to state that the energy is quantized and absorption of
radiation causes a molecule to move to a higher internal energy level. This is achieved by the alternating electric field of the
radiation interacting with the molecule and causing a change in the movement of the molecule. There are multiple possibilities
for the different possible energy levels for the various types of transitions.
The energy levels can be rated in the following order: electronic > vibrational > rotational. Each of these transitions differs by
an order of magnitude. Rotational transitions occur at lower energies (longer wavelengths) and this energy is insufficient and
cannot cause vibrational and electronic transitions but vibrational (near infra-red) and electronic transitions (ultraviolet region
of the electromagnetic spectrum) require higher energies.

Figure 1: Energy levels for a molecule. Possible transitions that occur: (A): Pure rotational Transitions, (B) rotational-
Vibrational Transitions, (C) Rotational-Vibrational-Electronic Transitions

7/5/2021 1 https://chem.libretexts.org/@go/page/1847
The energy of IR radiation is weaker than that of visible and ultraviolet radiation, and so the type of radiation produced is
different. Absorption of IR radiation is typical of molecular species that have a small energy difference between the rotational
and vibrational states. A criterion for IR absorption is a net change in dipole moment in a molecule as it vibrates or rotates.
Using the molecule HBr as an example, the charge distribution between hydrogen and bromine is not evenly distributed since
bromine is more electronegative than hydrogen and has a higher electron density. H Br thus has a large dipole moment and is
thus polar. The dipole moment is determined by the magnitude of the charge difference and the distance between the two
centers of charge. As the molecule vibrates, there is a fluctuation in its dipole moment; this causes a field that interacts with
the electric field associated with radiation. If there is a match in frequency of the radiation and the natural vibration of the
molecule, absorption occurs and this alters the amplitude of the molecular vibration. This also occurs when the rotation of
asymmetric molecules around their centers results in a dipole moment change, which permits interaction with the radiation
field.

Molecules such as O2, N2, Br2, do not have a changing dipole moment (amplitude nor orientation) when they undergo
rotational and vibrational motions, as a result, they cannot cannot absorb IR radiation.

Diatomic Molecular Vibration


The absorption of IR radiation by a molecule can be likened to two atoms attached to each other by a massless spring.
Considering simple diatomic molecules, only one vibration is possible. The Hook's law potential on the other hand is based on
an ideal spring
F = −kx (1)

dV (x)
=− (2)
dx

this results in one dimensional space


1 2
V (r) = k(r − req ) (3)
2

One thing that the Morse and Harmonic oscillator have in common is the small displacements (x = r − r ) from the eq

equilibrium. Solving the Schrödinger equation for the harmonic oscillator potential results in the energy levels results in
1
Ev = (v + ) h ve (4)
2

with v = 0, 1, 2, 3, . . . , inf inity



1 k
ve = √ (5)
2π μ

When calculating the energy of a diatomic molecule, factors such as anharmonicity (has a similar curve with the harmonic
oscillator at low potential energies but deviates at higher energies) are considered. The energy spacing in the harmonic
oscillator is equal but not so with the anharmonic oscillator. The anharmonic oscillator is a deviation from the harmonic
oscillator. Other considered terms include; centrifugal stretching, vibrational and rotational interactions have to be taken into
account. The energy can be expressed mathematically as
2
1 1 2 2
1
Ev = (v + ) h ve − (v + ) Xe h ve + Be J(J + 1) − De J (J + 1 ) − αe (v + ) J(J + 1) (6)
2 2 Rigid Rotor centrifugal stretching
2
Harmonic Oscillator anharmonicity rovibrational coupling

The first and third terms represent the harmonicity and rigid rotor behavior of a diatomic molecule such as HCl. The second
term represents anharmonicity and the fourth term represents centrifugal stretching. The fifth term represents the interaction
between the vibration and rotational interaction of the molecule.

Polyatomic Molecular Vibration

7/5/2021 2 https://chem.libretexts.org/@go/page/1847
The bond of a molecule experiences various types of vibrations and rotations. This causes the atom not to be stationary and to
fluctuate continuously. Vibrational motions are defined by stretching and bending modes. These movements are easily defined
for diatomic or triatomic molecules. This is not the case for large molecules due to several vibrational motions and interactions
that will be experienced. When there is a continuous change in the interatomic distance along the axis of the bond between two
atoms, this process is known as a stretching vibration. A change in the angle occurring between two bonds is known as a
bending vibration. Four bending vibrations exist namely, wagging, twisting, rocking and scissoring. A CH2 group is used as an
example to illustrate stretching and bending vibrations below.

Symmetric Stretch Asymmetric Stretch Twisting

Wagging Scissoring Rocking


Figure 3: Types of Vibrational Modes. To ensure that no center of mass motion occurs, the center atom (yellow ball) will also
move. Figure from Wikipedia
As stated earlier, molecular vibrations consist of stretching and bending modes. A molecule consisting of (N) number of atoms
has a total of 3N degrees of freedom, corresponding to the Cartesian coordinates of each atom in the molecule. In a non-linear
molecule, 3 of these degrees of freedom are rotational, 3 are translational and the remainder is fundamental vibrations. In a
linear molecule, there are 3 translational degrees of freedom and 2 are rotational. This is because in a linear molecule, all of
the atoms lie on a single straight line and hence rotation about the bond axis is not possible. Mathematically the normal modes
for a linear and non linear can be expressed as
Linear Molecules: (3N - 5) degrees of freedom
Non-Linear molecules: (3N - 6) degrees of freedom

Example 1: Vibrations of Water


Diagram of Stretching and Bending Modes for H2O.
Solution
H2O molecule is a non-linear molecule due to the uneven distribution of the electron density. O2 is more electronegative
than H2 and carries a negative charge, while H has a partial positive charge. The total degrees of freedom for H2O will be
3(3)-6 = 9-6 = 3 degrees of freedom which correspond to the following stretching and bending vibrations. The vibrational
modes are illustrated below:

Figure 4 The vibrational modes of H2O

7/5/2021 3 https://chem.libretexts.org/@go/page/1847
Example Vibrations of CO 2

Diagram of Stretching and Bending Modes for CO2.


Solution
CO2 is a linear molecule and thus has the formula (3N-5). It has 4 modes of vibration (3(3)-5). CO2 has 2 stretching
modes, symmetric and asymmetric. The CO2 symmetric stretch is not IR active because there is no change in dipole
moment because the net dipole moments are in opposite directions and as a result, they cancel each other. In the
asymmetric stretch, O atom moves away from the C atom and generates a net change in dipole moments and hence
absorbs IR radiation at 2350 cm-1. The other IR absorption occurs at 666 cm-1. CO2 symmetry with D CO2 has a total ∞h

of four of stretching and bending modes but only two are seen. Two of its bands are degenerate and one of the vibration
modes is symmetric hence it does not cause a dipole moment change because the polar directions cancel each other. The
vibrational modes are illustrated below:

Figure 5 The vibrational modes of CO2

The Deduction of Frequency


The second law of Newton states that
F = ma (7)

where m is the mass and a is the acceleration, acceleration is a 2nd order differential equation of distance with respect to time.
Thus "a" can be written as
2
d y
a = (8)
dt

Substituting this into Equation 1 gives


2
md y
= −ky (9)
2
dt

−k
the 2nd order differential equation of this equation is equal to displacement of mass and time can be stated as
m

y = A cos 2π νm t (10)

where vm is the natural vibrational frequency and A is the maximum amplitude of the motion. On differentiating a second time
the equation becomes
2
d y
2 2
= −4 π νm A cos 2π νm t
2
dt

substituting the two equations above into Newton's second law for a harmonic oscillator,
2 2
m ∗ (−4 π νm Acos 2π νm t) = −k ∗ (Acos 2π νm t) (11)

If we cancel out the two functions y ,


2 2
4m π νm = k (12)

from above, we obtain the natural frequency of the oscillation.

7/5/2021 4 https://chem.libretexts.org/@go/page/1847

−−
1 k
νm = √ (13)
2π m

νm which is the natural frequency of the mechanical oscillator which depends on the force constant of the spring and the mass
of the attached body and independent of energy imparted on the system. when there are two masses involved in the system
then the mass used in the above equation becomes
m1 m2
μ = (14)
m1 + m2

The vibrational frequency can be rewritten as




1 k
νm = √ (15)
2π μ

The Deduction of Wave Number


Using the harmonic oscillator and wave equations of quantum mechanics, the energy can be written as


1 h k
E = (v + ) √ (16)
2 2π μ

where h is Planck's constant and v is the vibrational quantum number and ranges from 0,1,2,3.... infinity.
1
E = (v + ) h vm (17)
2

where ν is the vibrational frequency. Transitions in vibrational energy levels can be brought about by absorption of radiation,
m

provided the energy of the radiation exactly matches the difference in energy levels between the vibrational quantum states and
provided the vibration causes a change in dipole moment. This can be expressed as


h k
△E = h vm = √ (18)
2π μ

At room temperature, the majority of molecules are in the ground state v = 0, from the equation above
1
Eo = h vm (19)
2

following the selection rule, when a molecule absorbs energy, there is a promotion to the first excited state
3
E1 = h vm (20)
2

3 1
( h vm − h vm ) = h vm (21)
2 2

The frequency of radiation v that will bring about this change is identical to the classical vibrational frequency of the bond vm
and it can be expressed as


h k
Eradiation = hv = △E = h vm = √ (22)
2π μ

The above equation can be modified so that the radiation can be expressed in wave numbers


h k
ν̃ = √ (23)
2πc μ

where
c is the velocity of light (cm s-1) and

7/5/2021 5 https://chem.libretexts.org/@go/page/1847
ν̃ is the wave number of an absorption maximum (cm-1)

Theory of IR
Molecular vibrational frequencies lie in the IR region of the electromagnetic spectrum, and they can be measured using the IR
technique. In IR, polychromatic light (light having different frequencies) is passed through a sample and the intensity of the
transmitted light is measured at each frequency. When molecules absorb IR radiation, transitions occur from a ground
vibrational state to an excited vibrational state (Figure 1).
For a molecule to be IR active there must be a change in dipole moment as a result of the vibration that occurs when IR
radiation is absorbed. Dipole moment is a vector quantity and depends on the orientation of the molecule and the photon
electric vector. The dipole moment changes as the bond expands and contracts. When all molecules are aligned as in a crystal
and the photon vector points along a molecular axis such as z. Absorption occurs for the vibrations that displace the dipole
along z. Vibrations that are totally x or y polarized would be absent. Dipole moment in a heteronuclear diatomic molecule can
be described as uneven distribution of electron density between the atoms. One atom is more electronegative than the other
and has a net negative charge.
The dipole moment can be expressed mathematically as
μ = er

The relationship between IR intensity and dipole moment is given as


2

IIR ∝ ( )
dQ

relating this to intensity of the IR radiation, we have have the following equation below.
where μ is the dipole moment and Q is the vibrational coordinate. The transition moment integral, that gives information
about the probability of a transition occurring, for IR can also be written as
^ |ψ ⟩
⟨ψ| M f

i and f represent are initial and final states. ψ is the wave function. Relating this to IR intensity we have
i

^ |ψ ⟩
IIR ∝ ⟨ψ| M f

where M ^
is the dipole moment and has the Cartesian coordinates, ^ ^
Mx My , , ^
Mz . In order for a transition to occur by dipole
selection rules , at least one of the integrals must be non zero.

Region of IR
The IR region of the electromagnetic spectrum ranges in wavelength from 2 -15 µm. Conventionally the IR region is
subdivided into three regions, near IR, mid IR and far IR. Most of the IR used originates from the mid IR region. The table
below indicates the IR spectral regions
Region Wavelength Wavenumbers (V), cm-1 Frequencies (v), HZ

Near 0.78 -2.5 12800 - 4000 3.8 x 1014 - 1.2 x 1014

Middle 2.5 - 50 4000 - 200 3.8 x 1014 - 1.2 x 1014


Far 50 -100 200 -10 3.8 x 1014 - 1.2 x 1014
Most Used 2.5 -15 4000 -670 3.8 x 1014 - 1.2 x 1014

IR deals with the interaction between a molecule and radiation from the electromagnetic region ranging (4000- 40 cm-1). The
cm-1 is the wave number scale and it can also be defined as 1/wavelength in cm. A linear wavenumber is often used due to its
direct relationship with both frequency and energy. The frequency of the absorbed radiation causes the molecular vibrational
frequency for the absorption process. The relationship is given below
1 μm v(H z)
−1 4
v̄ (c m ) = × 10 ( ) = (24)
λ(μm) cm c(cm/s)

7/5/2021 6 https://chem.libretexts.org/@go/page/1847
Near InfraRed Spectroscopy: Absorption bands in the near infrared (NIR) region (750 - 2500 nm) are weak because they
arise from vibrational overtones and combination bands. Combination bands occur when two molecular vibrations are
excited simultaneously. The intensity of overtone bands reduces by one order of overtone for each successive overtone.
When a molecule is excited from the ground vibrational state to a higher vibrational state and the vibrational quantum
number v is greater than or equal to 2 then an overtone absorption results. The first overtone results from v = 0 to v = 2.
The second overtone occurs when v =0 transitions to v = 3. Transitions arising from the near ir absorption are weak, hence
they are referred to as forbidden transitions but these transitions are relevant when non-destructive measurements are
required such as a solid sample. Near IR spectra though have low absorption they have a high signal to noise ratio owing to
intense radiation sources and NIR is able to penetrate undiluted samples and use longer path lengths; it becomes very
useful for rapid measurement of more representative samples.
Far InfraRed Spectroscopy: The far IR region is particularly useful for inorganic studies due to stretching and bending
vibrations of bonds between the metal atoms and ligands. The frequencies, which these vibrations are observed, are usually
lower than 650 cm-1. Pure rotational absorption of gases is observed in the far IR region when there is a permanent dipole
moment present. Examples include H2O, O3, HCl.

IR Analysis
Qualitative Analysis
IR spectroscopy is a great method for identification of compounds, especially for identification of functional groups.
Therefore, we can use group frequencies for structural analysis. Group frequencies are vibrations that are associated with
certain functional groups. It is possible to identify a functional group of a molecule by comparing its vibrational frequency on
an IR spectrum to an IR stored data bank.
Here, we take the IR spectrum of Formaldehyde for an example. Formaldehyde has a C=O functional group and C-H bond.
The value obtained from the following graph can be compared to those in reference data banks stored for Formaldehyde. A
molecule with a C=O stretch has an IR band which is usually found near 1700 cm-1 and around 1400 cm-1 for CH2 bend. It's
important to note that this value is dependent on other functional groups present on the molecule. The higher 1700 cm-1
indicates a large dipole moment change. It is easier to bend a molecule than stretch it, hence stretching vibrations have higher
frequencies and require higher energies than bending modes. The finger print region is a region from 1400-650 cm-1. Each
molecule has it's own characteristic print and is often cumbersome to attach any values to this region.

Figure 6 IR Spectrum of Formaldehyde

Quantitative Analysis
Infrared spectroscopy can also be applied in the field of quantitative analysis, although sometimes it's not as accurate as other
analytical methods, like gas chromatography and liquid chromatography. The main theory of IR quantification is Beer's law or
Beer-Lambert law, which is written as
I0
A = log( ) = ϵlc (25)
I

Where A is the absorbance of the sample, I is the intensity of transmitted light, I0 is the intensity of incident light, l is the path
length, a is the molar absorptivity of the substance, and c is the concentration of the substance.

7/5/2021 7 https://chem.libretexts.org/@go/page/1847
From the Beer's Law, we could figure out the relation between the absorbance and the concentration of the sample since the
analytes have a particular molar absorptivity at a particular wavelength. Therefore, we could use IR spectroscopy and Beer's
Law to find the concentration of substance or the components of mixture. This is how the IR quantification operated.

Selection Rules of IR
In order for vibrational transitions to occur, they are normally governed by some rules referred to as selection rules.
1. An interaction must occur between the oscillating field of the electromagnetic radiation and the vibrational molecule for a
transition to occur. This can be expressed mathematically as

( ) ≠0
dr
req

△v = +1 and △J = +1
2. This holds for a harmonic oscillator because the vibrational levels are equally spaced and that accounts for the single peak
observed in any given molecular vibration. For gases J changes +1 for the R branch and -1 for the P branch.△J = 0 is a
forbidden transition and hence a q branch for a diatomic will not be present. For any anharmonic oscillator, the selection
rule is not followed and it follows that the change in energy becomes smaller. This results in weaker transitions called
overtones, then △v = +2 (first overtone) can occur, as well as the 2nd overtone △v = +3 . The frequencies of the 1st and
2nd overtones provides information about the potential surface and about two to three times that of the fundamental
frequency.
3. For a diatomic, since μ is known, measurement of ue provides a value for k, the force constant.
2
d V (r)
k =( )
2
dr
req

where k is the force constant and indicates the strength of a bond.

Influence Factors of IR
Isotope Effects: It's been observed that the effect on k when an atom is replaced by an isotope is negligible but it does
have an effect on ν due to changes in the new mass. This is because the reduced mass has an effect on the rotational and
vibrational behavior.
Solvent Effects: The polarity of solvent will have an influence on the IR spectra of organic compounds due to the
interactions between solvent and compounds, which is called solvent effects. If we place a compound, which contains n, pi
and pi* orbitals, into a polar solvent, the solvent will stabilizes these three orbitals in different extent. The stabilization
effects of polar solvent on n orbital is the largest one, the next larger one is pi* orbital, and the effects on pi orbital is the
smallest one. The spectra of n→pi* transition will shift to blue side, which means it will move to shorter wavelengths and
higher energies since the polar solvent causes the energy difference between n orbital and pi* orbital to become bigger. The
spectra of pi→pi* transition will shift to red side, which means it will move to longer wavelengths and lower energies
since the polar solvent causes the energy difference between n orbital and pi* orbital to become smaller.

Advantages of IR
High Scan Speed: Infrared spectroscopy can get information for the whole range of frequency simultaneously, within one
second. Therefore, IR can be used to analyze a substance that is not very stable and finish the scan before it start to
decompose.
High Resolution: The resolution of general prism spectrometer is only about 3 cm-1, but the resolution of infrared
spectrometer is much higher. For example, the resolution of Grating infrared spectrometer could be 0.2 cm-1, the resolution
of FT infrared spectrometer could be 0.1-0.005 cm-1.
High Sensitivity: With Fourier Transform, the infrared spectrometer doesn't need to use the slit and monochromator. In
this way, the reflection specularity will be increased and the loss of energy in the analysis process will be decreased.
Therefore the energy that reaches the detector is large enough and even very small amount of analytes could be detected.
Nowadays, the infrared spectroscopy could detect the sample as small as 1-10 grams.
Wide Range of Application: Infrared spectroscopy could be used to analyze almost all organic compounds and some
inorganic compounds. It has a wide range of application in both qualitative analysis and quantitative analysis. Also, the

7/5/2021 8 https://chem.libretexts.org/@go/page/1847
sample of Infrared spectroscopy doesn't have phase constraints. It could be gas, liquid or solid, which has enlarged the
range of analytes a lot.
Large Amount of Information: Infrared Spectra could give us lots of structural information of the analytes, such as the
type of compound, the functional group of compound, the stereoscopic structure of compound, the number and position of
substituent group and so on. Depending on the available information form the functional part and the fingerprint part,
infrared spectroscopy has become a great method to identify different kinds of compounds.
Non-Destructive: Infrared Spectroscopy is non-destructive to the sample.

Disadvantages of IR
Sample Constraint: Infrared spectroscopy is not applicable to the sample that contains water since this solvent strongly
absorb IR light.
Spectrum Complication: The IR spectrum is very complicated and the interpretation depends on lots of experience.
Sometimes, we cannot definitely clarify the structure of the compound just based on one single IR spectrum. Other
spectroscopy methods, such as ( Mass Spectrometry) MS and ( Nuclear Magnetic Resonance) NMR, are still needed to
further interpret the specific structure.
Quantification: Infrared spectroscopy works well for the qualitative analysis of a large variety of samples, but quantitative
analysis may be limited under certain conditions such as very high and low concentrations.

Symmetry & IR Spectroscopy


One of the most importance applications of IR spectroscopy is structural assignment of the molecule depending on the
relationship between the molecule and observed IR absorption bands. Every molecule is corresponding to one particular
symmetry point group. Then we can predict which point group the molecule is belonging to if we know its IR vibrational
bands. Vice versa, we can also find out the IR active bands from the spectrum of the molecule if we know its symmetry. These
are two main applications of group theory. We'll take the following problem as an example to illustrate how this works.

Question
How do you distinguish whether the structure of transition metal complex molecule M(CO)2L4 is cis or trans by inspection of
the CO stretching region of the IR spectra?

Answer
For cis-M(CO)2L4, the symmetry point group of this molecule is C2v.

C2v E C2 σ (xz) σ (yz)

γ co 2 0 2 0

γ co = A1 + B1
Since A1 has a basis on z axis and B1 has a basis on x axis, there are two IR vibrational bands observed in the spectrum.
For trans-M(CO)2L4, the symmetry point group of this molecule is D4h.

D4h E C4 C2 C2' C2" i S4 σh σv σd

γ co 2 2 2 0 0 0 0 0 2 2

γ co = A1g + A2u
Since A2u has a basis on z axis, there is only one IR vibrational band observed in the spectrum.
Therefore, from what have been discussed above, we can distinguish these two structures based on the number of IR bands.

Problems
The frequency of C=O stretching is higher than that of C=C stretching. The Intensity of C=O stretching is stronger than that of
C=C stretching. Explain it.

References

7/5/2021 9 https://chem.libretexts.org/@go/page/1847
1. D. A. Skoog, F. J. Holler, S. R. Crouch. Principles of Instrumental Analysis, 6th ed. Belmont, CA. Thomson Higher
Education. 2007
2. G. D. Christain. Analytical Chemistry, 5th ed. New York. John Wiley & Sons, INC. 1994
3. R. S. Drago. Physical Methods, 2nd ed. Mexico.Saunders College Publishing.1992
4. S.M. Blinder. Introduction to Quantum Mechanics. Academic Press. 2004
5. D. C. Harris, M. D. Bertolucci. Symmetry and Spectroscopy: An Introduction to Vibrational and Electronic Spectroscopy.
New York. Dover Publications, INC

Contributors and Attributions


Richard Osibanjo, Rachael Curtis, Zijuan Lai

7/5/2021 10 https://chem.libretexts.org/@go/page/1847
Interpreting Infrared Spectra
This chapter will focus on infrared (IR) spectroscopy. The wavelengths found in infrared radiation are a little longer than those
found in visible light. IR spectroscopy is useful for finding out what kinds of bonds are present in a molecule, and knowing
what kinds of bonds are present is a good start towards knowing what the structure could be.

Topic hierarchy

Contributors
Chris P Schaller, Ph.D., (College of Saint Benedict / Saint John's University)

Chris Schaller 7/4/2021 1 https://chem.libretexts.org/@go/page/4165


Carbon Nitrogen Bonds
The IR spectra of nitrogen-containing compounds can be messier than the ones you have seen so far. N-H bends and C-N
stretches tend to be broader and weaker than peaks involving oxygen atoms. However, some peaks in nitrogen compounds are
useful. The problems in this section will guide you through some of these features.

Contributors and Attributions


Chris P Schaller, Ph.D., (College of Saint Benedict / Saint John's University)

Chris Schaller 7/7/2021 1 https://chem.libretexts.org/@go/page/4174


IR10. More Practice with IR Spectra
Contributors and Attributions
Chris P Schaller, Ph.D., (College of Saint Benedict / Saint John's University)

Chris Schaller 6/28/2021 1 https://chem.libretexts.org/@go/page/4167


IR11. Appendix: IR Table of Organic Compounds
Table of IR Absorptions Common. Note: strong, medium, weak refers to the length of the peak (in the y axis direction). Note:
spectra taken by ATR method (used at CSB/SJU) have weaker peaks between 4000-2500 cm-1 compared to reference spectra
taken by transmittance methods (typical on SDBS and other sites).

Approximate Frequency (cm-1) Description Bond Vibration Notes

much broader, lower frequency


3500 - 3200 broad, round O-H (3200-2500)
if next to C=O
3400-3300 weak, triangular N-H stronger if next to C=O
3300 medium-strong =C-H (sp C-H)
can get bigger if lots of bonds
3100-3000 weak-medium =C-H (sp2 C-H)
present
can get bigger if lots of bonds
3000-2900 weak-medium -C-H (sp3 C-H)
present
2800 and 2700 medium C-H in O=C-H two peaks; "alligator jaws"
2250 medium C=N
stronger if near electronegative
2250-2100 weak-medium C=C
atoms
lower frequency (1650-1550)
if attached to O or N
middle frequency if attached to C,
1800-1600 strong C=O
H
higher frequency (1800) if
attached to Cl
lower frequency (1600-1450) if
conjugated
1650-1450 weak-medium C=C
(i.e. C=C-C=C)
often several if benzene present
1450 weak-medium H-C-H bend
higher frequency (1200-1300) if
1300 - 1000 medium-strong C-O conjugated
(i.e. O=C-O or C=C-O)
1250-1000 medium C-N
1000-650 strong C=C-H bend often several if benzene present

Contributors and Attributions


Chris P Schaller, Ph.D., (College of Saint Benedict / Saint John's University)

Chris Schaller 6/16/2021 IR11. Appendix.1 https://chem.libretexts.org/@go/page/4168


IR2. Hydrocarbon Spectra
All organic and biological compounds contain carbon and hydrogen, usually with various other elements as well.
Hydrocarbons are compounds containing only carbon and hydrogen, but no other types of atoms. Since all organic compounds
contain carbon and hydrogen, looking at hydrocarbon spectra will tell us what peaks are due to the basic C&H part of these
molecules. It is sometimes useful to think of the C&H part of a molecule as the basic skeleton or scaffolding used to construct
the molecule. The other atoms often form more interesting and active features, like the doors, windows and lights on a
building.
The simplest hydrocarbons contain only single bonds between their carbons, and no double or triple bonds. These
hydrocarbons are variously referred to as saturated hydrocarbons, paraffins or alkanes. Examples of alkanes include hexane
and nonane. (You can take a look at the Glossary to see what these names tell you about the structure.)

Look at the IR spectrum of hexane. You should see:


a set of peaks dipping down from the baseline at about 2900 cm-1.
another set of peaks dipping down from the baseline at about 1400-1500 cm-1.

Figure IR2.IR spectrum of hexane.


Source: SDBSWeb : http://riodb01.ibase.aist.go.jp/sdbs/ (National Institute of Advanced Industrial Science and Technology of
Japan, 14 July 2008)
If you look at an IR spectrum of any other alkane, you will also see peaks at about 2900 and 1500 cm-1. The IR spectra of
many organic compounds will show these peaks because the compound may contain paraffinic parts in addition to parts with
other elements in them.
These two kinds of peaks tell you that C-H bonds are present.
Specifically, the bonds involve sp3 or tetrahedral carbons.
Stretching C-H bonds in alkanes absorb light at around 2900 cm-1.
Bending H-C-H angles in alkanes absorb light at around 1500 cm-1.

Contributors and Attributions


Chris P Schaller, Ph.D., (College of Saint Benedict / Saint John's University)

Chris Schaller 6/23/2021 1 https://chem.libretexts.org/@go/page/4169


IR3. Subtle Points of IR Spectroscopy
Alkanes show two sets of peaks in the IR spectrum. Alkanes contain two kinds of bonds: C-C bonds and C-H bonds. However,
these two facts are not related. The reasons are explained through bond polarity and molecular vibrations.
Bond polarity can play a role in IR spectroscopy.
nature rules that only bonds that contain dipoles can absorb infrared light.
C-C bonds are usually nonpolar and usually do not show up as peaks in the IR spectrum.
C-H bonds are not very polar and do not give rise to strong peaks in the IR spectrum.
a whole lot of small C-H peaks can add up together to look like one big peak. This would happen if a molecule contained
many C-H bonds (a common situation).
Molecular vibrations play a major role in IR spectroscopy.
IR light interacts with vibrating bonds. When light is absorbed, the bond has a little more energy and vibrates at a higher
frequency.
a bond does not have an exact, fixed length; it can stretch and compress. This is called a bond stretching vibration.
Stretching C-H bonds in alkanes absorb light at around 2900 cm-1.
bond angles can also bend; for instance, the H-C-H bond angle can compress and stretch. This is called a bending
vibration.
Bending H-C-H angles in alkanes absorb light at around 1500 cm-1.
The factors that govern what bonds (and what vibrations) show up at what frequencies are easily handled by computational
chemistry software. In fact, prediction of absorption frequencies in IR spectra can be done using 17th century classical
mechanics, specifically Hooke's Law (devised to explain the vibrational frequencies of springs). Computation is not the focus
of this chapter but it may help you keep track of what kinds of vibrations absorb at what frequencies.
Hooke's Law states:
the vibrational frequency is proportional to the strength of the spring; the stronger the spring, the higher the frequency.
the vibrational frequency is inversely proportional to the masses at the ends of the spring; the lighter the weights, the higher
the frequency.
IR light is absorbed if it is in resonance with a vibrating bond; that means the light's frequency is the same as the frequency of
the bond vibration, or else an exact multiple of it (2x, 3x, 4x...). It's a little like pushing a child on a swing: unless you are
pushing at the same frequency that the swing is swinging, you will not be able to transfer your energy to the swing.
Hooke's Law in IR spectroscopy means:
stronger bonds absorb at higher frequencies.
weaker bonds absorb at lower frequencies.
bonds between lighter atoms absorb at higher frequencies.
bonds between heavier atoms absorb at lower frequencies.
Remember, there are two factors here, so you won't be able to make predictions knowing only one factor. Some strong bonds
may not absorb at high frequency because they are between heavy atoms. The information is presented mostly to help you
organize what bonds absorb at what general frequencies after you have learned about them.
The reasons explaining why C-H bending vibrations are at lower frequency than C-H stretching vibrations are also related to
Hooke's Law. An H-C-H bending vibration involves three atoms, not just two, so the mass involved is greater than in a C-H
stretch. That means lower frequency. Also, it turns out that the "stiffness" of a bond angle (analogous to the strength of a
spring) is less than the "stiffness" of a bond length; the angle has a little more latitude to change than does the length. Both
factors lead to a lower bending frequency.

Contributors and Attributions


Chris P Schaller, Ph.D., (College of Saint Benedict / Saint John's University)

Chris Schaller 6/25/2021 1 https://chem.libretexts.org/@go/page/4170


IR4. Carbon Carbon Multiple Bonds
Unsaturated hydrocarbons contain only carbon and hydrogen, but also have some multiple bonds between carbons. One type
of unsaturated hydrocarbon is an olefin, also known as an alkene. Alkenes contain double bonds between carbons. One
example of an alkene is 1-heptene. It looks similar to hexane, except for the double bond from the first carbon to the second.

Look at the IR spectrum of 1-heptene. You should see:


a set of peaks dipping down from the baseline at about 2900 cm-1.
another set of peaks dipping down from the baseline at about 1500 cm-1.

Figure IR3. IR spectrum of 1-heptene.


Source: SDBSWeb : http://riodb01.ibase.aist.go.jp/sdbs/ (National Institute of Advanced Industrial Science and Technology of
Japan, 14 July 2008)
So far, these peaks are the same as the ones seen for hexane. We can assign them as the C-H stretching and bending
frequencies, respectively.
Looking further, you will also see:
a small peak around 3100 cm-1.
a small peak near 1650 cm-1.
medium peaks near 800 and 1000 cm-1.
The peak at 3100 cm-1 hardly seems different from the C-H stretch seen before. It is also a C-H stretch, but from a different
type of carbon. This stretch involves the sp2 or trigonal planar carbon of the double bond, whereas the peak at 2900 involves
an sp3 or tetrahedral carbon.
The peak at 1650 cm-1 can be identified via computational methods as arising from a carbon-carbon double bond stretch. It is a
weak stretch because this bond is not very polar. Sometimes it is obscured by other, larger peaks.
The larger peaks near 800 and 1000 cm-1 are bending vibrations. They are due to a C=C-H bond angle that bends out of the
plane of the double bond (remember that the carbons on either end of the double bond are trigonal planar). They are called oop
bends. Oop bends are often prominent in alkenes and are easier to spot than an sp2 C-H stretching mode or a C=C stretching
mode.

Contributors and Attributions


Chris P Schaller, Ph.D., (College of Saint Benedict / Saint John's University)

Chris Schaller 7/4/2021 1 https://chem.libretexts.org/@go/page/4171


IR5. Carbon Oxygen Single Bonds
These bonds are pretty polar, so they show up strongly in IR spectroscopy. IR spectroscopy is therefore a good way to
determine what heteroatom-containing functional groups are present in a molecule.

Contributors and Attributions


Chris P Schaller, Ph.D., (College of Saint Benedict / Saint John's University)

Chris Schaller 6/16/2021 1 https://chem.libretexts.org/@go/page/4172


IR6. Carbon Oxygen Double Bonds
The largest class of oxygen-containing molecules is carbonyl compounds, which contain C=O bonds. A C=O stretch is
normally easy to find in an IR spectrum, because it is very strong and shows up in a part of the spectrum that is not cluttered
with other peaks. Examples of carbonyl compounds include 2-octanone, a ketone, and butanal, an aldehyde. In an aldehyde,
the carbonyl is at the end of a chain, with a hydrogen attached to the carbonyl carbon.

If you look at the IR spectrum of 2-octanone:


there are sp3 C-H stretching and CH2 bending modes at 2900 and 1500 cm-1.
there is a very strong peak around 1700 cm-1.

Figure IR10. IR spectrum of 2-octanone.


Source: SDBSWeb : http://riodb01.ibase.aist.go.jp/sdbs/ (National Institute of Advanced Industrial Science and Technology of
Japan, 14 July 2008)
Even though there is just one C=O bond, the carbonyl stretch is often the strongest peak in the spectrum. That makes carbonyl
compounds easy to identify by IR spectroscopy.
If you look at the IR spectrum of butanal:
there are sp3 C-H stretching and CH2 bending modes at 2900 and 1500 cm-1.
there is a very strong C=O peak around 1700 cm-1.
there is a pair of medium peaks around 2700 and 2800 cm-1. This is the aldehyde C-H stretching mode.

Figure IR11. IR spectrum of butanal.

Chris Schaller 7/4/2021 1 https://chem.libretexts.org/@go/page/4173


Source: SDBSWeb : http://riodb01.ibase.aist.go.jp/sdbs/ (National Institute of Advanced Industrial Science and Technology of
Japan, 14 July 2008)
The aldehyde C-H bond absorbs at two frequencies because it can vibrate in phase with the C=O bond (a symmetric stretch)
and out of phase with the C=O bond (an asymmetric stretch), and these vibrations are of different energies. The probability of
the symmetric stretch and the asymmetric stretch are about equal, so the two peaks are always about the same size. This
unusual C-H peak can often be used to distinguish between an aldehyde and a ketone.

Contributors and Attributions


Chris P Schaller, Ph.D., (College of Saint Benedict / Saint John's University)

Chris Schaller 7/4/2021 2 https://chem.libretexts.org/@go/page/4173


IR8. More Complicated IR Spectra
Sometimes more complicated heteroatomic functional groups, containing bonds to more than one heteroatom, have slightly
different spectra. Carboxylic acids feature a hydroxyl group bonded to a carbonyl. Hexanoic acid, a carboxylic acid in a six-
atom chain, is one example.

If you look at the IR spectrum of hexanoic acid:


there are CH2 bending modes at 1500 cm-1.
there is a very strong C=O peak around 1700 cm-1.
there is a medium C-O peak around 1250 cm-1.
the sp3 C-H and O-H stretching modes are less clear.

Figure IR12. IR spectrum of hexanoic acid.


Source: SDBSWeb : http://riodb01.ibase.aist.go.jp/sdbs/ (National Institute of Advanced Industrial Science and Technology of
Japan, 14 July 2008)
At first, the O-H peak appears to be absent. The C-H stretch appears to be very broad. The wide peak between 3000 and 2600
cm-1 is really the usual C-H stretch with a broad O-H stretch superimposed on it. The low frequency vibration of this O-H
bond is related to the partial dissociation of protons due to strong hydrogen bonding.

Contributors and Attributions


Chris P Schaller, Ph.D., (College of Saint Benedict / Saint John's University)

Chris Schaller 7/7/2021 1 https://chem.libretexts.org/@go/page/4175


IR9. Misleading Peaks
There are some practical problems that can make IR interpretation in real life more difficult. Being aware of these problems
may make you double-check your suspicions:
water in the sample. Since water contains O-H bonds, water in a sample will make it appear as if the compound contains O-
H bonds. There are experimental techniques for removing water from a sample but they must be done carefully.
overtones in a spectrum. Overtones are absorptions occurring at different multiples of the normal frequency. Strong
overtones of carbonyl peaks often occur at about twice the normal wavenumber. A large peak at 1750 cm-1 might be
accompanied by a smaller peak at 3500 cm-1, and could be confused with an O-H peak.
In addition, there are complications that you may run into based on the instrument or technique used to obtain the spectrum.
Note that the x-axis scale on the spectrum is not uniform: the scale is indexed every 1000 cm-1 between 4000 and 2000 cm-
1
and every 500 cm-1 below that. Spectra from different instruments may display this scale differently.
Different techniques may lead to increased absorption at one end of the spectrum vs. the other. For example, an OH stretch
is much more dominant if the sample is prepared in a KBr mull than if the spectrum is obtained via attenuated total
reflectance.

Contributors and Attributions


Chris P Schaller, Ph.D., (College of Saint Benedict / Saint John's University)

Chris Schaller 7/7/2021 1 https://chem.libretexts.org/@go/page/4176


What Does an IR Spectrum Look Like?
A spectrum is a graph in which the amount of light absorbed is plotted on the y-axis and frequency is plotted on the x-axis. An
example is shown below. You can run your finger along the graph and see whether any light of a particular frequency is
absorbed; if so, you will see a "peak" at that frequency. If not, you will see "the baseline" at that frequency.

Figure IR1. IR spectrum of benzene. The x-axis labels are, from right to left, 500, 1000, 1500, 2000, 3000 and 4000 cm-1.
Source: SDBSWeb : http://riodb01.ibase.aist.go.jp/sdbs/ (National Institute of Advanced Industrial Science and Technology of
Japan, 14 July 2008)
In IR spectra (spectra = plural of spectrum):
the y axis is usually labeled "transmittance". Transmittance is the amount of light that passes through the sample.
the unit of transmittance is percent (%).
the x axis is labeled "wavenumbers". Wavenumbers are proportional to frequency, so the higher the frequency, the higher
the wavenumber.
the symbol for wavenumber is reciprocal centimeters (cm-1).
the x-axis is usually displayed with high wavenumber on the left and lower wavenumber on the right.
As you run your finger from left to right across an IR spectrum, you can see whether or not light is absorbed at particular
frequencies. When the curve dips down, less light is transmitted. That means light is absorbed. The dip in the graph is called a
peak. Different bonds absorb different frequencies of light, so the peaks tell you what kinds of bonds are present.

Contributors and Attributions


Chris P Schaller, Ph.D., (College of Saint Benedict / Saint John's University)

Chris Schaller 6/21/2021 1 https://chem.libretexts.org/@go/page/4166


IR Spectroscopy Background
This page describes what an infra-red spectrum is and how it arises from bond vibrations within organic molecules.

How an infra-red spectrum is produced


You probably know that visible light is made up of a continuous range of different electromagnetic frequencies - each
frequency can be seen as a different color. Infra-red radiation also consists of a continuous range of frequencies - it so happens
that our eyes can't detect them.
If you shine a range of infra-red frequencies one at a time through a sample of an organic compound, you find that some
frequencies get absorbed by the compound. A detector on the other side of the compound would show that some frequencies
pass through the compound with almost no loss, but other frequencies are strongly absorbed.
How much of a particular frequency gets through the compound is measured as percentage transmittance. A percentage
transmittance of 100 would mean that all of that frequency passed straight through the compound without any being absorbed.
In practice, that never happens - there is always some small loss, giving a transmittance of perhaps 95% as the best you can
achieve. A transmittance of only 5% would mean that nearly all of that particular frequency is absorbed by the compound. A
very high absorption of this sort tells you important things about the bonds in the compound.

What an infra-red spectrum looks like


A graph is produced showing how the percentage transmittance varies with the frequency of the infra-red radiation.

Notice that an unusual measure of frequency is used on the horizontal axis. Wavenumber is defined like this:

Similarly, don't worry about the change of scale half-way across the horizontal axis. You will find infra-red spectra where the
scale is consistent all the way across, infra-red spectra where the scale changes at around 2000 cm-1, and very occasionally
where the scale changes again at around 1000 cm-1. As you will see when we look at how to interpret infra-red spectra, this
does not cause any problems - you simply need to be careful reading the horizontal scale.

What causes some frequencies to be absorbed?


Each frequency of light (including infra-red) has a certain energy. If a particular frequency is being absorbed as it passes
through the compound being investigated, it must mean that its energy is being transferred to the compound.
Energies in infra-red radiation correspond to the energies involved in bond vibrations.

Bond stretching
In covalent bonds, atoms aren't joined by rigid links - the two atoms are held together because both nuclei are attracted to the
same pair of electrons. The two nuclei can vibrate backwards and forwards - towards and away from each other - around an
average position.
The diagram shows the stretching that happens in a carbon-oxygen single bond. There will, of course, be other atoms attached
to both the carbon and the oxygen. For example, it could be the carbon-oxygen bond in methanol, CH3OH.

Jim Clark 6/25/2021 1 https://chem.libretexts.org/@go/page/3732


The energy involved in this vibration depends on things like the length of the bond and the mass of the atoms at either end.
That means that each different bond will vibrate in a different way, involving different amounts of energy.
Bonds are vibrating all the time, but if you shine exactly the right amount of energy on a bond, you can kick it into a higher
state of vibration. The amount of energy it needs to do this will vary from bond to bond, and so each different bond will absorb
a different frequency (and hence energy) of infra-red radiation.

Bond bending
As well as stretching, bonds can also bend. The diagram shows the bending of the bonds in a water molecule. The effect of
this, of course, is that the bond angle between the two hydrogen-oxygen bonds fluctuates slightly around its average value.
Imagine a lab model of a water molecule where the atoms are joined together with springs. These bending vibrations are what
you would see if you shook the model gently.

Again, bonds will be vibrating like this all the time and, again, if you shine exactly the right amount of energy on the bond,
you can kick it into a higher state of vibration. Since the energies involved with the bending will be different for each kind of
bond, each different bond will absorb a different frequency of infra-red radiation in order to make this jump from one state to a
higher one.

Tying all this together


Look again at the infra-red spectrum of propan-1-ol, CH3CH2CH2OH:

In the diagram, three sample absorptions are picked out to show you the bond vibrations which produced them. Notice that
bond stretching and bending produce different troughs in the spectrum.

Contributor
Jim Clark (Chemguide.co.uk)

Jim Clark 6/25/2021 2 https://chem.libretexts.org/@go/page/3732


The Fingerprint Region
This page explains what the fingerprint region of an infra-red spectrum is, and how it can be used to identify an organic
molecule.

Contributors and Attributions


Jim Clark (Chemguide.co.uk)

Jim Clark 6/29/2021 1 https://chem.libretexts.org/@go/page/3733


CHAPTER OVERVIEW
RAMAN SPECTROSCOPY
Raman spectroscopy is a chemical instrumentation technique that exploits molecular vibrations.

RAMAN: APPLICATION
If one can extract all of the vibrational information corresponds a molecule, its molecular structure
can then be determined. In the field of spectroscopy, two main techniques are applied in order to
detect molecular vibrational motions: Infrared spectroscopy (IR) and Raman spectroscopy. Raman
Spectroscopy has its unique properties which have been used very commonly and widely.

RAMAN: THEORY
The phenomenon of Raman scattering of light was first postulated by Smekai in 1923 and first
observed experimentally in 1928 by Raman and Krishnan. Raman scattering is most easily seen as
the change in frequency for a small percentage of the intensity in a monochromatic beam as the result of coupling between the
incident radiation and vibrational energy levels of molecules. A vibrational mode will be Raman active only when it changes the
polariazbility of the moleculeat.

RESONANT VS. NONRESONANT RAMAN SPECTROSCOPY


In this section readers will be introduced to the theory behind resonance and non-resonance Raman spectroscopy. Each technique has
its share of advantages and challenges. Each of these aspects will be explored.

1 7/7/2021
Raman: Application
If one can extract all of the vibrational information corresponds a molecule, its molecular structure can then be determined. In
the field of spectroscopy, two main techniques are applied in order to detect molecular vibrational motions: Infrared
spectroscopy (IR) and Raman spectroscopy. Raman Spectroscopy has its unique properties which have been used very
commonly and widely in Inorganic, Organic, Biological systems [1] and Material Science [2], [3], etc.

Introduction
Generally speaking, vibrational and rotational motions are unique for every molecule. The uniqueness to molecules are in
analogous to fingerprint identification of people hence the term molecular fingerprint. Study the nature of molecular vibration
and rotation is particularly important in structure identification and molecular dynamics. Two of the most important techniques
in studying vibration/rotation information are IR spectroscopy and Raman spectroscopy. IR is an absorption spectroscopy
which measures the transmitted light. Coupling with other techniques, such as Fourier Transform, IR has been highly
successful in both organic and inorganic chemistry.
Unlike IR, Raman spectroscopy measures the scattered light (Figure 2). There are three types of scattered lights: Rayleigh
scattering, Stokes scattering, and anti-stokes scattering. Rayleigh scattering is elastic scattering where there is no energy
exchange between the incident light and the molecule. Stokes scattering happens when there is an energy absorption from the
incident light, while anti-stokes scattering happens when the molecule emites energy to the incident light. Thus, Stokes
scattering results in a red shift, while anti-stokes scattering results in a blue shift. (Figure 1) Stokes and Anti-Stokes scattering
are called Raman scattering which can provide the vibration/rotation information

The intensity of Rayleigh scattering is around 107 times that of Stokes scattering. [4]According to the Boltzmann distribution,
anti-Stokes is weaker than Stokes scattering. Thus, the main difficulty of Raman spectroscopy is to detect the Raman
scattering by filtering out the strong Rayleigh scattering. In order to reduce the intensity of the Rayleigh scattering, multiple
monochromators are applied to selectively transmit the needed wave range. An alternative way is to use Rayleigh filters. There
are many types of Rayleigh filters. One common way to filter the Rayleigh light is by interference.
Because of the weakness of Raman scattering, the resolving power of a Raman spectrometer is much higher than an IR
specctrometer. A resolution of 105 is needed in Raman while 103 is sufficient in IR. [5] In order to achieve high resolving
power, prisms, grating spectrometers or interferometers are applied in Raman instruments.
Despite the limitations above, Raman spectroscopy has some advantages over IR spectroscopy as follows:

5/22/2021 Raman.1 https://chem.libretexts.org/@go/page/1849


1. Raman Spectroscopy can be used in aqueous solutions (while water can absorb the infrared light strongly and affect the IR
spectrum).
2. Because of the different selection rules, vibrations inactive in IR spectroscopy may be seen in Raman spectroscopy. This
helps to complement IR spectroscopy.
3. There is no destruction to the sample in Raman Spectroscopy. In IR spectroscopy, samples need to disperse in transparent
matrix. For example grind the sample in solid KBr. In RS, no such destructions are needed.
4. Glass vials can be used in RS (this should only work in the visible region. If in UV region, glass is not applicable because it
can strongly absorb light too.)
5. Raman Spectroscopy needs relative short time. So we can do Raman Spectroscopy detection very quickly.
After analysis of the advantages and disadvantages of Raman Spectroscopy technique, we can begin to consider the
application of Raman Spectroscopy in inorganic, organic, biological systems and Material Science, etc.

Applications
Raman Spectroscopy application in inorganic systems
X-ray diffraction (XRD) has been developed into a standard method of determining structure of solids in inorganic systems.
Compared to XRD, it is usually necessary to obtain other information (NMR, electron diffraction, or UV-Visible) besides
vibrational information from IR/Raman in order to elucidate the structure. Nevertheless, vibrational spectroscopy still plays an
important role in inorganic systems. For example, some small reactive molecules only exist in gas phase and XRD can only be
applied for solid state. Also, XRD cannot distinguish between the following bonds: –CN vs. –NC, –OCN vs. –NCO,–CNO vs.
–ONC, -SCN vs. –NCS. [7] Furthermore, IR and Raman are fast and simple analytical method, and are commonly used for the
first approximation analysis of an unknown compound.
Raman spectroscopy has considerable advantages over IR in inorganic systems due to two reasons. First, since the laser beam
used in RS and the Raman-scattered light are both in the visible region, glass (Pyrex) tubes can be used in RS. On the other
hand, glass absorbs infrared radiation and cannot be used in IR. However, some glass tubes, which contain rare earth salts, will
gives rises to fluorescence or spikes. Thus, using of glass tubes in RS still need to be careful. Secondly, since water is a very
weak Raman scatter but has a very broad signal in IR, aqueous solution can be directly analyzed using RS.
Raman Spectroscopy and IR have different selection rules. RS detects the polarizability change of a molecule, while IR detects
the dipole momentum change of a molecule. Principle about the RS and IR can be found at Chemwiki Infrared Theory and
Raman Theory. Thus, some vibration modes that are active in Raman may not be active IR, vice versa. As a result, both of
Raman and IR spectrum are provided in the stucture study. As an example, in the study of Xenon Tetrafluoride. There are 3
strong bands in IR and solid Raman shows 2 strong bands and 2 weaker bands. These information indicates that Xenon
Tetrafluoride is a planar molecule and has a symmetry of D4h. [8] Another example is the application of Raman Spectroscopy
in homonuclear diatomic molecules. Homonuclear diatomic molecules are all IR inactive, fortunately, the vibration modes for
all the homonuclear diatomic molecules are always Raman Spectroscopy active.

Raman Spectroscopy Application in Organic Systems


Unlike inorganic compounds, organic compounds have less elements mainly carbons, hydrogens and oxygens. And only a
certain function groups are expected in organic specturm. Thus, Raman and IR spectroscopy are widely used in organic
systems. Characteristic vibrations of many organic compounds both in Raman and IR are widely studied and summarized in
many literature. [5] Qualitative analysis of organic compounds can be done base on the characteristic vibrations table.
Table 1: Characteristic frequencies of some organic function group in Raman and IR
Vibration Region(cm-1) Raman intensity IR intensity

v(O-H) 3650~3000 weak strong

v(N-H) 3500~3300 medium medium


v(C=O) 1820~1680 strong~weak very strong
v(C=C) 1900~1500 very strong~medium 0~weak

“RS is similar to IR in that they have regions that are useful for functional group detection and fingerprint regions that permit
the identification of specific compounds.”[1] While from the different selection rules of Raman Spectroscopy and IR, we can

5/22/2021 Raman.2 https://chem.libretexts.org/@go/page/1849


get the Mutual Exclusion rule [5], which says that for a molecule with a center of symmetry, no mode can be both IR and
Raman Spectroscopy active. So, if we find a strong bond which is both IR and Raman Spectroscopy active, the molecule
doesn't have a center of symmetry.

Non-classical Raman Spectroscopy


Although classical Raman Spectroscopy has been successfully applied in chemistry, this technique has some major limitations
as follows[5]:
1. The probability for photon to undergo Raman Scattering is much lower than that of Rayleigh scattering, which causes low
sensitivity of Raman Spectroscopy technique. Thus, for low concentration samples, we have to choose other kinds of
techniques.
2. For some samples which are very easily to generate fluorescence, the fluorescence signal may totally obscure the Raman
signal. We should consider the competition between the Raman Scattering and fluorescence.
3. In some point groups, such as C6 , D6 , D6h , C4h , D2h, there are some vibrational modes that is neither Raman or IR active.
4. The resolution of the classical Raman Spectroscopy is limited by the resolution of the monochromator.
In order to overcome the limitations, special techniques are used to modify the classical Raman Spectroscopy. These non-
classical Raman Spectroscopy includes: Resonance Raman Spectroscopy, surface enhanced Raman Spectroscopy, and
nonlinear coherent Raman techniques, such as hyper Raman spectroscopy

Resonance Raman Scattering (RRS)


The resonance effect is observed when the photon energy of the exciting laser beam is equal to the energy of the allowed
electronic transition. Since only the allowed transition is affected, (in terms of group theory, these are the totally symmetric
vibrational ones.), only a few Raman bands are enhanced (by a factor of 106). As a result, RRS can increase the resolution of
the classical Raman Spectroscopy, which makes the detection of dilution solution possible (concentrations as low as 10-3 M).
RRS is extensively used for biological molecules because of its ability to selectively study the local environment. As an
example, the Resonance Raman labels are used to study the biologically active sites on the bond ligand. RRS can also be used
to study the electronic excited state. For example, the excitation profile which is the Raman intensity as a function of incident
laser intensity can tell the interaction between the electronic states and the vibrational modes. Also, it can be used to measure
the atomic displacement between the ground state and the excited state.

Surface Enhanced Raman Scattering (SERS)


At 1974, Fleischmann discovered that pyridine adsorbed onto silver electrodes showed enhanced Raman signals. This
phenomenon is now called surface enhanced Raman Scattering (SERS). Although the mechanism of SERS is not yet fully
understood, it is believed to result from an enhancement either of transition polarizability, α,or the electric field, E, by the
interaction with the rough metallic support.
Unlike RRS, SERS enhances every band in the Raman spectrum and has a high sensitivity. Due to the high enhancement (by a
factor of 1010~11), the SERS results in a rich spectrum and is an ideal tool for trace analysis and in situ study of interfacial
process. Also, it is a better tool to study highly diluted solutions. A concentration of 4x10-12 M was reported by Kneipp using
SERS. [5]

Nonlinear Raman Spectroscopy


In a nonlinear process, the output is not linearly proportional to its input. This happens when the perturbation become large
enough that the response to the perturbation doesn’t follows the perturbation’s magnitude. Nonlinear Raman Spectroscopy
includes: Hyper Raman spectroscopy, coherent anti-Stokes Raman Spectroscopy, coherent Stokes Raman spectroscopy,
stimulated Raman gain and inverse Ramen spectroscopy. Nonlinear Raman spectroscopy is more sensitive than classical
Raman spectroscopy and can effectively reduce/remove the influence of fluorescence. The following paragraph will focus on
the most useful nonlinear Raman spectroscopy---coherent anti-Stokes Raman Spectroscopy (CARS):

5/22/2021 Raman.3 https://chem.libretexts.org/@go/page/1849


Figure 3) The two laser sources generate a coherent beam at frequency ν3. There is a signal enhancement when ν3 equal the
anti-Stokes scattering (νa), and the vibrational transition equals to the energy difference between two light sources. Since
CARS signal is at anti-Stoke region (a higher energy region than fluorescence), the influence of fluorescence is eliminated.
Thus, CARS is very useful for molecules with high fluorescence effect, for example, some biological molecules. Another
important advantage of CARS is that the resolution is no longer limited by the monochromator as in classical Raman because
only the anti-Stokes frequency is studied in CARS. High-resolution CARS has been developed as a tool for small-time scale
process, such as photochemical analysis and chemical kinetics studies. [6]

References
1. Principles of Instrumental Analysis, fifth edition. Skoog, Holler and Nieman.
2. Infrared and Raman Spectra of Inorganic and Coordination Compounds, fifth edition. Kazuo Nakamoto.
3. Symmetry and Spectroscopy an introduction to vibrational and electronic spectroscopy. Daniel C. Harris, etc.
4. P. Bisson, G. Parodi, D. Rigos, J.E. Whitten, The Chemical Educator, 2006, Vol. 11, No. 2
5. B. Schrader, Infrared and Raman Spectroscopy, VCH, 1995, ISBN:3-527-26446-9
6. S.A. Borman, Analytical Chemistry, 1982, Vol. 54, No. 9, 1021A-1026A
7. K. Nakamoto, Infrared Spectra of Inorganic and Coordination Compounds, 3rd edition. Wiley Intrsc John Wiley & Sons,
New York London Sydney Toronto, 1978
8. H.H. Claassen, C.L. Chernick, J.G. Malm, 1963 J. Am. Chem. Soc., 85, 1927

Problems
1. What are the advantages and disadvantages for Raman spectroscopy, comparing with IR spectroscopy?
2. Please briefly explain the mutual exclusive principle in Raman and IR spectroscopy.

Contributors and Attributions


Xu Yuntao (UCD), Qiuting Hong (UCD)

5/22/2021 Raman.4 https://chem.libretexts.org/@go/page/1849


Raman: Theory
The phenomenon of Raman scattering of light was first postulated by Smekai in 1923 and first observed experimentally in 1928 by
Raman and Krishnan. Raman scattering is most easily seen as the change in frequency for a small percentage of the intensity in a
monochromatic beam as the result of coupling between the incident radiation and vibrational energy levels of molecules. A vibrational
mode will be Raman active only when it changes the polariazbility of the molecule.

Introduction
When monochromatic radiation with a wavenumber ν~ is incident on systems, most of it is transmitted without change, but, in addition,
0

some scattering of the radiation occurs. If the frequency content of the scattered radiation is analyzed, there will be observed to be
present not only the wavenumber ν~ associated with the incident radiation but also, in general, pairs of new wavenumbers of the type
0

~ ~
ν = ν +ν
0
~
M . In molecular systems, the wavenumbers ν~ are found to lie principally in the ranges associated with transitions
M

between rotational, vibrational, and electronic levels.


Such scattering of radiation with change of wavenumber is called Raman Scattering, after the Indian scientist C. V. Raman who, with K.
S. Krishnan, first observed this phenomenon in liquids in 1928. The effect had been predicted on theoretical grounds in 1923 by A.
Smekal. Due to its very low scattering efficiency, Raman spectroscopy did not become popular until powerful laser systems were
available after the 1960s. Now, Raman spectroscopy has become one of the most popular approaches to study the vibrational structures
of molecules together with infrared spectrum.
The origin of the modified frequencies found in Raman scattering is explained in terms of energy transfer between the scattering system
and the incident radiation. When a system interacts with radiation of wavenumber ν~ , it makes an upward transition from a lower
0

energy level E1 to an upper energy level E2. It must then acquire the necessary energy, ΔE= E2- E1, from the incident radiation. The
energy ΔE is expressed in terms of a wavenumber ν~ associated with the two levels involved, where
M

~
ΔE = hcν M (Raman.1)

This energy requirement is regarded as being provided by the absorption of one photon of the incident radiation of energy hcν~ and the0

simultaneous emission of a photon of smaller energy hc (ν~ − ν~ ) , so that scattering of radiation of lower wavenumber, ν~ − ν~ ,
0 M 0 M

occurs. Alternatively, the interaction of the radiation with the system may cause a downward transition from a higher energy level E2 to
a lower energy level E1, in which case it makes available energy
~
E2 − E1 = hc ν M (Raman.2)

Again a photon of the incident radiation of energy ~


hcν 0 and the simultaneous emission of a photon of higher energy ~ ~
hc (ν 0 + ν M ) , so
that scattering of radiation of higher wavenumber, ν~ 0
~
+νM , occurs.
In the case of Rayleigh scattering, although there is no resultant change in the energy state of the system, the system still participates
directly in the scattering act, causing one photon of incident radiation hcν~ to be absorbed and a photon of the same energy to be
0

emitted simultaneously, so that scattering of radiation of unchanged wavenumber, ν~ , occurs.


0

It is clear that, as far as wavenumber is concerned, a Raman band is to be characterized not by its absolute wavenumber,

~ ~
ν = ν ±ν
0
~
M , but by the magnitude of its wavenumber shift ν~ from the incident wavenumber. Such wavenumber shifts are often
M

referred to as Raman wavenumbers. Where it is necessary to distinguish Stokes and anti-Stokes Raman scattering we shall define Δν~ to

be positive for Stokes scattering and negative for anti-Stokes scattering, that is Δν~ = ν~ + ν~ (see Fig. 1).
0

Figure 1. Diagrammatic representation of an energy transfer model of Rayleigh scattering, Stokes Raman and anti-Stokes Raman
scattering.

5/28/2021 Raman.1 https://chem.libretexts.org/@go/page/1850


The intensity of anti-Stokes relative to Stokes Raman scattering decreases rapidly with increase in the wavenumber shift. This is
because anti-Stokes Raman scattering involves transitions to a lower energy state from a populated higher energy states.

Classical theory
According to the classical theory of electromagnetic radiation, electric and magnetic fields oscillating at a given frequency are able to
give out electromagnetic radiation of the same frequency. One could use electromagnetic radiation theory to explain light scattering
phenomena.
For a majority of systems, only an induced electric dipole moment μ is taken into consideration. This dipole moment which is induced
by the electric field E could be expressed by the power series
(1) (2) (3)
μ =μ +μ +μ +⋯ (Raman.3)

where
(1)
μ =α⋅E (Raman.4)

(2) 1
μ = β ⋅ EE
2

(3) 1
μ = γ ⋅ EEE
6

α is termed the polarizability tensor. It is a second-rank tensor with all the components in the unit of CV-1m2. Typically, orders of
magnitude for components in α, β, and γ are as follows, α, 10-40 CV-1m2; β, 10-50 CV-2m3; and γ, 10-61 CV-3m4. According to the values,
the contributions of μ(2)and μ(3) are quite small unless electric field is very high. Since Rayleigh and Raman scattering are observed
quite readily with very much lower electric field intensities, one may expect to explain Rayleigh and Raman scattering in terms of μ(1)
only.
We shall now consider the interaction of a molecular system with the harmonically oscillating electric field in the frequency ω0. To
make the explanation easily, we shall ignore the rotation but just consider the vibration part. It is to be expected that the polarizability
will be a function of the nuclear coordinates. The variation of components in polarizability tensor with vibrational coordinates is
expressed in a Taylor series
2
∂αij ∂ αij
1
αij = (αij ) +∑ ( ) Qk + ∑ ( ) Qk Ql + ⋯
0 k ∂Qk 2 k,l ∂ Qk ∂ Ql
0
0

where (αij)0 is the αij value at the equilibrium configuration, Qk, Ql are normal coordinates of vibration at frequencies ωk, ωl. We shall
make a harmonic approximation to neglect the terms which involve powers of Q higher than first. After initially fixing our attention on
one normal mode, Qk, we could get
∂αij
αij = (αij ) +( ) Qk
0 ∂Qk
0

As for a harmonic vibration,


Qk = Qk0 cos(ωk t + δk ) (Raman.5)

Then we could get the expression of α tensor resulting from k-th vibration,
∂αk
αk = α0 + ( ) Qk0 cos(ωk t + δk )
∂Q
k 0

Now, under the influence of electromagnetic radiation at frequency ω0, the induced electric dipole moment μ(1) is expressed,
∂αk
(1)
μ = αk ⋅ E0 cos ω0 t = α0 ⋅ E0 cos ω0 t + ( ) ⋅ E0 Qk0 cos ω0 t cos(ωk t + δk )t = (Raman.6)
∂Qk
0

1 ∂αk 1 ∂αk
α0 ⋅ E0 cos ω0 t + ( ) ⋅ E0 Qk0 cos(ω0 t + ωk t + δk ) + ( ) ⋅ E0 Qk0 cos(ω0 t − ωk t − δk ) (Raman.7)
2 ∂Qk 2 ∂Qk
0 0

We see that the linear induced dipole moment μ(1) has three components with different frequencies,
α0 ⋅ E0 cos ω0 t which gives rise to radiation at ω0 and accounts for the Rayleigh scattering;
1 ∂αk
( ) ⋅ E0 Qk0 cos(ω0 t + ωk t + δk )
2 ∂Qk
0

which gives rise to radiation at ω0+ωk and accounts for the anti-Stokes Raman scattering; and
1 ∂αk
( ) ⋅ E0 Qk0 cos(ω0 tωk tδk ) (Raman.8)
2 ∂Qk 0

5/28/2021 Raman.2 https://chem.libretexts.org/@go/page/1850


which gives rise to radiation at ω0-ωk and accounts for the Stokes Raman scattering.
From these mathematical manipulations, there emerges a useful qualitative picture of the mechanisms of Rayleigh and Raman
scattering in terms of classical radiation theory. Rayleigh scattering comes from the dipole oscillating at ω0 induced in the molecule by
the electric field of the incident radiation at frequency ω0. Raman scattering arises from the dipole moment oscillating at ω0±ωk
produced by the modulation of dipole oscillating at ω0 with molecular vibration at frequency ωk. In other words, the frequencies we
observe in Raman scattering are beat frequencies of the radiation frequency ω0 and the molecular vibrational frequency ωk.

Quantum mechanical treatment


According to the quantum theory, radiation is emitted or absorbed as a result of a system making a downward or upward transition
between two discrete energy levels. A quantum theory of spectroscopic processes should, therefore, treat the radiation and molecule
together as a complete system, and explore how energy is transferred between the radiation and the molecule as a result of their
interaction.
A transition between energy levels of the molecular systems takes place with the emission or absorption of radiation, provided a
transition moment associated with the initial and final molecular states is non-zero. The transition moment could be defined as
Mf i = ⟨Ψf ∣ μ ∣ Ψi ⟩

in the Dirac bracket notation, where Ψi and Ψf are the wave function of initial and final states, respectively, and μ is the dipole moment
operator.
As we have discussed in the classical part, the linear induced dipole moment could be expressed by
(1)
μ =α⋅E (Raman.9)

Therefore, in quantum mechanical treatment, if a transition from an initial state to a final state is induced by incident radiation at
frequency ω0, the transition moment is given by
(1)
μ = ⟨Ψf ∣ α ∣ Ψi ⟩ ⋅ E (Raman.10)
fi

Now we will examine in more detail the nature of a typical matrix element of the polarizability tensor, like [αxy]fi, for Raman
scattering. Just like in the classical theory, we will ignore the rotational wave function and consider the vibrational part only.
[ αxy ] = ⟨Φf ∣ αxy ∣ Φi ⟩ (Raman.11)
fi

where Φ is the vibrational wave function.


In classical theory,
2
∂αxy 1 ∂ αij
αxy = (αxy ) +∑( ) Qk + ∑( ) Qk Ql + ⋯ (Raman.12)
0
∂Qk 2 ∂ Qk ∂ Ql
k 0 k,l
0

Introducing the quantum part, we obtain (just consider the first order term)
∂αxy
[ αxy ] = (αxy ) ⟨Φf ∣ Φi ⟩ + ∑ ( ) ⟨Φf ∣ Qk ∣ Φi ⟩
fi 0 k ∂Q
k 0

In the harmonic oscillator model, the total vibrational wave function is product of the harmonic oscillator wave functions for each of the
normal modes of vibration. Thus, for Φ,

Φi = ∏ Φvi (Qk ) (Raman.13)


k

where Φ (Q ) is the harmonic oscillator wave function associated with the normal coordinate Qk, which has a vibrational quantum
v
i
k
k

number vki in a certain state. For harmonic oscillator functions, we have


f i
0 for v ≠v
k k
⟨Φ f (Qk ) ∣ Φvi (Qk )⟩ = { (Raman.14)
v k f i
k
1 for v =v
k k

and

5/28/2021 Raman.3 https://chem.libretexts.org/@go/page/1850


f i

⎪ 0 f or v =v


k k
⎪ 1
f
⟨Φ f (Qk ) ∣ Qk ∣ Φ i (Qk )⟩ = ⎨ (vi + 1) 2
bv f or v =v
i
+1 (Raman.15)
v v k k k k
k k





1
f
i 2 i
(v ) bvk f or v =v −1
k k k

where
−−−−−
h
bvk = √ (Raman.16)
8π 2 vk

Selection rules
We are now able to find out the conditions which have to be satisfied if the transition moment is non-zero. We will consider the zero
order term which accounts for the Rayleigh scattering part first. This term is none-zero only if v = v which means none of the f

k
i

vibrational quantum numbers change during this transition from initial state i to final state f. Thus, for Rayleigh scattering, the quantum
mechanical treatment and the classical theory give the same results.
∂αxy
Then we could go to the first order term ∑
k
(
∂Qk
) ⟨Φf ∣ Qk ∣ Φi ⟩ which accounts for the Raman scattering part. For the k-th
0

summand, it is zero unless every term in the product is non-zero, and to achieve this the following conditions must be satisfied, for all
f
modes except the k-th: the vibrational quantum numbers must be the same during the transition, i.e. v = v where j≠k; and, for the k-th j
i
j

mode, and the vibrational quantum number must change by one unit, i.e. v = v ± 1 . The transition moment is associated with Stokes
f

k
i
k

Raman scattering for Δvk=1, and with anti-Stokes Raman scattering for Δvk=-1. These conditions are a result of the properties of
harmonic oscillator wave functions.
It follows from these arguments that, in the harmonic approximation, only vibrational fundamentals, i.e., transitions with only one
vibrational quantum number changes by one unit, can be observed in the Raman scattering. However, this Δvk=±1 restriction is only a
necessary but not a sufficient condition for the occurrence of Raman scattering at the k-th vibrational mode. The vibrational mode
should be Raman active, i.e., at least one of the elements of the derived polarizability tensor should be non-zero.
It can be rigorously established by group theory that the elements of the derived polarizability will be non-zero only if they have the
same symmetry with the second order terms, i.e., x2, y2, z2, xy, yz, xz. In other words, the irreducible representation of a certain
vibrational mode should have a basis in x2, y2, z2, xy, yz or xz.
Finally, we could establish a much better basis for determining selection rules for vibrational transitions in the Raman effect, if we
consider the properties of the vibrational transition polarizability components, rather than the derived prolarizability tensor components.
For fundamental vibrational transitions, where in the initial state all vibrational quantum numbers are zero and in the final state only the
k-th vibrational quantum number has changed to unity,
[ αxy ] = ⟨Φ1 ∣ αxy ∣ Φ0 ⟩
fi

According to the group theory, this integral will be non-zero, only if αxy and Φ1(Qk) belong to the same symmetry species, which
implies that, under each symmetry operation of the molecule in question, αxy and Φ1(Qk) transform in the same way. This constitutes a
general selection rule for the Raman activity of a fundamental transition.
In its most general way, covering all types of transitions, the selection rule is as follows: a transition between two states, Ψiand Ψf, is
Raman forbidden unless at least one of the triple products of the type ΨiαxyΨf belongs to a representation whose structure contains the
totally symmetric species.

References
1. Long, D.A., Raman Spectoscopy, McGraw-Hill, Inc., New York, 1977.
2. Szymanski, H. A., Raman Spectroscopy: Theory and Practice, Plenum Press, New York, 1967.
3. McCreery, R. L., Raman Spectroscopy for Chemical Analysis, John Wiley & Sons, Inc., New York, 2000.
4. Stencel, J. M., Raman Spectroscopy for Catalysis, Van Nostrand Reinhold, New York, 1990.
5. Grasselli, J. G.; Bulkin, B. J., Analytical Raman Spectroscopy, John Wiley & Sons, Inc., New York, 1991.
6. Harris, D. C.; Bertolucci, M. D.; Symmetry and Spectroscopy: an introduction to vibrational and electronic spectroscopy, Dover
Publications, Inc., New York, 1989.

Outside links
http://en.Wikipedia.org/wiki/Raman_spectroscopy

5/28/2021 Raman.4 https://chem.libretexts.org/@go/page/1850


Contributors and Attributions
Yang Liu (Department of Chemistry, UC Davis), Hong Gao (Department of Chemistry, UC Davis)

5/28/2021 Raman.5 https://chem.libretexts.org/@go/page/1850


Resonant vs. Nonresonant Raman Spectroscopy
Raman spectroscopy is a chemical instrumentation technique that exploits molecular vibrations. It does not require large
sample sizes and is non-destructive to samples. It is capable of qualitative analysis of samples and the intensity of spectral
bands produced assist in quantitative analysis as well. Raman spectroscopy is even being used in areas outside of physical
science (i.e. archeology and art preservation) due to the characteristics mentioned above.

Introduction
Raman spectroscopy is based on scattering of radiation (Raman scattering), which is a phenomenon discovered in 1928 by
physicist Sir C. V. Raman. The field of Raman spectroscopy was greatly enhanced by the advent of laser technology during the
1960s.1 Resonance Raman also helped to advance the field. This technique is more selective compared to non-resonance
Raman spectroscopy. It works by exciting the analyte with incident radiation corresponding to the electronic absorption
bands.2 This causes an augmentation of the emission up to a factor of 106 in comparison to non-resonance Raman.2,3
In this section readers will be introduced to the theory behind resonance and non-resonance Raman spectroscopy. Each
technique has its share of advantages and challenges. Each of these aspects will be explored.

Theory
Raman scattering is the basis of the two Raman techniques. A molecule must have polarizability to Raman scatter and its
symmetry must be even (or gerade) for it to have polarizability2. Furthermore, the more electrons a molecule has gererally
increases its polarizability. Polarizability (α ) is a measure of an applied electronic field’s (E) ability to generate a dipole
moment (µ) in the molecule.5 In other words, it is an alteration of a molecule's electron cloud. Mathematically, this can be
determined by the following equation:
μ = αE (1)

To help provide a better visualization of how Raman spectroscopy works, a generic diagram can be seen in Figure 2. A sample
is irradiated with monochromatic laser light; which is then scattered by the sample. The scattered light passes through a filter
to remove any stray light that may have also been scattered by the sample.2 The filtered light is then dispersed by the
diffraction grating and collected on the detector. This set-up works for both the non-resonance and resonance Raman
techniques.

Figure 1: Simplified diagram of how a Raman spectrometer works. Images used with permission from
http://www.doitpoms.ac.uk
Non-resonance Raman scattering occurs when the radiation interacts with a molecule resulting in polarization of the
molecule’s electrons.4 The increase in energy from the radiation excites the electrons to an unstable virtual state; therefore, the
interaction is almost immediately discontinued and the radiation is emitted (scattered) at a slightly different energy than the
incident radiation.4
Resonance Raman scattering occurs in a similar fashion. However, the incident radiation is at a frequency near the frequency
of an electronic transition of the molecule of interest. This provides enough energy to excite the electrons to a higher electronic
state. Figure 1 provides a visual depiction of what non-resonance and resonance Raman scattering looks like in terms of
energy levels.

6/6/2021 1 https://chem.libretexts.org/@go/page/1851
Figure 2: Diagram depicting the different processes for non-resonance and resonance Raman spectroscopy. Notice that
fluorescence is more likely to be a problem with resonance Raman than with non-resonance. Image used with permisison from
Wikipedia (Credit: Moxfyre).

Advantages of Non-Resonance and Resonance Raman


Instrumental techniques each have certain strengths that make them better suited for some jobs as oppose to others. Non-
resonance is a good example of this notion. It is considered better suited for analyzing water containing samples due to water’s
low polarizability. Non-resonance and resonance Raman each have the capability to analyze samples in the gaseous, liquid, or
solid state. Their non-destructive nature makes it a great candidate for doing analysis of delicate materials. Archeologists and
art historians even find resonance Raman spectroscopy useful for studying and authentication of artifacts and artwork.4
Monochromatic light in the ultraviolet or near-infrared regions is generally used for both resonance and non-resonance Raman
spectroscopy. A tunable laser is preferred for resonance Raman and can be an advantage. That is because only one laser is
necessary to do analyses of multiples samples in which each one requires a different excitation wavelength.4 This allows the
user to switch out samples without having to switch out the lasers as well. It becomes a matter of just changing the setting on
the tunable laser. If the laboratory is not equipped with a tunable laser, any laser that is available can be used to achieve the
enhancement of the Raman signal. The only stipulation being that the laser available must have a frequency as near as possible
to one of the analyte’s electronic transitions.2 Therefore, researchers conducting resonance Raman spectroscopy without a
tunable laser are at the mercy of whatever laser they do have in the laboratory.
Resonance Raman spectroscopy has greater sensitivity compared to its non-resonance counterpart. It is capable of analyzing
samples with concentrations as low as 10-8 M. Non-resonance Raman can analyze samples with concentrations no lower than
0.1 M. Resonance Raman spectroscopy produces a spectrum with relatively few lines. The reason being that the technique
only augments Raman signals affiliated with chromophores in the analyte.2,4 This makes the technique particularly useful for
analysis of larger molecules like biomolecules.

Fluorescence Disadvantage
Fluorescence is a problem for both Resonance Raman techniques, particularly when using sources in the visible range.2 Non-
resonance Raman signals are generally weak and can be easily overwhelmed by fluorescence signals.6 In addition,
fluorescence has a longer excited state lifetime compared to Raman scattering, causing an inability to detect Raman signals.2,6
Even when the analyte is not a fluorescent molecule, the signal could be a result of the sample matrix content (i.e. solvent or
contaminants). Resonance Raman is particularly at risk of inducing fluorescence because it uses sources at frequencies near to
that of a molecule’s electronic transition. The radiation is more likely to absorb resulting in fluorescence as a possible
mechanism for the electrons return to the ground state. Thus, highly fluorescent molecules should be avoided when using
Raman spectroscopy; especially resonance Raman. Figure 3 is a general illustration of how a fluorescence signal can
overwhelm Raman signals.

6/6/2021 2 https://chem.libretexts.org/@go/page/1851
Raman shift (cm-1)
Figure 3: Two generic Raman spectra overlaid. The blue Raman spectrum represents one obtained via excitation source in the
visible range. The black Raman spectrum represents one obtained via excitation source in the near-infrared range. The black
Raman signals are free of fluorescence interference.
There are techniques that spectroscopists use to avoid fluorescence interference. For instance, background subtraction could be
done. Another example is to use near-infrared radiation to excite the sample as a means to overcome fluorescence.6 A more
elaborate method was used by Matousek et al. They took advantage of the differences in excitation lifetimes for Raman and
fluorescence. It required implementing shifted excitation Raman difference spectroscopy (SERDS) in conjunction with a
device known as a Kerr gate to successfully obtain a resonance Raman spectrum of the rhodamine 6G dye.6 SERDS is a
technique that uses two excitation wavelengths to produce two Raman spectra. The excitation wavelengths have a difference in
value that corresponds to the bandwidth of the Raman signal.6 The two spectra are subtracted from each other and the
difference spectra is recreated by means of mathematical processes.6
A Kerr gate can be used to remove fluorescence from a Raman signal based on their different lifetimes.6 The device consists
of a couple of crossed polarizers, Kerr medium, and an additional laser to provide a gating pulse.7 Now consider a sample that
has been irradiated resulting in fluorescence and Raman scattering. The fluorescence and Raman scatter would pass through
the crossed polarizer and then through the Kerr medium (Matousek et al. used carbon disulfide as the Kerr medium). The Kerr
gate is referred to as being open when a laser pulse (the gating pulse) strikes the Kerr medium as the fluorescence and Raman
scattered light pass through.7 Furthermore, the gate remains open for a length of time corresponding to the lifetime of the
Raman scatter,7 The interaction of the gating pulse with the Kerr medium causes the light to become anisotropic and transmit
beyond the Kerr medium.7 The light then goes from being polarized in a linear direction to elliptical polarization. However,
Raman scattered light can be selectively switched back to linear polarization by selecting the appropriate propagation length
for Kerr medium transmission; or by altering the degree of anisotropy.7 The Raman scattered light is then allowed to pass
through the second crossed polarizer and on to the spectrometer. Any fluorescence that passes through the Kerr medium is
prevented from entering the spectrometer due to its inability to transmit through the second crossed polarizer on account of its
new elliptical polarization. Figure 4 should assist with the visualization of such process.

Figure 4: This is an illustration of a Kerr gate. The thick black arrow represents the direction of the signal produced. The blue
and red arrows are fluorescence and Raman scattering, respectively. All other components are labeled in the figure.

Conclusions
This module was meant to provide an introduction to the similarities and differences between non-resonance and resonance
Raman spectroscopy. Notice that they each have their own advantages making both of them powerful analytical techniques.

6/6/2021 3 https://chem.libretexts.org/@go/page/1851
Non-resonance Raman is more advantageous when compared to IR spectroscopy. However, resonance Raman appears to have
the upper hand when compared to its non-resonance counterpart. The important thing is to choose the technique that is most
appropriate for the work to be done.

Problems
1. What does Dr. Nivens' quote mean and what can be done to avoid the problem?
2. Archeology and art preservation were mentioned as fields in which Raman spectroscopy is useful due to the following
characteristics: non-destructive to sample, qualitative analysis, and quantitative analysis. Name an additional area/field that
could benefit from Raman due to these characteristics and why.
3. Does non-resonance Raman or resonance Raman have a better limit of detection?
4. What is polarizability?
5. You are working on a project that is trying to determine if lycopene is absorbed better via supplements or diet. You only
have a Raman spectrometer available to conduct your analysis. Thus you decide to determine the amount absorbed in the
body by subtracting the amount excreted in urine from the total intake. In general, would it be better to detect lycopene in
your biological sample using non-resonance or resonance Raman spectroscopy and why?

References
1. Efremov, E.; Ariese, F.; Gooijer, C. “Achievements in resonance Raman spectroscopy: A Review of a technique with a
distinct analytical chemistry potential” Analytica Chimica Acta, 606, 2008, 119-134.
2. Smith, W.; Dent, G. Introduction, Basic Theory, and Principles and Resonance Raman Scattering. Modern Raman
Spectroscopy, John Wiley & Sons, Ltd, England, 2005; 1-7 & 93-97.
3. Schmitt, M.; Popp, J. “Raman spectroscopy at the beginning of the twenty-first century” J. Raman Spectrosc., 37, 2006,
20-28.
4. Krishnan, R. S.; Shankar, R. K. “Raman effect: History of the discovery” J. Raman Spectrosc., 10, 1981, 1-8.
5. Ball, D.W. “Theory of Raman Spectroscopy” Spectroscopy, 16 (11), 2001, 32.
6. Matousek, P.; Towrie, M.; Parker, A. W. “Fluorescence background suppression in Raman spectroscopy using combined
Kerr gated and shifted excitation Raman difference techniques” J. Raman Spectrosc., 33, 2002, 238-242.

Solutions
1. "Fluorescence is the enemy of Raman" refers to the fact that fluorescence induced during Raman spectroscopy will inhibit
detection of Raman signals. One way to avoid such a problem is to use an excitation source that is in the ultra-violet or
near-infrared range.
2. Answers will vary, but here is an example: Forensic science is a field that could benefit for the indicated characteristics of
Raman spectroscopy. The non-destructive nature will insure that precious evidence is not damage or destroyed during
analysis. The evidence can then be preserved for additional testing. The qualitative and quantitative factors could provide
information to investigators as to what is present in their evidence (i.e. bodily fluids, drugs, accelerants, etc.).
3. Resonance Raman is capable of analyte detection at concentrations as low as 10-8 M. The limit of detection is much higher
for non-resonance Raman.
4. Polarizability measures the capability of an applied electric field to cause a dipole moment.
5. The sample will most likely contain a multitude of things because it is biological. Resonance Raman would be the better
choice because lycopene contains chromophores which are targeted for excitation in that technique. Thus the spectra will
not be cluttered by all the other contents in the biological sample.

6/6/2021 4 https://chem.libretexts.org/@go/page/1851
CHAPTER OVERVIEW
ELECTRONIC SPECTROSCOPY
Electron spectroscopy is an analytical technique to study the electronic structure and its dynamics in
atoms and molecules. In general an excitation source such as x-rays, electrons or synchrotron
radiation will eject an electron from an inner-shell orbital of an atom.

CIRCULAR DICHROISM
Circular Dichroism, an absorption spectroscopy, uses circularly polarized light to investigate
structural aspects of optically active chiral media. It is mostly used to study biological molecules,
their structure, and interactions with metals and other molecules.

ELECTRONIC SPECTROSCOPY: APPLICATION


Electronic Absorption and Fluorescence spectroscopy are both analytical methods that center
around the idea that when one perturbs a known or unknown solution with a spectrum of energetic photons, those photons that have
the correct energy to interact with the molecules in solution will do so, and those molecules under observation will always interact
with photons of energies characteristic to that molecule.

ELECTRONIC SPECTROSCOPY - INTERPRETATION


Electronic Spectroscopy relies on the quantized nature of energy states. Given enough energy, an electron can be excited from its
initial ground state or initial excited state (hot band) and briefly exist in a higher energy excited state. Electronic transitions involve
exciting an electron from one principle quantum state to another. Without incentive, an electron will not transition to a higher level.
Only by absorbing energy, can an electron be excited.

ELECTRONIC SPECTROSCOPY BASICS


Explains the origin of UV-visible absorption spectra, how they are measured, and how they can be used in the analysis of organic
compounds.

A DOUBLE BEAM ABSORPTION SPECTROMETER


BONDING THEORY FOR UV-VISIBLE ABSORPTION SPECTRA
ELECTROMAGNETIC RADIATION
THE BEER-LAMBERT LAW
The Beer-Lambert law relates the attenuation of light to the properties of the material through which the light is traveling. This page
takes a brief look at the Beer-Lambert Law and explains the use of the terms absorbance and molar absorptivity relating to UV-visible
absorption spectrometry.

USING UV-VISIBLE ABSORPTION SPECTROSCOPY


WHAT CAUSES MOLECULES TO ABSORB UV AND VISIBLE LIGHT
This page explains what happens when organic compounds absorb UV or visible light, and why the wavelength of light absorbed
varies from compound to compound.

FLUORESCENCE AND PHOSPHORESCENCE


Fluorescence and phosphorescence are types of molecular luminescence methods. A molecule of analyte absorbs a photon and excites
a species. The emission spectrum can provide qualitative and quantitative analysis. The term fluorescence and phosphorescence are
usually referred as photoluminescence because both are alike in excitation brought by absorption of a photon.

JABLONSKI DIAGRAM
A Jablonski diagram is basically an energy diagram, arranged with energy on a vertical axis. The energy levels can be quantitatively
denoted, but most of these diagrams use energy levels schematically. The rest of the diagram is arranged into columns. Every column
usually represents a specific spin multiplicity for a particular species. However, some diagrams divide energy levels within the same
spin multiplicity into different columns.

JABLONSKI DIAGRAM: SHOCKWAVE


METAL TO LIGAND AND LIGAND TO METAL CHARGE TRANSFER BANDS
In the field of inorganic chemistry, color is commonly associated with d–d transitions. If this is the case, why is it that some transition
metal complexes show intense color in solution, but possess no d electrons? In transition metal complexes a change in electron
distribution between the metal and a ligand gives rise to charge transfer (CT) bands when performing Ultraviolet-visible spectroscopy
experiments. A brief introduction to electron transfer reactions and Marcus-Hush theory is given.

1 7/7/2021
RADIATIVE DECAY
Spontaneous emission is the process in which a quantum mechanical system (such as an atom, molecule or subatomic particle)
transitions from an excited energy state to a lower energy state (e.g., its ground state) and emits a quantum in the form of a photon.

CHEMILUMINESCENCE
Most chemiluminescence methods involve only a few chemical components to actually generate light.

FLUORESCENCE
Fluorescence, a type of luminescence, occurs in gas, liquid or solid chemical systems. Fluorescence is brought about by absorption of
photons in the singlet ground state promoted to a singlet excited state. The spin of the electron is still paired with the ground state
electron, unlike phosphorescence. As the excited molecule returns to ground state, it involves the emission of a photon of lower
energy, which corresponds to a longer wavelength, than the absorbed photon.

PHOSPHORESCENCE
Unlike florescence, phosphorescence does not re-emit the light immediately. Instead, phosphorescence releases light very slowly in
the dark due to its energy transition state. When light such as ultraviolet light is shined upon a glow in dark object, the object emits
light, creating phosphorescence.

SELECTION RULES FOR ELECTRONIC SPECTRA OF TRANSITION METAL COMPLEXES


The Selection Rules governing transitions between electronic energy levels.

DERIVATION OF LAPORTE RULE


The Laporte Rule is a selection rule in electron absorption spectroscopy that applies to centrosymmetric molecules. It says that
transitions between states of the same symmetry with respect to inversion are forbidden. Electronic transitions from waveunfunctions
with g symmetry to wavefunctions with g symmetry are forbidden, as are transitions from wavefunctions with u symmetry to
wavefunctions with u symmetry. Transitions from g to u and u to g, where the symmetry switches, may be allowed.

SPIN-ORBIT COUPLING
Spin-orbit coupling refers to the interaction of a particle's "spin" motion with its "orbital" motion.

ATOMIC TERM SYMBOLS


In electronic spectroscopy, an atomic term symbol specifies a certain electronic state of an atom (usually a multi-electron one), by
briefing the quantum numbers for the angular momenta of that atom. The form of an atomic term symbol implies Russell-Saunders
coupling. Transitions between two different atomic states may be represented using their term symbols, to which certain rules apply.

MOLECULAR TERM SYMBOLS


Molecular term symbols specify molecular electronic energy levels. Term symbols for diatomic molecules are based on irreducible
representations in linear symmetry groups, derived from spectroscopic notations. They usually consist of four parts: spin multiplicity,
azimuthal angular momentum, total angular momentum and symmetry. All molecular term symbols discussed here are based on
Russel-Saunders coupling.

THE RUSSELL SAUNDERS COUPLING SCHEME


The ways in which the angular momenta associated with the orbital and spin motions in many-electron-atoms can be combined
together are many and varied. In spite of this seeming complexity, the results are frequently readily determined for simple atom
systems and are used to characterize the electronic states of atoms.

TWO-PHOTON ABSORPTION
Two-photon absorption is one of a variety of two-photon processes. In this specific process, two photons are absorbed by a sample
simultaneously. Neither photon is at resonance with the available energy states of the system, however, the combined frequency of the
photons is at resonance with an energy state.

2 7/7/2021
Circular Dichroism
Circular Dichroism, an absorption spectroscopy, uses circularly polarized light to investigate structural aspects of optically
active chiral media. It is mostly used to study biological molecules, their structure, and interactions with metals and other
molecules.

Introduction
Circular Dichroism (CD) is an absorption spectroscopy method based on the differential absorption of left and right circularly
polarized light. Optically active chiral molecules will preferentially absorb one direction of the circularly polarized light. The
difference in absorption of the left and right circularly polarized light can be measured and quantified. UV CD is used to
determine aspects of protein secondary structure. Vibrational CD, IR CD, is used to study the structure of small organic
molecules, proteins and DNA. UV/Vis CD investigates charge transfer transitions in metal-protein complexes.

Circular Polarization of Light


Electromagnetic radiation consists of oscillating electric and magnetic fields perpendicular to each other and the direction of
propagation. Most light sources emit waves where these fields oscillate in all directions perpendicular to the propagation
vector. Linear polarized light occurs when the electric field vector oscillates in only one plane. In circularly polarized light, the
electric field vector rotates around the propagation axis maintaining a constant magnitude. When looked at down the axis of
propagation the vector appears to trace a circle over the period of one wave frequency (one full rotation occurs in the distance
equal to the wavelength). In linear polarized light the direction of the vector stays constant and the magnitude oscillates. In
circularly polarized light the magnitude stays constant while the direction oscillates.

Figure 1: Diagram of linearly polarized and circularly polarized light


As the radiation propagates the electric field vector traces out a helix. The magnetic field vector is out of phase with the
electric field vector by a quarter turn. When traced together the vectors form a double helix.

Light can be circularly polarized in two directions: left and right. If the vector rotates counterclockwise when the observer
looks down the axis of propagation, the light is left circularly polarized (LCP). If it rotates clockwise, it is right circularly
polarized (RCP). If LCP and RCP of the same amplitude, they are superimposed on one another and the resulting wave will be
linearly polarized.

Figure 2: The superposition of LCP and RCP light of the same amplitude produces linearly polarized light

Interaction with Matter

6/24/2021 1 https://chem.libretexts.org/@go/page/1761
As with linear polarized light, circularly polarized light can be absorbed by a medium. An optically active chiral compound
will absorb the two directions of circularly polarized light by different amounts
ΔA = Al − Ar (1)

This can be extended to the Beer-Lambert Law. The molar absorpitivty of a medium will be different for LCP and RCP. The
Beer-Lambert Law can be rewritten as
A = (εl − εr )cl (2)

The difference in molar absorptivity is also known as the molar circular dichroism
Δε = εl − εr (3)

The molar circular dichroism is not only wavelength dependent but also depends on the absorbing molecules conformation,
which can make it a function of concentration, temperature, and chemical environment.
Any absorption of light results in a change in amplitude of the incident wave; absorption changes the intensity of the light and
intensity of the square of the amplitude. In a chiral medium the molar absorptivities of LCP and RCP light are different so they
will be absorbed by the medium in different amounts. This differential absorption results in the LCP and RCP having different
amplitudes which means the superimposed light is no longer linearly polarized. The resulting wave is elliptically polarized.

Figure 3: The diagrams of the superposition of LCP and RCP light when viewed down the axis of propagation. On the left the
two circular waves (red and green) have the same amplitude which produces linearly polarized light (blue). On the right the
LCP (red) has a larger amplitude than the RCP (green), the superposition of the two waves (blue) forms an ellipse.

Molar Ellipticity
The CD spectrum is often reported in degrees of ellipticity, θ , which is a measure of the ellipticity of the polarization given by:
El − Er
tanθ = (4)
El + Er

where E is the magnitude of the electric field vector.

Figure 4: Elliptically polarized light (purple) is the superposition of LCP (red) and RCP (blue) light. θ is the angle between the
magnitude of the electric field vector at its maximum and its minimum
The change in polarization is usually small and the signal is often measured in radians where θ =
2.303

4
(Al − Ar ) and is a
function of wavelength. θ can be converted to degrees by multiplying by 180

π
which gives θ = 32.98ΔA

The historical reported unit of CD experiments is molar ellipticity, [θ] , which removes the dependence on concentration and
path length
[θ] = 3298Δε (5)

where the 3298 converts from the units of molar absorptivity to the historical units of degrees⋅ cm2⋅dmol-1.

6/24/2021 2 https://chem.libretexts.org/@go/page/1761
Applications
Instrumentation
Most commercial CD instruments are based on the modulation techniques introduced by Grosjean and Legrand. Light is
linearly polarized and passed through a monochromator. The single wavelength light is then passed through a modulating
device, usually a photoelastic modulator (PEM), which transforms the linear light to circular polarized light. The incident light
on the sample switches between LCP and RCP light. As the incident light swtches direction of polarization the absorption
changes and the differention molar absorptivity can be calculated.

Figure 5: The instrumentation for a common CD spectrometer showing the polarization of light and the differential absorption
of LCP and RCP light.

Biological molecules
The most widely used application of CD spectroscopy is identifying structural aspects of proteins and DNA. The peptide
bonds in proteins are optically active and the ellipticity they exhibit changes based on the local conformation of the molecule.
Secondary structures of proteins can be analyzed using the far-UV (190-250 nm) region of light. The ordered α -helices, β-
sheets, β-turn, and random coil conformations all have characteristic spectra. These unique spectra form the basis for protein
secondary structure analysis. It should be noted that in CD only the relative fractions of residues in each conformation can be
determined but not specifically where each structural feature lies in the molecule. In reporting CD data for large biomolecules
it is necessary to convert the data into a normalized value that is independent of molecular length. To do this the molar
ellipticity is divided by the number of residues or monomer units in the molecule.
The real value in CD comes from the ability to show conformational changes in molecules. It can be used to determine how
similar a wild type protein is to mutant or show the extent of denaturation with a change in temperature or chemical
environment. It can also provide information about structural changes upon ligand binding. In order to interpret any of this
information the spectrum of the native conformation must be determined.
Some information about the tertiary structure of proteins can be determined using near-UV spectroscopy. Absorptions between
250-300 nm are due to the dipole orientation and surrounding environment of the aromatic amino acids, phenylalanine,
tyrosine, and tryptophan, and cysteine residues which can form disulfide bonds. Near-UV techniques can also be used to
provide structural information about the binding of prosthetic groups in proteins.
Metal containing proteins can be studied by visible CD spectroscopy. Visible CD light excites the d-d transitions of metals in
chiral environments. Free ions in solution will not absorb CD light so the pH dependence of the metal binding and the
stoichiometry can be determined.
Vibrational CD (VCD) spectroscopy uses IR light to determine 3D structures of short peptides, nucleic acids, and
carbohydrates. VCD has been used to show the shape and number of helices in A-, B-, and Z-DNA. VCD is still a relatively
new technique and has the potential to be a very powerful tool. Resolving the spectra requires extensive ab initio calculations,
as well as, high concentrations and must be performed in water, which may force the molecule into a nonnative conformation.

References
1. Woody, R. W. Circular-Dichroim. Methods in Enzymology 246, 34-71 (1995).
2. Johnson, W. C. Protein secondary structure and circular dichroism: A practical guide. Proteins: Structure, Function, and
Genetics 7, 205–214 (1990).

6/24/2021 3 https://chem.libretexts.org/@go/page/1761
3. Drake, A. F. Polarisation modulation-the measurement of linear and circular dichroism. Journal of Physics E: Scientific
Instruments 19, 170–181 (1986).
4. Neidig, M. L., Wecksler, A. T., Schenk, G., Holman, T. R. & Solomon, E. I. Kinetic and spectroscopic studies of N694C
lipoxygenase: a probe of the substrate activation mechanism of a nonheme ferric enzyme. Journal of the American
Chemical Society 129, 7531–7537 (2007).
5. Polavarapu, P. L. & Zhao, C. X. Vibrational circular dichroism: a new spectroscopic tool for biomolecular structural
determination. Fresenius Journal Anaytical Chemistry 366, 727–734 (2000).

Contributors and Attributions


Nick Hurlburt, UCDavis

6/24/2021 4 https://chem.libretexts.org/@go/page/1761
Electronic Spectroscopy: Application
Electronic Absorption and Fluorescence spectroscopy are both analytical methods that center around the idea that when one
perturbs a known or unknown solution with a spectrum of energetic photons, those photons that have the correct energy to
interact with the molecules in solution will do so, and those molecules under observation will always interact with photons of
energies characteristic to that molecule. These methods have tremendously helped scientists elucidate and characterize the
physical properties of a variety of molecules by giving a means to probe the fundamental electronic structure of those
molecules.

Introduction
The theory behind Electronic Absorption and Fluorescence was described in a previous text. If you are unfamiliar with
electronic spectroscopy, browsing the theory might help paint a better picture of what will be discussed in this module.But
even if you understand what a Jablonski diagram represents or the mathematical description behind the transition dipole
moment operator, eventually you want to know how to apply the information you've learned to your world.
What this means to you is that you now have the ability that allow you to find the reason why carrots are orange or why plants
are green. You could take leaves and carrots, chop them into fine pieces, extract and separate the molecules inside of them, and
use Electronic Absorption or Fluorescence spectroscopy to figure out that the orange color in carrots is due to a family of
molecules known as the carotenoids and that the green color in the leaves is due to a variety of different chlorophyll
molecules. So excited by this knowledge, you could take the carotenoids that you now know are responsible for the color
orange and dye cloth to make yourself a carrot costume!
But what if you wanted a different color? What if you wanted a lighter shade of orange? What if you wanted a very exact and
very specific color? Before you can go explore the world (or lab) to find (or design) a molecule that has those properties you
need to be able to understand the instruments that allow you to do the work so you can be certain you get the results you
desire.
The remainder of this wikitext describes the design and components of a photospectrometer to measure the the Ultra Violet
through Visable (UV-Vis) region of the electromagnetic spectrum. Basic instrumental design and theory behind a
spectrofluorometer will also be discussed along with the errors and limits known to plague these instruments. By the end of
this text, you should be fully able to design a a photospectrometer in order to characterize the electronic behavior of molecules
in order to prove to yourself that carotenoids are responsible for the color orange (building of a carrot costume is optional)

Basic Design
As stated in the introduction, the purpose of a spectrophotometer is to characterize how molecules react with photons of
varying wavelengths. This easiest way to do this is to design an instrument that can control what type of light is in contact with
the sample and the amount of that light that makes it through. Shown below is a simple spectrophotometer that does just what
was described. A light source produces light that can be separated and controlled by the monochrometer which then allows
only the desired light to impinge upon the sample. The light that travels through the sample is detected and sent to a recorder.
Keep this overall picture in mind when thinking about the components and what each of their parts are in whole experiment.

Sources
As shown above, the first piece of equipment needed is a source of photons (light). Since we are interested in the UV-Vis
region of the electromagnetic spectrum, we need to assure that this source is able to produce photons with the correct
wavelengths (180-780 nm) that are characteristic to the spectrums of interest so that the experiment can be completed. Typical
sources come in two forms, either in the form of a lamp, or in the form of a laser. Lamps can provide a wide spectrum of light

5/23/2021 Electronic Spectroscopy.1 https://chem.libretexts.org/@go/page/1763


that can be used for almost any purpose at a very reasonable cost, while lasers provide very intense monochromatic light and
tend can be used for fluorescence and other specialized experiments due to their high costs.

Tungsten Filament Lamps


A common lamp source that might be familiar is the tungsten filament lamp or sometimes referred to as the incandescent light
bulb. This type of lamp produces a spectrum of light perfect for the visible spectrum (320 nm - 3.5 µm)1 by applying a voltage
through a thin filament of tungsten until it heats up enough to produce black body radiation. The Tungsten-Halogen lamp is a
significant improvement over the normal tungsten filament lamp due to its longer lifetime and the wider range of light it
produces. This is achieved by incasing the filament with quartz instead of silica so the filament can be forced to higher
temperatures (thus increasing the range of light produced) and by trapping a halogen gas (normally Iodine) that can react with
sublimated tungsten atoms to form WI2 that can then redeposit the tungsten by coming into contact with the filament (thus
increasing the lifespan).

Hydrogen/Deuterium Lamps
This type of lamp is used to obtain a spectrum of light in the Ultra Violet region (190-400 nm). The spectrum is obtained by
trapping Hydrogen at low pressures inside of a glass tube and creating an electrical arc inside of the cell. This electrical arc
creates high energy hydrogen gas atoms that dissociate into two hydrogen atoms along with producing photons. The
wavelength of the photons created from the dissociation is dependent on the resulting kinetic energies of the two hydrogen
atoms after the dissociation. Since the kinetic energies of the dissociated atoms can vary from zero to the original energy of the
excited atom, the photon produced can also vary which leads to a spectrum of photons. Deuterium is often used in lieu of
hydrogen because it can produce more a more intense spectrum and has a longer lifetime. Both types of this lamp are housed
within a quartz shell since silica absorbs in the Ultraviolet range.

Lasers
Tunable dye and fixed wavelength lasers that produce photons within the UV-Vis spectrum are commonly used for
fluorescence experiments where the intensity of the resulting fluorescence is directly proportional to the intensity of the
source. Lasers are inherently focused which provides the benefit of being able to work with very small or dilute samples. The
caveats of using a laser are that the spectrum produced is physically limited and that care must also be taken to avoid
photodegredation of your samples by attenuating the power of the laser. A description of the construction and theory behind
lasers won't be discussed here, but can be found through a variety of sources.1,2,3,4,5,6

Filters and Monochrometers


Unless a laser is used, a device is needed to restrict and isolate the wavelengths that our sources provide in order to control the
experiment. If a isolated wavelength is needed or isn't needed, interference or absorption filters can be used. If a spectrum of
wavelengths needs to be passed through the sample, monochrometers provide the ability to separate and control the light
passed through the sample while preserving the available spectrum.

Filters
Intereference filters are designed to provide constructive or destructive interference of light by taking advantage of the
refraction of light through different materials. As light passes from one medium to the other the direction and wavelength of
light can be changed based on the index of refraction of both mediums involved and the angle of the incident and exiting light
(For more look at Snell's Law). Due to this behavior, constructive and destructive interference can be controlled by varying the
thickness (d) of a transparent dielectric material between two semi-reflective sheets and the angle the light is shined upon the
surface. As light hits the first semi-reflective sheet, a portion is reflected, while the rest travels through the dielectric to be bent
and reflected by the second semi-reflective sheet. If the conditions are correct, the reflected light and the initial incident light
will be in phase and constructive interference occurs for only a particular wavelength.

5/23/2021 Electronic Spectroscopy.2 https://chem.libretexts.org/@go/page/1763


Another common filter is the Absorption Filter. Absorption filters work on the premise of being able to filter light by
absorbing all other wavelengths that aren't of interest. They are normally constructed from a colored glass that absorbs over a
wide range. Although normally cheaper than interference filters, absorbance filters tend to be less precise at filtering for a
selected wavelength and also have the added penalty of absorbing some of the selected light thus lowering the intensity.

Monochrometers
Monochrometers, as previously mentioned, are used to control and separate light so that a sample can be subjected to a span of
wavelengths. Light entering a monochrometer is filtered by a thin slit. The filtered light is then focused by a mirror onto a
dispersing element that separates the light into its different wavelengths. The separated light is then focused again and angled
toward an exit slit which then filters against all wavelengths except the desired one. The wavelength selected can then easily
be changed by rotating the dispersing element to support the transmission of the new desired wavelength. Though
monochrometers nearly all have the same practical design, the difference normally is determined by the type of dispersing
element.
One type of dispersing element is a prism which disperses light by refraction through two angled surfaces. In order to separate
the light into different wavelengths, the prism needs to be made of material that has a change in the index of refraction with
respect to wavelength so that each wavelength is bent at a different angle. The larger the difference in the index of refraction
the better the separation is between wavelengths. For the UV-Vis range, a typical prism is cut from left handed quartz at a
thirty degree angle and attached to another piece of quarts cut the same way except from right handed quartz to make a Conru
Prism. The pitfall of prism based dispersing elements is the fact that the index of refraction for most materials varies
nonlinearly when compared to wavelength which results in a smaller degree of separation at longer wavelengths.

More commonly used due to less expensive fabrication costs and its ability to separate light in a linear fashion are grating
dispersing elements. Gratings are designed to diffract light which is a separation based on the angle of the incident light to the
grating normal and the spacing between groves. When beams of light impinge upon the grating some beams travel farther than
others and this causes an effect akin to a interference filter which allows for constructive and destructive interference and
provides the means to reflect a specific wavelength based on the angle of incidence. The most common type of grating is the

5/23/2021 Electronic Spectroscopy.3 https://chem.libretexts.org/@go/page/1763


echellette grating in which the grooves are angled in order to provide maximum reflection of the incident light into a single
order of reflected light.

Sample Cell
Care must be taken when considering a sample container to avoid unwanted absorption in the range of interest. While glass
might be perfect for the visible range, it absorbs in the UV range. Quartz can be used for both but most likely would cost
considerably more. Disposable Plastic cuvets created from polystyrene or polymethyl methacrylate are in common use today
as they are cheap to purchase and eliminate the need for cleaning cuvettes in order to analyze multiple samples. The shape of
the sample holder is also very important as unwanted scattering of light should be minimized and pathlength through the cell
should remain constant . Square (1 cm) cuvettes with frosted sides have been designed to minimize scattering and provide a
surface to hold the cuvette, thereby reducing smudges or smears of the optical window all the while keeping a fixed path
length. Cuvettes for fluorescent experiments cannot allow the frosted sides due to the 90° angle of most instruments and have
to be carefully cleaned as bodily oils have the ability to fluoresce. Placement of the sample within either a absorption or
fluorescence device is imperative and is normally fixed by a sample holder to insure reproducible results.

Detectors
After the light has passed through the sample, we want to be able to detect and measure the resulting light. These types of
detectors come in the form of transducers that are able to take energy from light and convert it into an electrical signal that can
be recorded, and if necessary, amplified.Photomultiplier tubes are a common example of a transducer that is used in a variety
of devices. The idea is that when a photon hits the top of the tube electrons are released, which are pulled toward the other end
of the tube by an electric field. The way the electrons are multiplied is due to the fact that along the length of the tube there are
several dynodes that have a slightly less negative potential than surface before it which causes the electrons traveling down the
tube to hit each surface which then in turn produces more electrons until at the very end there is a large amount of electrons
(~1,000,000) representing the one photon that started the cascade.

Another type of transducer is a Charge injection device (CID). This device works by having a p-type and n-type
semiconductor next to each other in which the n-type material is separated from the anodes by a silica layer that acts as a
capacitor. As a photon interacts with the n-type layer, an electron migrates to the p-type semiconductor and a positive charge is
generated that migrates toward the silica capacitor. This process continues to happen until the potential is measured by means
of comparing the more negative of the anodes to ground (Vi) . The charge is then transferred to the other anode and measured

5/23/2021 Electronic Spectroscopy.4 https://chem.libretexts.org/@go/page/1763


(Vf). The difference between Vf and Vi is directly proportional to the amount of photons that collided with the n-type layer.
The positive charges are then repelled when the anodes switch momentarily to having a positive charge and then the cycle can
be repeated. Thousands of these little detectors can be aligned and used to describe the light that interacts with it and can rival
the performance of the photomultiplier tube.

Signal Processor
In the end of the experiment all of the data collected needs to be stored or written down. In the past printers have been used to
record the transmittance as the experiment progresses. These days, computers are used to record, store, and even manipulate
data along with controlling the devices within the spectrometer. This ability allowed by having a computer that can quickly
integrate, compare, or annotate spectra dramatically decreases the learning curve needed to operate the instrument as well as
reduces the labor associated with simple or even complex experiments. Instead of having to record the transmittance, convert it
to absorbance and then subtract the background noise, the computer can accomplish all that for you in the matter of a few
seconds .

Instrumental Designs
Now that we've covered the instrument and its components in some detail, there are some different modifications that can be
made in order to tailor the device to the experiment it'll be performing. The device introduced in the beginning of this text is
the basic single beam spectrometer. This device can measure the transmittance or absorbance of a particular analyte at a given
wavelength provided by the source and monochrometer. Background subtraction in these machines needs to be done
separately before the analyte is inserted and can be stored and subtracted by the signal processor attached. Although this is the
simplest design, the cost of these types of instruments can vary greatly depending on the components that the machine is
comprised of.

For fluorescence spectroscopy the equivalent to a single beam spectrometer has a few slight modifications in that the detector
is perpendicular to the source, and that there is an additional monochrometer that can be used to vary the wavelength detected.

5/23/2021 Electronic Spectroscopy.5 https://chem.libretexts.org/@go/page/1763


In this fashion, the source light is unable to interfere with the fluorescence light being measured, and the fluorescent light
produced can be separated and described as the resulting wavelengths of fluorescence as a function of incident light.

A double beam spectrometer takes a more analytical approach to the design. Because of fluctuations in the source intensity and
inconsistencies in the transducer, the source light is split into two beams, one which travels through the sample, and another
that is sent through a blank or standard solution. Both beams are then read by separate transducers and the difference between
the two is recorded as the corrected transmittance. This allows for quick screening of analytes and negates the need for two
separate scans to complete a background subtraction. The same idea can be applied to a fluorometer as well to obtain the same
benefit

The latest instrument designs use a multichannel detector, such as an array of CID's, that allow for the spectrum of an analyte
to be gathered in seconds due to the fact that the light transmitted through the sample can be split and a spectrum of
wavelengths can be monitored simultaniously instead of individually. Also fiber optic cables can be used to transfer the light
from the source to the sample or from the sample to the dispersion grating and negate the need to consider the slit when
thinking about resolution.

Instrumental Noise
As with any instrument, the measurements we make are limited by the tools we use. In spectrophotometers, the only
measurement they are designed to monitor is the transmittance of light through the sample. Thus any noise associated with the
transmittance will correlate with errors in our analyses. The analysis of the noise in spectrophotometers has been done7 and is
outlined by Skoog and others in "Principles of Analytical Chemistry"3. In the outline they summarize that the types of errors
commonly experienced fall into three cases and are listed below.
In the first case, the standard deviation of the transmittance is a constant.
ST = k1 (Electronic Spectroscopy.1)

Errors associated with this situation are due to detectors with limited readout resolution and dark currant and amplifier noise.
Limited readout resolution keeps the deviation constant because the measurements more precise than the readout can not be

5/23/2021 Electronic Spectroscopy.6 https://chem.libretexts.org/@go/page/1763


expressed or displayed by the machine. Dark Currant and Amplifier noise are only an issue when the lamp intensity and
transducer sensitivity are low and random fluctuations in the currant become the dominant source of error.
In the second case, the deviation in transmittance varies with the equation.
− −− −−−
2
ST = K2 √ T +T (Electronic Spectroscopy.2)

Errors associated with this equation are those related to shot noise or the transfer of electrons across any sort of junction or
barrier like those found in the photomultiplier tube in the detector. Since the currant is dependant on this transfers, which
happen randomly, the currant becomes a random distribution that centers around and average. This error becomes considerable
when dealing with either really low or really high transmittances.
The last situation deals with errors that relate proportionally to the transmittance. Errors related to this type of noise have to do
with source flickering along with cell positioning. Both of these errors can be easily corrected either by attaching a constant
voltage source on the lamp or by placing a fixed cuvette holde
ST = k3 T (Electronic Spectroscopy.3)

Other big sources of error can be introduced from the entrance and exit slits not having the appropriate separation or if the
experiment being conducted is outside or near the limit of the instrumentation. In older machines, a scan time that was faster
than the recording printer could be a source of error as the printer would not be able to record the data as fast it was being
given.

References
1. G. W Ewing. "Instrumental Methods of Chemical Analysis", 5th Edition, McGraw-Hill 1985
2. H. H. Willard, Et Al. "Instrumental Methods of Analysis", 7th Edition, Wadsworth Inc. 1988
3. D. A. Skoog, Et Al. "Principles of Instrumental Analysis" 6th Edition, Thomson Brooks/Cole. 2007
4. R. A. Serway, J. W. Jewett, "Physics: For Scientists and Engineers with modern physics", Thomson Brooks/Cole, 2004
5. G. W. Ewing, J. Cazes, "Ewing's Analytical Instrumentation Handbook", 3rd Edition, Marcel Dekker, 2005
6. E. D. Stokes, Et Al. Optics Communications, Vol. 5, Num. 4, 267-270, 1972
7. L.D. Rothman, S. R. Crouch, J. D. Ingle. Analytical Chemistry, Vol 47, Num. 8, 1226-1233, 1975

5/23/2021 Electronic Spectroscopy.7 https://chem.libretexts.org/@go/page/1763


Electronic Spectroscopy - Interpretation
Electronic Spectroscopy relies on the quantized nature of energy states. Given enough energy, an electron can be excited from
its initial ground state or initial excited state (hot band) and briefly exist in a higher energy excited state. Electronic transitions
involve exciting an electron from one principle quantum state to another. Without incentive, an electron will not transition to a
higher level. Only by absorbing energy, can an electron be excited. Once it is in the excited state, it will relax back to it's
original more energetically stable state, and in the process, release energy as photons.

Introduction
Often, during electronic spectroscopy, the electron is excited first from an initial low energy state to a higher state by
absorbing photon energy from the spectrophotometer. If the wavelength of the incident beam has enough energy to promote an
electron to a higher level, then we can detect this in the absorbance spectrum. Once in the excited state, the electron has higher
potential energy and will relax back to a lower state by emitting photon energy. This is called fluorescence and can be detected
in the spectrum as well.

Embedded into the electronic states (n=1,2,3...) are vibrational levels (v=1,2,3...) and within these are rotational energy levels
(j=1,2,3...). Often, during electronic transitions, the initial state may have the electron in a level that is excited for both
vibration and rotation. In other words, n=0, v does not = 0 and r does not =0. This can be true for the ground state and the
excited state. In addition, due to the Frank Condon Factor, which describes the overlap between vibrational states of two
electronic states, there may be visible vibrational bands within the absorption bands. Therefore, vibrational fine structure that
can be seen in the absorption spectrum gives some indication of the degree of Frank Condon overlap between electronic states.
When interpreting the absorbance and fluorescence spectra of a given molecule, compound, material, or an elemental material,
understanding the possible electronic transitions is crucial. Assigning the peaks in the absorption spectrum can become easier
when considering which transitions are allowed by symmetry, the Laporte Rules, electron spin, or vibronic coupling. Knowing
the degree of allowedness, one can estimate the intensity of the transition, and the extinction coefficient associated with that
transition. These guidelines are a few examples of the selection rules employed for interpreting the origin of spectral bands.
Only a complete model of molecular energy diagrams for the species under investigation can make clear the possible
electronic transitions. Every different compound will have unique energy spacing between electronic levels, and depending on
the type of compound, one can categorize these spacings and find some commonality. For example, aromatic compounds pi to
pi* and n to pi* transitions where as inorganic compounds can have similar transitions with Metal to Ligand Charge Transfer
(MLCT) and Ligand to Metal Charge Transfer (LMCT) in addition to d-d transitions, which lead to the bright colors of
transition metal complexes. Although surprises in science often lead to discovery, it is more fortuitous for the interpreter to
predict the spectra rather than being baffled by the observation.
The following section will discuss the interpretation of electronic absorption spectra given the nature of the chemical species
being studied. This includes an understanding of the molecular or elemental electronic state symmetries, Russell-Sanders
states, spin multiplicities, and forbidden and allowed transitions of a given species.
As the light passes through the monochrometer of the spectrophotometer, it hits the sample with some wavelength and
corresponding energy. The ratio of the initial intensity of this light and the final intensity after passing through the sample is
measured and recorded as absorbance (Abs). When absorbance is measured at different wavelengths, an absorbance spectrum
of Abs vs wavelength can be obtained. This spectra reveals the wavelengths of light that are absorbed by the chemical specie,
and is specific for each different chemical. Many electronic transitions can be visible in the spectrum if the energy of the

6/18/2021 1 https://chem.libretexts.org/@go/page/1764
incident light matches or surpasses the quantum of energy separating the ground
state and that particular excited state. An example of an absorbance spectrum is
given below.

Figure 3.

Temperature Effects
Here we can see the effect of temperature and also the effect of solvents on the clarity of the spectrum. We can see from these
anonymous compounds that decreasing the temperature allows the vibrational fine structure to emerge. These vibrational
bands embedded within the electronic bands represent the transitions from v=n to v'=n. Generally, the v=0 to v'=0 transition is
the one with the lowest frequency. From there, increasing energy, the transitions can be from v=0 to v'=n, where n=1,2,3...
With a higher temperature, the vibrational transitions become averaged in the spectrum due to the presence of vibrational hot
bands and Fermi Resonance, and with this, the vibrational fine structure is lost at higher temperatures.

Figure 2.

Solvent Effects
The effect that the solvent plays on the absorption spectrum is also very important. It is clear that polar solvents give rise to
broad bands, non-polar solvents show more resolution, though, completely removing the solvent gives the best resolution. This
is due to solvent-solute interaction. The solvent can interact with the solute in its ground state or excited state through
intermolecular bonding. For example, a polar solvent like water has the ability of hydrogen bonding with the solute if the
solute has a hydrogen bonding component, or simply through induced dipole-dipole interactions. The non-polar solvents can
interact though polarizability via London interactions also causing a blurring of the vibronic manifold. This is due to the
solvent's tendency to align its dipole moment with the dipole moment of the solute. Depending on the interaction, this can
cause the ground state and the excited state of the solute to increase or decrease, thus changing the frequency of the absorbed
photon. Due to this, there are many different transition energies that become average together in the spectra. This causes peak-
broadening. The effects of peak broadening are most severe for polar solvent, less so for non-polar solvents, and absent when
the solute is in vapor phase.

6/18/2021 2 https://chem.libretexts.org/@go/page/1764
Group Theory and The Transition Moment Integral
When estimating the intensities of the absorption peaks, we use the molar absorptivity constant (epsilon). If the transition is
"allowed" then the molar absorptivity constant from the Beer's Law Plot will be high. This means that the probability of
transition is large. If the transition is not allowed, then there will be no intensity and no peak on the spectrum. Transitions can
be "partially allowed" as well, and these bands appear with a lower intensity than the full allowed transitions. One way to
decide whether a transition will be allowed or not is to use symmetry arguments with Group Theory.
If the symmetries of the ground and final state of a transition are correct, then the transition is symmetry allowed. We express
this by modifying the transition moment integral from an integral of eigenstates to an orthogonally expressed direct product of
the symmetries of the states.

∫ ψ2 μψ1 dT ⟶ Γ2 ⊗ Γμxyz ⊗ Γ1 (1)

∫ ψv ψel μψv ψel dT ⟶ Γv ⊗ Γel ⊗ Γμxyz ⊗ Γv ⊗ Γel (2)


2 2 1 1 2 2 1 1

The conversions of integration to direct products of symmetry as shown gives spectroscopists a short cut into deciding whether
the transition will be allowed or forbidden. A transition will be forbidden if the direct products of the symmetries of the
electronic states with the coupling operator is odd. More specifically, if the direct product does not contain the totally
symmetric representation, then the transition is forbidden by symmetry arguments. If the product does contain the totally
symmetric representation (A, A1, A1g...etc) then the transition is symmetry allowed.
Some transitions are forbidden by the Equation 1 and one would not expect to be able to see the band that corresponds to the
transition; however, a weak absorbance band is quite clear on the spectrum of many compounds. The transition may be
forbidden via pure electronic symmetries; however, for an octahedral complex for example since it has a center of inversion,
the transition is weakly allowed because of vibronic coupling. When the octahedra of a transition metal complex is completely
symmetric (without vibrations), the transition cannot occur. However, when vibrations exist, they temporarily perturb the
symmetry of the complex and allow the transition by equation (2). If the product of all of these representations contains the
totally symmetric representation, then the transition will be allowed via vibronic coupling even if it forbidden electronically.

Hot Bands
Some transitions are forbidden by symmetry and do not appear in the absorption spectrum. If the symmetries are correct, then
another state besides the ground state can be used to make the otherwise forbidden transition possible. This is accomplished by
hot bands, meaning the electrons in the ground state are heated to a higher energy level that has a different symmetry. When
the transition moment integral is solved with the new hot ground state, then the direct product of the symmetries may contain
the totally symmetric representation. If we employ the old saying, "You can't get there from here!" then we would be referring
to the transition from the ground state to the excited state. However, if we thermally excite the molecules from out of the
ground state, then, "we can get there from here!"
Knowing whether a transition will be allowed by symmetry is an essential component to interpreting the spectrum. If the
transition is allowed, then it should be visible with a large extinction coefficient. If it is forbidden, then it should only appear as
a weak band if it is allowed by vibronic coupling. In addition to this, a transition can also be spin forbidden. The examples
below of excited state symmetries, give an indication of what spin forbidden means:

6/18/2021 3 https://chem.libretexts.org/@go/page/1764
Figure 1:
These states are derived from the electron configuration of benzene. Once we have the molecular orbital energy diagram for
benzene, we can assign symmetries to each orbital arrangement of the ground state. From here, we can excite an electron from
the Highest Occupied Molecular Orbital (HOMO) to the Lowest Unoccupied Molecular Orbital (LUMO). This is the lowest
energy transition. Other transitions include moving the electron above the LUMO to higher energy molecular orbitals. To solve
for the identity of the symmetry of the excited state, one can take the direct product of the HOMO symmetry and the excited
MO symmetry. This give a letter (A, B, E..) an the subscript (1u, 2u, 1g...). The superscript is the spin multiplicity, and from
single electron transitions, the spin multiplicity is 2S+1 = M, where S = 1 with two unpaired electrons having the same spin
and S=0 when the excited electron flips its spin so that the two electrons have opposite spin. This gives M=1 and M=3 for
benzene above. From the results above, we have three transitions that are spin allowed and three that are spin forbidden.
Once we take the direct product of the symmetries and the coupling operator for each of these states given above, we find that
only the A1g to E1u transition is allowed by symmetry. Therefore, we have information regarding spin and symmetry
allowedness and we have an idea of what the spectra will look like:

When interpreting the spectrum, it is clear that some transitions are more probable than others. According to the symmetry of
excited states, we can now order them from low energy to high energy based on the position of the peaks (E1u is the highest,
then B1u, and B2u is lowest). The A1g to E1u transition is fully allowed and therefore the most intense peak. The A1g to B1u
and A1gto B2u transitions are symmetry forbidden and thus have a lower probability which is evident from the lowered
intensity of their bands. The singlet A1g to triplet B1u transition is both symmetry forbidden and spin forbidden and therefore

6/18/2021 4 https://chem.libretexts.org/@go/page/1764
has the lowest intensity. This transition is forbidden by spin arguments; however, a phenomenon known as spin-orbit coupling
can allow this transition to be weakly allowed as well. If spin-orbit coupling exists, then the singlet state has the same total
angular momentum as the triplet state so the two states can interact. A small amount of singlet character in the triplet state
leads to a transition moment integral that is non-zero, so the transition is allowed.

Organic Molecule Spectra


From the example of benzene, we have investigated the characteristic pi to pi* transitions for aromatic compounds. Now we
can move to other organic molecules, which involves n to pi* as well as pi to pi*. Two examples are given below:

The highest energy transition for both of these molecules has an intensity around 10,000 cm-1 and the second band has an
intensity of approximately 100 cm-1. In the case of formaldehyde, the n → π transition is forbidden by symmetry where as

the pi to pi* is allowed. The opposite is true for As(Ph)3 and the difference in molar absorptivity is evidence of this.

n → π

transitions
These transitions involve moving an electron from a nonbonding electron pair to a antibonding π orbital. They tend to have

molar absorbtivities less than 2000 and undergo a blue shift with solvent interactions (a shift to higher energy and shorter
wavelengths). This is because the lone pair interacts with the solvent, especially a polar one, such that the solvent aligns itself
with the ground state. When the excited state emerges, the solvent molecules do not have time to rearrange in order to stabilize
the excited state. This causes a lowering of energy of the ground state and not the excited state. Because of this, the energy of
the transition increases, hence the "blue shift".

π → π

transitions
These transitions involve moving an electron from a bonding π orbital to an antibonding π orbital. They tend to have molar

absorptivities on the order of 10,000 and undergo a red shift with solvent interactions (a shift to lower energy and longer
wavelengths). This could either be due to a raising of the ground state energy or lowering of the excited state energy. If the
excited state is polar, then it will be solvent stabilized, thus lowering its energy and the energy of the transition.

Inorganic Molecule Spectra


Speaking of transition probabilities in organic molecules is a good introduction to interpreting the spectra of inorganic
molecules. Three types of transitions are important to consider are Metal to Ligand Charge Transfer (MLCT), Ligand to Metal
Charge Transfer (LMCT), and d-d transitions. To understand the differences of these transitions we must investigate where
these transitions originate. To do this, we must define the difference between pi accepting and pi donating ligands:

6/18/2021 5 https://chem.libretexts.org/@go/page/1764
d-d Transitions
From these two molecular orbital energy diagrams for transition metals, we see that the pi donor ligands lie lower in energy
than the pi acceptor ligands. According to the spectral chemical series, one can determine whether a ligand will behave as a pi
accepting or pi donating. When the ligand is more pi donating, its own orbitals are lower in energy than the t2g metal orbitals
forcing the frontier orbitals to involve an antibonding pi* (for t2g) and an antibonding sigma* (for eg). This is in contrast to
the pi accepting ligands which involve a bonding pi (t2g) and an antibonding sigma* (eg). Because of this, the d-d transition
(denoted above by delta) for the pi acceptor ligand complex is larger than the pi donor ligand. In the spectra, we would see the
d-d transitions of pi acceptor ligands to be of a higher frequency than the pi donor ligands. In general though, these transitions
appear as weakly intense on the spectrum because they are Laporte forbidden. Due to vibronic coupling; however, they are
weakly allowed and because of their relatively low energy of transition, they can emit visible light upon relaxation which is
why many transition metal complexes are brightly colored. The molar extinction coefficients for these transition hover around
100.

LMCT Transitions
At an even higher energy are the LMCT which involve pi donor ligands around the metal. These transitions arise because of
the low-lying energy of the ligand orbitals. Therefore, we can consider this as a transition from orbitals that are ligand in
character to orbitals that are more metal in character, hence the name, Ligand to Metal Charge Transfer. The electron travels
from a bonding pi or non-bonding pi orbital into a sigma* orbital. These transitions are very strong and appear very intensely
in the absorbance spectrum. The molar extinction coefficients for these transitions are around 104. Examples of pi donor
ligands are as follows: F-, Cl-, Br-, I-, H2O, OH-, RS-, S2-, NCS-, NCO-,...

MLCT Transitions
The somewhat less common MLCT has the same intensity and energy of the LMCT as they involve the transition of an
electron from the t2g (pi) and the eg (sigma*) to the t1u (pi*/sigma*). These transitions arise from pi acceptor ligands and
metals that are willing to donate electrons into the orbitals of Ligand character. This is the reason that they are less frequent
since metals commonly accept electrons rather than donate them. All the same, both types of Charge Transfer bands are more
intense than d-d bands since they are not Laporte Rule forbidden. Examples of pi accepting ligands are as follows: CO, NO,
CN-, N2, bipy, phen, RNC, C5H5-, C=C double bonds, C=C triple bonds,...
From this spectra of an octahedral Chromium complex, we see that the d-d transitions are far weaker than the LMCT. Since
Chlorine is a pi donor ligand in this example, we can label the CT band as LMCT since we know the electron is transitioning

6/18/2021 6 https://chem.libretexts.org/@go/page/1764
from a MO of ligand character to a MO of metal character. The Laporte forbidden
(symmetry forbidden) d-d transitions are shown as less intense since they are only
allowed via vibronic coupling.
In addition, the d-d transitions are lower in energy than the CT band because of
the smaller energy gap between the t2g and eg in octahedral complexes (or eg to
t2g in tetrahedral complexes) than the energy gap between the ground and excited
states of the charge transfer band.
These transitions abide by the same selection rules that organic molecules follow:
spin selection and symmetry arguments. The Tanabe and Sugano diagrams for
transition metal complexes can be a guide for determining which transitions are
seen in the spectrum. We will use the [CrCl(NH3)5]2+ ion as an example for determining the types of transitions that are spin
allowed. To do this we look up the Tanabe and Sugano diagrams for Octahedral fields. Since Cr in the complex has three
electrons, it is a d3 and so we find the diagram that corresponds to d3 metals:
Based on the TS diagram on the left, and the information we have already learned, can
you predict which transition will be spin allowed and which ones will be forbidden? From
the diagram we see that the ground state is a 4A2. This is because of the three unpaired
electrons which make M=2S+1= 4. The A comes from the fact that there is only one
combination of electrons possible. With a spin multiplicity of 4, by the spin selection
rules, we can only expect intense transitions between the ground state 4A2 and 4T2, 4T1,
and the other 4T1 excited state. The other transitions are spin forbidden. Therefore, we
would expect to see three d-d transitions on the absorption spectra.
For us to visualize this, we can draw these transitions in order of increasing energy and
then plot the spectrum as we would expect it for only the d-d transitions in a d3 octahedral
complex:

From three spin allowed transitions, we would expect to see three d-d bands appear on the spectrum. In addition to these of
course, the LMCT band will appear as well.

Fluorescence
Now that we have discussed the nature of absorption involving an electron absorbing photon energy to be excited to a higher
energy level, now we can discuss what happens to that excited electron. Due to its higher potential energy, the electron will
relax back to its initial ground state, and in the process, emit electromagnetic radiation. The energy gap between the excited
state and the state to which the electron falls determines the wavelength of light that will be emitted. This process is called
fluorescence. Generally, the wavelengths of fluorescence are longer than absorbance, can you explain why? Given the
following diagram, one can see that vibrational relaxation occurs in the excited electronic state such that the electronic
relaxation occurs from the ground vibrational state of the excited electronic state. This causes lower energy electronic
relaxations than the previous energy of absorption.

6/18/2021 7 https://chem.libretexts.org/@go/page/1764
Here we see that the absorption transitions by default involve a greater energy change than the emission transitions. Due to
vibrational relaxation in the excited state, the electron tends to relax only from the v'=0 ground state vibrational level. This
gives emission transitions of lower energy and consequently, longer wavelength than absorption. When obtaining fluorescence,
we have to block out the transmitted light and only focus on the light being emitted from the sample, so the detector is usually
90 degrees from the incident light. Because of this emission spectra are generally obtained separately from the absorption
spectra; however, they can be plotted on the same graph as shown.

Generally separated by ~10 nm, the fluorescence peak follows the absorption peak according to the spectrum. With that, we
conclude our discussion of electronic spectroscopy interpretation. Refer to outside links and references for additional
information.

References
1. Cotton, Albert. Chemical Applications of Group Theory. John Wiley & Sons, New York, 1990.
2. Drago, Russell. Physical Methods for Chemists. Surfside Scientific Publishers, Gainesville, Fl, 1992.
3. Harris, Daniel; Bertolucci, Michael. Symmetry and Spectroscopy. An Introduction to Vibrational and Electronic
Spectroscopy. Dover Publications, Inc., New York, 1989.
4. Miessler, Gary; Tarr, Donald. Inorganic Chemistry. Pearson Education Inc., New Jersey, 2004.

Practice Problems!
1. From what we've discussed so far, if we change the solvent from non-polar to polar what effect will this have on the
frequency of absorption if the ground state is non-polar and the excited state is polar? Will it increase or decrease?
2. What causes peak broadening in absorption spectra?

6/18/2021 8 https://chem.libretexts.org/@go/page/1764
3. What are the little spikes in the more broad electronic transition bands? Draw potential energy wells to show their order
and use the Frank Condon factor to describe your answer.
4. Why are fluorescence bands lower in energy than absorption bands?
5. If an electronic transition is symmetry forbidden and spin forbidden, list two ways of overcoming this to explain why the
bands are still seen in the spectrum.
6. Define MLCT, LMCT, and d-d transitions and label the molar extinction coefficients associated with each.
7. How do the spectra of transition metal complexes differ with organic molecule?
8. What is a "blue shift" and a "red shift" and what solvent conditions would cause these to occur?
9. From the Tanabe Sugano diagram of a d2 metal complex, list all of the transitions that are spin allowed.
10. Define the coupling operator that sits between the excited state wave function and the ground state wave function in the
transition moment integral!

Contributors and Attributions


Troy Townsend

6/18/2021 9 https://chem.libretexts.org/@go/page/1764
Electronic Spectroscopy Basics
Explains the origin of UV-visible absorption spectra, how they are measured, and how they can be used in the analysis of
organic compounds.

Topic hierarchy

A Double Beam Absorption Spectrometer

Bonding Theory for UV-visible Absorption Spectra

Electromagnetic Radiation

The Beer-Lambert Law


The Beer-Lambert law relates the attenuation of light to the properties of the material through which the light is traveling.
This page takes a brief look at the Beer-Lambert Law and explains the use of the terms absorbance and molar absorptivity
relating to UV-visible absorption spectrometry.

Using UV-visible Absorption Spectroscopy

What Causes Molecules to Absorb UV and Visible Light


This page explains what happens when organic compounds absorb UV or visible light, and why the wavelength of light
absorbed varies from compound to compound.

Jim Clark 6/3/2021 1 https://chem.libretexts.org/@go/page/3742


A Double Beam Absorption Spectrometer
This page describes a double beam UV-visible absorption spectrometer.

Contributors and Attributions


Jim Clark (Chemguide.co.uk)

Jim Clark 6/18/2021 1 https://chem.libretexts.org/@go/page/3744


Bonding Theory for UV-visible Absorption Spectra
This page takes an introductory look at two areas of bonding theory needed for a proper understanding of how organic
compounds absorb some of the UV or visible light that passes through them.
It looks simply at anti-bonding orbitals, and what is meant by conjugation in compounds and how it contributes to the
delocalization of electrons.

Anti-bonding orbitals
Bonding and anti-bonding orbitals in a simple hydrogen molecule
I am assuming that you know how a simple covalent bond between two atoms forms. Half-filled atomic orbitals on each
atom overlap in space to form a new orbital (a molecular orbital) containing both electrons. In the case of two hydrogen
atoms, each has one electron in the 1s orbital. These come together to make a new orbital surrounding both of the
hydrogen nuclei.

It is important to understand exactly what this molecular orbital means. The two electrons are most likely to be found in this
region of space - and the most likely place to find them within this space is on the line between the two nuclei.
The molecule holds together because both nuclei are strongly attracted to this same pair of electrons. This most simple of
bonds is called a sigma bond - a sigma bond is one where the electron pair is most likely to be found on the line between
the two nuclei.
However . . . This is all a bit of a simplification! Molecular orbital theory demands that if you start with two atomic orbitals,
you must end up with two molecular orbitals - and we seem to be only producing one. A second molecular orbital is
formed, but in most cases (including the hydrogen molecule) it is left empty of electrons. It is described as an anti-bonding
orbital. The anti-bonding orbital has a quite different shape and energy from the bonding orbital.
The next diagram shows the relative shapes and energies of the various atomic and molecular orbitals when two hydrogen
atoms combine.

An anti-bonding orbital is always shown by the use of a star after its symbol.
Notice that when a bonding orbital forms, it is at a lower energy than the original atoms. Energy is released when the
bonding orbital is formed, and the hydrogen molecule is more energetically stable than the original atoms. However, an
anti-bonding orbital is less energetically stable than the original atoms.
A bonding orbital is stable because of the attractions between the nuclei and the electrons. In an anti-bonding orbital there
are no equivalent attractions - instead you get repulsions. There is very little chance of finding the electrons between the
two nuclei - and in fact half-way between the nuclei there is zero chance of finding them. There is nothing to stop the two
nuclei from repelling each other apart.

Jim Clark 5/28/2021 1 https://chem.libretexts.org/@go/page/3745


So in the hydrogen case, both of the electrons go into the bonding orbital, because that produces the greatest stability -
more stable than having separate atoms, and a lot more stable than having the electrons in the anti-bonding orbital.

Why doesn't helium form an He2 molecule?


You might reasonably say that helium can't form an He2 molecule because it doesn't have any unpaired electrons to share.
Fine! But let's also look at it from the point of molecular orbital theory. The diagram for helium is just a small modification of
the last one.

This time we have a total of 4 electrons in the original atomic orbitals. Two atomic orbitals have to form two molecular
orbitals. That means that this time, we would have to use both the bonding and anti-bonding molecular orbitals to
accommodate them.
But any gain in energetic stability due to the formation of the bonding orbitals would be countered by the loss of energetic
stability because of the anti-bonding ones. There is no energetic advantage in He2 forming - and so it doesn't.

Anti-bonding orbitals in double bonds


You are probably familiar with a picture of the double bond in ethene shown as:

The pi bond shown in red is, of course, a normal bonding orbital. It was formed by sideways overlap between a half-filled
p-orbital on each of the two carbon atoms. Remember that the two red shapes shown in the diagram are part of the same
pi bonding orbital. But if you overlap two atomic orbitals, you must get two molecular orbitals according to molecular orbital
theory. The second one is an anti-bonding pi orbital - and we never draw it under normal circumstances.
The anti-bonding pi orbital is (just like the anti-bonding sigma one) at a higher energy than the bonding orbital - and so isn't
used to hold electrons. Both of the electrons in the pi bond are found in the pi bonding orbital.

Summarizing the relative energies of various kinds of orbitals


The next diagram gives a general impression of how the energies of various types of orbital relate to each other in the sort
of compounds we will be looking at when we try to explain the absorption of light. It is not to scale. You will find that a new
sort of orbital has crept into the diagram - labelled "n" (for non-bonding). The sort of non-bonding orbitals that we will be
interested in contain lone pairs of electrons on, for example, oxygen, nitrogen and halogen atoms.
So . . . think of non-bonding orbitals as those containing lone pairs of electrons at the bonding level.

Jim Clark 5/28/2021 2 https://chem.libretexts.org/@go/page/3745


When light passes through a compound, some of the energy in the light kicks an electron from one of the bonding or non-
bonding orbitals into one of the anti-bonding ones. The energy gaps between these levels determine the frequency (or
wavelength) of the light absorbed, and those gaps will be different in different compounds. This is covered in detail on
another page.

Conjugation
We are going to leave explaining what conjugation is for a while - it is necessary to look at some more bonding first.

The simple ethene double bond


To understand about conjugated double bonds, you first need to be sure that you understand simple double bonds. Ethene
contains a simple double bond between two carbon atoms, but the two parts of this bond are different. Part of it is a simple
sigma bond formed from end-to-end overlap between orbitals on each carbon atom, and part is caused by sideways
overlap between a p-orbital on each carbon.
The important diagram is the one leading up to the formation of the pi bond - where the two p-orbitals are overlapping
sideways:

. . . giving the familiar pi bond.

Conjugated double bonds in buta-1,3-diene


Bonding in buta-1,3-diene
Buta-1,3-diene has the structure:

Now picture the formation of the various molecular orbitals as if you were thinking about two ethene molecules joined
together. You would have sigma bonds formed by the end-to-end overlap of various orbitals on the carbons and
hydrogens. That would leave you with a p-orbital on each carbon atom.

Those p-orbitals will overlap sideways - all of them! A system of delocalised pi bonds is formed, similar to the benzene
case that you are probably familiar with. The diagram shows one of those molecular orbitals.

To stress again - the diagram shows only one of the delocalised molecular orbitals. Remember that both of the red bits in
the diagram are part of the same orbital. The interaction of the two double bonds with each other to produce a delocalised
system of pi electrons over all four atoms is known as conjugation. Conjugation in this context literally means "joining
together".

Jim Clark 5/28/2021 3 https://chem.libretexts.org/@go/page/3745


In reality, if you start by overlapping four atomic orbitals, you will end up with four molecular orbitals. The four electrons will
go into the two lowest energy of these - two in each. That means that you get two pi bonding orbitals. We just draw one of
these for simplicity - the other one has a different shape.
There are also two pi anti-bonding orbitals, but these are normally empty. For most purposes, we ignore these entirely -
although not for this topic because energy from light can promote electrons from a pi bonding orbital into one of the anti-
bonding orbitals (as you will see on the next page).

Recognizing conjugated double bonds in a molecule


You can recognize the presence of conjugated double bonds in a molecule containing more than one double bond
because of the presence of alternating double and single bonds. The double bonds don't have to be always between
carbon atoms. All of the following molecules contain conjugated double bonds, although in the last case, the conjugation
doesn't extend over the whole molecule:

However, although the next molecule contains two double bonds, they aren't conjugated. They are separated by two single
bonds.

The reason why it is important to have the double and single bonds alternating is that this is the only way you can get all
the p-orbitals overlapping sideways. In the last case, you will get sideways overlap at each end of the molecule to get two
individual pi bonds. But the extra single bond in the middle stops them from interacting with each other.

delocalization extending beyond conjugated double bonds


Benzene rings
You are almost certainly familiar with the delocalization which occurs in a benzene ring. If you think about benzene using
the Kelulé structure, you have a perfect system of alternating single and double bonds around the molecule.

These conjugate to give the familiar delocalised pi system.

Once again, remember that this only shows one of the molecular orbitals formed. There will actually be three bonding pi
orbitals and three anti-bonding ones - because they arise from combining a total of six atomic orbitals. The extra bonding
orbitals aren't usually drawn.

Phenylamine and phenol


delocalization can also extend beyond pi bonds to include lone pairs on atoms like nitrogen or oxygen. Two simple
examples of this are phenylamine (aniline) and phenol. Writing these using the Kekulé structure for benzene:

Jim Clark 5/28/2021 4 https://chem.libretexts.org/@go/page/3745


You can see the alternating double and single bonds around the benzene ring. This conjugation leads to the familiar
delocalised electron system in benzene which we usually show as:

But the delocalization doesn't stop at the ring. It extends out to the nitrogen or oxygen atoms. In the phenylamine case,
there is a lone pair on the nitrogen atom which can overlap with the ring electrons . . .

. . . leading to delocalization which takes in both ring and nitrogen.

Exactly the same thing happens with phenol. One of the oxygen lone pairs overlaps with the ring electrons. The other one
is pointing in the wrong direction to get involved.

So . . . if you are trying to work out how far delocalization extends in a molecule, don't forget to look for atoms with lone
pairs that might get involved in the delocalization.

Other groups to look out for


Look especially for benzene rings with groups attached which contain double bonds. Let's start with a couple of simple
ones - phenylethene (styrene) and benzaldehyde.

In each case, you've got delocalization over the ring. Does it extend to the attached group? Do you have anything like
alternating single and double bonds? Yes, you do. You have the double bond in the side group, then a single bond, then

Jim Clark 5/28/2021 5 https://chem.libretexts.org/@go/page/3745


the ring delocalization. Looking at this in the phenylethene case, and imagining the arrangement of orbitals just before
delocalization over the side group:

You can see that the double bond and ring electrons will overlap to form a delocalised system looking something like this:

The benzaldehyde case is very similar, except that this time instead of the CH2 group at the end there is an oxygen with
two lone pairs. The delocalization is just the same.

Be careful, though! Remember that to get this extended delocalization, any double bond in the side chain must be able to
conjugate with the ring electrons - the two bits must be close enough to join together. Molecules like those in the next
diagram don't have the delocalization extending out into the side chain. The extra CH2 group prevents the necessary
sideways overlap between the p orbitals of the double bond and the ring electrons.

The other side group which it would be useful to know about is the nitro group, NO2 - for example in nitrobenzene.

The bonding in the nitro group is surprisingly awkward to work out. It is often shown with a double bond between the
nitrogen and one of the oxygens, and a co-ordinate (dative covalent) bond to the other.

This structure is actually misleading. Both nitrogen-oxygen bonds are identical and the group already has delocalization.
This is often shown as:

The dotted half-circle suggests the delocalization. Think of it as being like the circle you draw in the middle of the benzene
hexagon. This delocalization is just one single bond away from the ring delocalization. You get conjugation between the
two, and the delocalization takes in the whole molecule.

A case to beware of

Jim Clark 5/28/2021 6 https://chem.libretexts.org/@go/page/3745


Another case we need to look at (because it occurs in a molecule we'll explore on the next page) is an SO3- group
attached to a benzene ring.

It looks as if the double bonds are in just the right position relative to the ring for the delocalization to extend out over this
group. However, I have been told on good authority that the delocalization doesn't extend from the side group into the ring.
I can't, though, find any reference to this anywhere on the web or in the textbooks that I have available. The bonding in the
sulfonate group isn't at all easy to describe in orbital terms. In fact, it is most easily explained in terms of co-ordinate
(dative covalent) bonds, but introducing that now is just complicating things pointlessly.

Summary
When you are trying to work out how far delocalization extends in a molecule, look for:
alternating double and single bonds - not just between carbon and carbon, but including C=O, C=N, N=N, N=O.
Carbon-carbon triple bonds can also be involved in place of a carbon-carbon double bond.
benzene rings.
possible involvement of lone pairs on nitrogen or oxygen.
NO2 groups.
And finally . . . why does all this matter? The wavelength of UV or visible light absorbed by organic compounds depends
largely on the extent of delocalization in the molecules. That means that you may well have to look at an unfamiliar
molecule and make a reasonable estimate of whether it is highly delocalised or not very delocalised. It will always be pretty
clear-cut at this level, so don't worry too much about it!

Contribotors
Jim Clark (Chemguide.co.uk)

Jim Clark 5/28/2021 7 https://chem.libretexts.org/@go/page/3745


Electromagnetic Radiation
This page is a basic introduction to the electromagnetic spectrum sufficient for chemistry students interested in UV-visible
absorption spectroscopy. If you are looking for any sort of explanations suitable for physics courses, then I'm afraid this
isn't the right place for you.

Contributors and Attributions


Jim Clark (Chemguide.co.uk)

Jim Clark 4/28/2021 1 https://chem.libretexts.org/@go/page/3743


The Beer-Lambert Law
The Beer-Lambert law relates the attenuation of light to the properties of the material through which the light is traveling. This
page takes a brief look at the Beer-Lambert Law and explains the use of the terms absorbance and molar absorptivity relating
to UV-visible absorption spectrometry.

The Absorbance of a Solution


For each wavelength of light passing through the spectrometer, the intensity of the light passing through the reference cell is
measured. This is usually referred to as I - that's I for Intensity.
o

Figure 1: Light absorbed by sample in a cuvette


The intensity of the light passing through the sample cell is also measured for that wavelength - given the symbol, I . If I is
less than I , then the sample has absorbed some of the light (neglecting reflection of light off the cuvette surface). A simple bit
o

of math is then done in the computer to convert this into something called the absorbance of the sample - given the symbol, A .
The absorbance of a transition depends on two external assumptions.
1. The absorbance is directly proportional to the concentration (c ) of the solution of the sample used in the experiment.
2. The absorbance is directly proportional to the length of the light path (l), which is equal to the width of the cuvette.
Assumption one relates the absorbance to concentration and can be expressed as
A ∝c (1)

The absorbance (A ) is defined via the incident intensity I and transmitted intensity I by
o

Io
A = log10 ( ) (2)
I

Assumption two can be expressed as


A ∝l (3)

Combining Equations 1 and 3:


A ∝ cl (4)

This proportionality can be converted into an equality by including a proportionality constant (ϵ).
A = ϵcl (5)

This formula is the common form of the Beer-Lambert Law, although it can be also written in terms of intensities:
Io
A = log10 ( ) = ϵlc (6)
I

The constant ϵ is called molar absorptivity or molar extinction coefficient and is a measure of the probability of the
electronic transition. On most of the diagrams you will come across, the absorbance ranges from 0 to 1, but it can go higher
than that. An absorbance of 0 at some wavelength means that no light of that particular wavelength has been absorbed. The
intensities of the sample and reference beam are both the same, so the ratio I /I is 1 and the log of 1 is zero.
o 10

Example 1
In a sample with an absorbance of 1 at a specific wavelength, what is the relative amount of light that was absorbed by the
sample?

Jim Clark 6/19/2021 1 https://chem.libretexts.org/@go/page/3747


Solution
This question does not need Beer-Lambert Law (Equation 5) to solve, but only the definition of absorbance (Equation 2)
Io
A = log ( )
10
I

The relative loss of intensity is


I − Io I
=1−
Io Io

Equation 2 can be rearranged using the properties of logarithms to solved for the relative loss of intensity:
Io
A
10 =
I

−A
I
10 =
Io

−A
I
1 − 10 =1−
Io

Substituting in A = 1
I 1
−1
1− = 1 − 10 =1− = 0.9
Io 10

Hence 90% of the light at that wavelength has been absorbed and that the transmitted intensity is 10% of the incident
intensity. To confirm, substituting these values into Equation 2 to get the absorbance back:
Io 100
= = 10 (7)
I 10

and
log 10 = 1 (8)
10

The Beer-Lambert Law


You will find that various different symbols are given for some of the terms in the equation - particularly for the concentration
and the solution length.

The Greek letter epsilon in these equations is called the molar absorptivity - or sometimes the molar absorption coefficient.
The larger the molar absorptivity, the more probable the electronic transition. In uv spectroscopy, the concentration of the
sample solution is measured in mol L-1 and the length of the light path in cm. Thus, given that absorbance is unitless, the units
of molar absorptivity are L mol-1 cm-1. However, since the units of molar absorptivity is always the above, it is customarily
reported without units.

Example 2 : Guanosine
Guanosine has a maximum absorbance of 275 nm. ϵ = 8400M cm 275 and the path length is 1 cm. Using a
−1 −1

spectrophotometer, you find the that A = 0.70 . What is the concentration of guanosine?
275

Solution
To solve this problem, you must use Beer's Law.

Jim Clark 6/19/2021 2 https://chem.libretexts.org/@go/page/3747


A = ϵlc (9)

-1 -1
0.70 = (8400 M cm )(1 cm)(c )
Next, divide both side by [(8400 M-1 cm-1)(1 cm)]
c = 8.33x10-5 mol/L

Example 3
There is a substance in a solution (4 g/liter). The length of cuvette is 2 cm and only 50% of the certain light beam is
transmitted. What is the extinction coefficient?
Solution
Using Beer-Lambert Law, we can compute the absorption coefficient. Thus,
It 0.5
− log( ) = − log( ) = A = 8ϵ
Io 1.0

Then we obtain that


ϵ = 0.0376

Example 4
In Example 3 above, what is the molar absorption coefficient if the molecular weight is 100?
Solution
It can simply obtained by multiplying the absorption coefficient by the molecular weight. Thus,
ϵ = 0.0376 x 100 = 3.76 L·mol-1·cm-1

The Importance of Concentration


The proportion of the light absorbed will depend on how many molecules it interacts with. Suppose you have got a strongly
colored organic dye. If it is in a reasonably concentrated solution, it will have a very high absorbance because there are lots of
molecules to interact with the light. However, in an incredibly dilute solution, it may be very difficult to see that it is colored at
all. The absorbance is going to be very low. Suppose then that you wanted to compare this dye with a different compound.
Unless you took care to make allowance for the concentration, you couldn't make any sensible comparisons about which one
absorbed the most light.

Example 4
In Example 3 above, how much is the beam of light is transmitted when 8 g/liter ?
Solution
Since we know ϵ, we can calculate the transmission using Beer-Lambert Law. Thus,
log(1) − log(It ) = 0 − log(It ) = 0.0376 x 8 x 2 = 0.6016
log(It ) = -0.6016
Therefore, I = 0.2503 = 25%
t

Example 5
The absorption coefficient of a glycogen-iodine complex is 0.20 at light of 450 nm. What is the concentration when the
transmission is 40 % in a cuvette of 2 cm?
Solution

Jim Clark 6/19/2021 3 https://chem.libretexts.org/@go/page/3747


It can also be solved using Beer-Lambert Law. Therefore,
− log(It ) = − log (0.4) = 0.20 × c × 2 (10)
10

Then c = 0.9948

The importance of the container shape


Suppose this time that you had a very dilute solution of the dye in a cube-shaped container so that the light traveled 1 cm
through it. The absorbance is not likely to be very high. On the other hand, suppose you passed the light through a tube 100 cm
long containing the same solution. More light would be absorbed because it interacts with more molecules. Again, if you want
to draw sensible comparisons between solutions, you have to allow for the length of the solution the light is passing through.
Both concentration and solution length are allowed for in the Beer-Lambert Law.

Molar Absorptivity
The Beer-Lambert law (Equation 5) can be rearranged to obtain an expression for ϵ (the molar absorptivity):
A
ϵ= (11)
lc

Remember that the absorbance of a solution will vary as the concentration or the size of the container varies. Molar
absorptivity compensates for this by dividing by both the concentration and the length of the solution that the light passes
through. Essentially, it works out a value for what the absorbance would be under a standard set of conditions - the light
traveling 1 cm through a solution of 1 mol dm-3. That means that you can then make comparisons between one compound and
another without having to worry about the concentration or solution length.
Values for molar absorptivity can vary hugely. For example, ethanal has two absorption peaks in its UV-visible spectrum - both
in the ultra-violet. One of these corresponds to an electron being promoted from a lone pair on the oxygen into a pi anti-
bonding orbital; the other from a π bonding orbital into a π anti-bonding orbital. Table 1 gives values for the molar
absorptivity of a solution of ethanal in hexane. Notice that there are no units given for absorptivity. That's quite common since
it assumes the length is in cm and the concentration is mol dm-3, the units are mol-1 dm3 cm-1.
Table1
electron jump wavelength of maximum absorption (nm) molar absorptivity

lone pair to π anti-bonding orbital 290 15


π bonding to π anti-bonding orbital 180 10,000

The ethanal obviously absorbs much more strongly at 180 nm than it does at 290 nm. (Although, in fact, the 180 nm
absorption peak is outside the range of most spectrometers.) You may come across diagrams of absorption spectra plotting
absorptivity on the vertical axis rather than absorbance. However, if you look at the figures above and the scales that are going
to be involved, you aren't really going to be able to spot the absorption at 290 nm. It will be a tiny little peak compared to the
one at 180 nm. To get around this, you may also come across diagrams in which the vertical axis is plotted as log10(molar
absorptivity).
If you take the logs of the two numbers in the table, 15 becomes 1.18, while 10,000 becomes 4. That makes it possible to plot
both values easily, but produces strangely squashed-looking spectra!

Contributors and Attributions


Jim Clark (Chemguide.co.uk)
Gamini Gunawardena from the OChemPal site (Utah Valley University)

Jim Clark 6/19/2021 4 https://chem.libretexts.org/@go/page/3747


Using UV-visible Absorption Spectroscopy
This page takes a brief look at how UV-visible absorption spectra can be used to help identify compounds and to measure
the concentrations of colored solutions. It assumes that you know how these spectra arise, and know what is meant by
terms such as absorbance, molar absorptivity and lambda-max. You also need to be familiar with the Beer-Lambert Law.

Contributors and Attributions


Jim Clark (Chemguide.co.uk)

Jim Clark 6/25/2021 1 https://chem.libretexts.org/@go/page/3748


What Causes Molecules to Absorb UV and Visible Light
This page explains what happens when organic compounds absorb UV or visible light, and why the wavelength of light
absorbed varies from compound to compound.

What happens when light is absorbed by molecules?


When we were talking about the various sorts of orbitals present in organic compounds on the introductory page (see above),
you will have come across this diagram showing their relative energies:

Remember that the diagram isn't intended to be to scale - it just shows the relative placing of the different orbitals. When light
passes through the compound, energy from the light is used to promote an electron from a bonding or non-bonding orbital into
one of the empty anti-bonding orbitals. The possible electron jumps that light might cause are:

In each possible case, an electron is excited from a full orbital into an empty anti-bonding orbital. Each jump takes energy
from the light, and a big jump obviously needs more energy than a small one. Each wavelength of light has a particular energy
associated with it. If that particular amount of energy is just right for making one of these energy jumps, then that wavelength
will be absorbed - its energy will have been used in promoting an electron.
We need to work out what the relationship is between the energy gap and the wavelength absorbed. Does, for example, a
bigger energy gap mean that light of a lower wavelength will be absorbed - or what? It is easier to start with the relationship
between the frequency of light absorbed and its energy:

You can see that if you want a high energy jump, you will have to absorb light of a higher frequency. The greater the
frequency, the greater the energy. That's easy - but unfortunately UV-visible absorption spectra are always given using
wavelengths of light rather than frequency. That means that you need to know the relationship between wavelength and
frequency.

Jim Clark 7/2/2021 1 https://chem.libretexts.org/@go/page/3746


You can see from this that the higher the frequency is, the lower the wavelength is. So, if you have a bigger energy jump, you
will absorb light with a higher frequency - which is the same as saying that you will absorb light with a lower wavelength.

Important summary: The larger the energy jump, the lower the wavelength of the light
absorbed.

Some jumps are more important than others for absorption spectrometry
An absorption spectrometer works in a range from about 200 nm (in the near ultra-violet) to about 800 nm (in the very near
infra-red). Only a limited number of the possible electron jumps absorb light in that region. Look again at the possible jumps.
This time, the important jumps are shown in black, and a less important one in grey. The grey dotted arrows show jumps
which absorb light outside the region of the spectrum we are working in.

Remember that bigger jumps need more energy and so absorb light with a shorter wavelength. The jumps shown with grey
dotted arrows absorb UV light of wavelength less that 200 nm. The important jumps are:
from pi bonding orbitals to pi anti-bonding orbitals;
from non-bonding orbitals to pi anti-bonding orbitals;
from non-bonding orbitals to sigma anti-bonding orbitals.
That means that in order to absorb light in the region from 200 - 800 nm (which is where the spectra are measured), the
molecule must contain either pi bonds or atoms with non-bonding orbitals. Remember that a non-bonding orbital is a lone pair
on, say, oxygen, nitrogen or a halogen.

Groups in a molecule which absorb light are known as chromophores.

What does an absorption spectrum look like


The diagram below shows a simple UV-visible absorption spectrum for buta-1,3-diene - a molecule we will talk more about
later. Absorbance (on the vertical axis) is just a measure of the amount of light absorbed. The higher the value, the more of a
particular wavelength is being absorbed.

You will see that absorption peaks at a value of 217 nm. This is in the ultra-violet and so there would be no visible sign of any
light being absorbed - buta-1,3-diene is colorless. You read the symbol on the graph as "lambda-max". In buta-1,3-diene,
CH2=CH-CH=CH2, there are no non-bonding electrons. That means that the only electron jumps taking place (within the
range that the spectrometer can measure) are from pi bonding to pi anti-bonding orbitals.

Jim Clark 7/2/2021 2 https://chem.libretexts.org/@go/page/3746


A chromophore producing two peaks
A chromophore such as the carbon-oxygen double bond in ethanal, for example, obviously has pi electrons as a part of the
double bond, but also has lone pairs on the oxygen atom. That means that both of the important absorptions from the last
energy diagram are possible. You can get an electron excited from a pi bonding to a pi anti-bonding orbital, or you can get one
excited from an oxygen lone pair (a non-bonding orbital) into a pi anti-bonding orbital.

The non-bonding orbital has a higher energy than a pi bonding orbital. That means that the jump from an oxygen lone pair into
a pi anti-bonding orbital needs less energy. That means it absorbs light of a lower frequency and therefore a higher
wavelength. Ethanal can therefore absorb light of two different wavelengths:
the pi bonding to pi anti-bonding absorption peaks at 180 nm;
the non-bonding to pi anti-bonding absorption peaks at 290 nm.
Both of these absorptions are in the ultra-violet, but most spectrometers won't pick up the one at 180 nm because they work in
the range from 200 - 800 nm.

The importance of conjugation and delocalisation


Consider these three molecules:

Ethene contains a simple isolated carbon-carbon double bond, but the other two have conjugated double bonds. In these cases,
there is delocalization of the pi bonding orbitals over the whole molecule. Now look at the wavelengths of the light which each
of these molecules absorbs.
molecule wavelength of maximum absorption (nm)

ethene 171
buta-1,3-diene 217
hexa-1,3,5-triene 258

All of the molecules give similar UV-visible absorption spectra - the only difference being that the absorptions move to longer
and longer wavelengths as the amount of delocalization in the molecule increases.
Why is this? You can actually work out what must be happening.
The maximum absorption is moving to longer wavelengths as the amount of delocalization increases.
Therefore maximum absorption is moving to shorter frequencies as the amount of delocalization increases.
Therefore absorption needs less energy as the amount of delocalization increases.
Therefore there must be less energy gap between the bonding and anti-bonding orbitals as the amount of delocalization
increases.
. . . and that's what is happening.
Compare ethene with buta-1,3-diene. In ethene, there is one pi bonding orbital and one pi anti-bonding orbital. In buta-1,3-
diene, there are two pi bonding orbitals and two pi anti-bonding orbitals. This is all discussed in detail on the introductory page
that you should have read.

Jim Clark 7/2/2021 3 https://chem.libretexts.org/@go/page/3746


The highest occupied molecular orbital is often referred to as the HOMO - in these cases, it is a pi bonding orbital. The lowest
unoccupied molecular orbital (the LUMO) is a pi anti-bonding orbital. Notice that the gap between these has fallen. It takes
less energy to excite an electron in the buta-1,3-diene case than with ethene.
In the hexa-1,3,5-triene case, it is less still.

If you extend this to compounds with really massive delocalisation, the wavelength absorbed will eventually be high enough to
be in the visible region of the spectrum, and the compound will then be seen as colored. A good example of this is the orange
plant pigment, beta-carotene - present in carrots, for example.

Why is beta-carotene orange?


Beta-carotene has the sort of delocalization that we've just been looking at, but on a much greater scale with 11 carbon-carbon
double bonds conjugated together. The diagram shows the structure of beta-carotene with the alternating double and single
bonds shown in red.

The more delocalization there is, the smaller the gap between the highest energy pi bonding orbital and the lowest energy pi
anti-bonding orbital. To promote an electron therefore takes less energy in beta-carotene than in the cases we've looked at so
far - because the gap between the levels is less.
Remember that less energy means a lower frequency of light gets absorbed - and that's equivalent to a longer wavelength.
Beta-carotene absorbs throughout the ultra-violet region into the violet - but particularly strongly in the visible region between
about 400 and 500 nm with a peak about 470 nm. If you have read the page in this section about electromagnetic radiation,
you might remember that the wavelengths associated with the various colors are approximately:
color region wavelength (nm)

violet 380 - 435

blue 435 - 500


cyan 500 - 520
green 520 - 565
yellow 565 - 590
orange 590 - 625
red 625 - 740

So if the absorption is strongest in the violet to cyan region, what color will you actually see? It is tempting to think that you
can work it out from the colors that are left - and in this particular case, you wouldn't be far wrong. Unfortunately, it isn't as
simple as that!

Jim Clark 7/2/2021 4 https://chem.libretexts.org/@go/page/3746


Sometimes what you actually see is quite unexpected. Mixing different wavelengths of light doesn't give you the same result
as mixing paints or other pigments. You can, however, sometimes get some estimate of the color you would see using the idea
of complementary colors.

Complementary colors
If you arrange some colors in a circle, you get a "color wheel". The diagram shows one possible version of this. An internet
search will throw up many different versions!

colors directly opposite each other on the color wheel are said to be complementary colors. Blue and yellow are
complementary colors; red and cyan are complementary; and so are green and magenta. Mixing together two complementary
colors of light will give you white light.
What this all means is that if a particular color is absorbed from white light, what your eye detects by mixing up all the other
wavelengths of light is its complementary color. In the beta-carotene case, the situation is more confused because you are
absorbing such a range of wavelengths. However, if you think of the peak absorption running from the blue into the cyan, it
would be reasonable to think of the color you would see as being opposite that where yellow runs into red - in other words,
orange.

Applying this to the color changes of two indicators


Phenolphthalein
You have probably used phenolphthalein as an acid-base indicator, and will know that it is colorless in acidic conditions and
magenta (bright pink) in an alkaline solution. How is this color change related to changes in the molecule? The structures of
the two differently colored forms are:

Both of these absorb light in the ultra-violet, but the one on the right also absorbs in the visible with a peak at 553 nm. The
molecule in acid solution is colorless because our eyes can't detect the fact that some light is being absorbed in the ultra-violet.
However, our eyes do detect the absorption at 553 nm produced by the form in alkaline solution.
553 nm is in the green region of the spectrum. If you look back at the color wheel, you will find that the complementary color
of green is magenta - and that's the color you see.
So why does the color change as the structure changes? What we have is a shift to absorption at a higher wavelength in
alkaline solution. As we've already seen, a shift to higher wavelength is associated with a greater degree of delocalisation.
Here is a modified diagram of the structure of the form in acidic solution - the colorless form. The extent of the delocalization
is shown in red.

Jim Clark 7/2/2021 5 https://chem.libretexts.org/@go/page/3746


Notice that there is delocalization over each of the three rings - extending out over the carbon-oxygen double bond, and to the
various oxygen atoms because of their lone pairs.
But the delocalization doesn't extend over the whole molecule. The carbon atom in the centre with its four single bonds
prevents the three delocalized regions interacting with each other.
Now compare that with the magenta form:

The rearrangement now lets the delocalization extend over the entire ion. This greater delocalization lowers the energy gap
between the highest occupied molecular orbital and the lowest unoccupied pi anti-bonding orbital. It needs less energy to make
the jump and so a longer wavelength of light is absorbed.

Increasing the amount of delocalization shifts the absorption peak to a higher


wavelength.

Methyl orange
You will know that methyl orange is yellow in alkaline solutions and red in acidic ones. The structure in alkaline solution is:

In acid solution, a hydrogen ion is (perhaps unexpectedly) picked up on one of the nitrogens in the nitrogen-nitrogen double
bond.

This now gets a lot more complicated! The positive charge on the nitrogen is delocalized (spread around over the structure) -
especially out towards the right-hand end of the molecule as we've written it. The normally drawn structure for the red form of
methyl orange is . . .

But this can be seriously misleading as regards the amount of delocalization in the structure for reasons discussed below (after
the red warning box) if you are interested.

Which is the more delocalized structure?


Let's work backwards from the absorption spectra to see if that helps. The yellow form has an absorption peak at about 440
nm. That's in the blue region of the spectrum, and the complementary color of blue is yellow. That's exactly what you would
expect. The red form has an absorption peak at about 520 nm. That's at the edge of the cyan region of the spectrum, and the
complementary color of cyan is red. Again, there's nothing unexpected here.
Notice that the change from the yellow form to the red form has produced an increase in the wavelength absorbed. An increase
in wavelength suggests an increase in delocalisation. That means that there must be more delocalization in the red form than in
the yellow one. Here again is the structure of the yellow form:

Jim Clark 7/2/2021 6 https://chem.libretexts.org/@go/page/3746


delocalization will extend over most of the structure - out as far as the lone pair on the right-hand nitrogen atom.
If you use the normally written structure for the red form, the delocalization seems to be broken in the middle - the pattern of
alternating single and double bonds seems to be lost.

But that is to misunderstand what this last structure represents.

Canonical forms
If you draw the two possible Kekulé structures for benzene, you will know that the real structure of benzene isn't like either of
them. The real structure is somewhere between the two - all the bonds are identical and somewhere between single and double
in character. That's because of the delocalization in benzene.

The two structures are known as canonical forms, and they can each be thought of as adding some knowledge to the real
structure. For example, the bond drawn at the top right of the molecule is neither truly single or double, but somewhere in
between. Similarly with all the other bonds.
The two structures we've previously drawn for the red form of methyl orange are also canonical forms - two out of lots of
forms that could be drawn for this structure. We could represent the delocalized structure by:

These two forms can be thought of as the result of electron movements in the structure, and curly arrows are often used to
show how one structure can lead to the other.

In reality, the electrons haven't shifted fully either one way or the other. Just as in the benzene case, the actual structure lies
somewhere in between these.
You must also realize that drawing canonical forms has no effect on the underlying geometry of the structure. Bond types or
lengths or angles don't change in the real structure.
For example, the lone pairs on the nitrogen atoms shown in the last diagram are both involved with the delocalisation. For this
to happen all the bonds around these nitrogens must be in the same plane, with the lone pair sticking up so that it can overlap
sideways with orbitals on the next-door atoms. The fact that in each of the two canonical forms one of these nitrogens is
shown as if it had an ammonia-like arrangement of the bonds is potentially misleading - and makes it look as if the
delocalization is broken.
The problem is that there is no easy way of representing a complex delocalized structure in simple structural diagrams. It is
bad enough with benzene - with something as complicated as methyl orange any method just leads to possible confusion if you

Jim Clark 7/2/2021 7 https://chem.libretexts.org/@go/page/3746


aren't used to working with canonical forms.
It gets even more complicated! If you were doing this properly there would be a host of other canonical forms with different
arrangements of double and single bonds and with the positive charge located at various places around the rings and on the
other nitrogen atom.
The real structure can't be represented properly by any one of this multitude of canonical forms, but each gives a hint of how
the delocalization works.
If we take the two forms we have written as perhaps the two most important ones, it suggests that there is delocalization of the
electrons over the whole structure, but that electron density is a bit low around the two nitrogens carrying the positive charge
on one canonical form or the other.

Why is the red form more delocalized


Finally, we get around to an attempt at an explanation as to why the delocalization is greater in the red form of methyl orange
in acid solution than in the yellow one in alkaline solution. The answer may lie in the fact that the lone pair on the nitrogen at
the right-hand end of the structure as we've drawn it is more fully involved in the delocalization in the red form. The canonical
form with the positive charge on that nitrogen suggests a significant movement of that lone pair towards the rest of the
molecule.
Doesn't the same thing happen to the lone pair on the same nitrogen in the yellow form of methyl orange? Not to the same
extent.
Any canonical form that you draw in which that happens produces another negatively charged atom somewhere in the rest of
the structure. Separating negative and positive charges like this is energetically unfavourable. In the red form, we aren't
producing a new separation of charge - just shifting a positive charge around the structure.

Contributors and Attributions


Jim Clark (Chemguide.co.uk)

Jim Clark 7/2/2021 8 https://chem.libretexts.org/@go/page/3746


Fluorescence and Phosphorescence
Fluorescence and phosphorescence are types of molecular luminescence methods. A molecule of analyte absorbs a photon and
excites a species. The emission spectrum can provide qualitative and quantitative analysis. The term fluorescence and
phosphorescence are usually referred as photoluminescence because both are alike in excitation brought by absorption of a
photon. Fluorescence differs from phosphorescence in that the electronic energy transition that is responsible for fluorescence
does not change in electron spin, which results in short-live electrons (<10-5 s) in the excited state of fluorescence. In
phosphorescence, there is a change in electron spin, which results in a longer lifetime of the excited state (second to minutes).
Fluorescence and phosphorescence occurs at longer wavelength than the excitation radiation.

Introduction
Fluorescence can occur in gaseous, liquid, and solid chemical systems. The simple kind of fluorescence is by dilute atomic
vapors. A fluorescence example would be if a 3s electron of a vaporized sodium atom is excited to the 3p state by absorption
of a radiation at wavelength 589.6 and 589.0 nm. After 10-8 s, the electron returns to ground state and on its return it emits
radiation of the two wavelengths in all directions. This type of fluorescence in which the absorbed radiation is remitted without
a change in frequency is known as resonance fluorescence. Resonance fluorescence can also occur in molecular species.
Molecular fluorescence band centers at wavelengths longer than resonance lines. The shift toward longer wavelength is
referred to as the Stokes Shift.

Singlet and Triplet Excited State


Understanding the difference between fluorescence and phosphorescence requires the knowledge of electron spin and the
differences between singlet and triplet states. The Pauli Exclusion principle states that two electrons in an atom cannot have
the same four quantum numbers (n , l, m , m ) and only two electrons can occupy each orbital where they must have opposite
l s

spin states. These opposite spin states are called spin pairing. Because of this spin pairing, most molecules do not exhibit a
magnetic field and are diamagnetic. In diamagnetic molecules, electrons are not attracted or repelled by the static electric field.
Free radicals are paramagnetic because they contain unpaired electrons have magnetic moments that are attracted to the
magnetic field.
Singlet state is defined when all the electron spins are paired in the molecular electronic state and the electronic energy levels
do not split when the molecule is exposed into a magnetic field. A doublet state occurs when there is an unpaired electron that
gives two possible orientations when exposed in a magnetic field and imparts different energy to the system. A singlet or a
triplet can form when one electron is excited to a higher energy level. In an excited singlet state, the electron is promoted in the
same spin orientation as it was in the ground state (paired). In a triplet excited stated, the electron that is promoted has the
same spin orientation (parallel) to the other unpaired electron. The difference between the spins of ground singlet, excited
singlet, and excited triplet is shown in Figure 1. Singlet, doublet and triplet is derived using the equation for multiplicity,
2S+1, where S is the total spin angular momentum (sum of all the electron spins). Individual spins are denoted as spin up (s =
+1/2) or spin down (s = -1/2). If we were to calculated the S for the excited singlet state, the equation would be 2(+1/2 +
-1/2)+1 = 2(0)+1 = 1, therefore making the center orbital in the figure a singlet state. If the spin multiplicity for the excited
triplet state was calculated, we obtain 2(+1/2 + +1/2)+1 = 2(1)+1 =3, which gives a triplet state as expected.

Figure 1: Spin in the ground and excited states.


The difference between a molecule in the ground and excited state is that the electrons is diamagnetic in the ground state and
paramagnetic in the triplet state.This difference in spin state makes the transition from singlet to triplet (or triplet to singlet)
more improbable than the singlet-to-singlet transitions. This singlet to triplet (or reverse) transition involves a change in
electronic state. For this reason, the lifetime of the triplet state is longer the singlet state by approximately 104 seconds fold
difference.The radiation that induced the transition from ground to excited triplet state has a low probability of occurring, thus
their absorption bands are less intense than singlet-singlet state absorption. The excited triplet state can be populated from the

7/7/2021 1 https://chem.libretexts.org/@go/page/1765
excited singlet state of certain molecules which results in phosphorescence. These spin multiplicities in ground and excited
states can be used to explain transition in photoluminescence molecules by the Jablonski diagram.

Jablonski Diagrams
The Jablonski diagram that drawn below is a partial energy diagram that represents the energy of photoluminescent molecule
in its different energy states. The lowest and darkest horizontal line represents the ground-state electronic energy of the
molecule which is the singlet state labeled as S . At room temperature, majority of the molecules in a solution are in this state.
o

Figure 2: Partial Jablonski Diagram for Absorption, Fluorescence, and Phosphorescence. from Bill Reusch.
The upper lines represent the energy state of the three excited electronic states: S1and S2 represent the electronic singlet state
(left) and T1 represents the first electronic triplet state (right). The upper darkest line represents the ground vibrational state of
the three excited electronic state.The energy of the triplet state is lower than the energy of the corresponding singlet state.
There are numerous vibrational levels that can be associated with each electronic state as denoted by the thinner lines.
Absorption transitions (blues lines in Figure 2) can occur from the ground singlet electronic state (So) to various vibrational
levels in the singlet excited vibrational states. It is unlikely that a transition from the ground singlet electronic state to the
triplet electronic state because the electron spin is parallel to the spin in its ground state (Figure 1). This transition leads to a
change in multiplicity and thus has a low probability of occurring which is a forbidden transition. Molecules also go through
vibration relaxation to lose any excess vibrational energy that remains when excited to the electronic states (S and S ) as 1 2

demonstrated in wavy lines in Figure 2. The knowledge of forbidden transition is used to explain and compare the peaks of
absorption and emission.

Absorption and Emission Rates


The table below compares the absorption and emission rates of fluorescence and phosphorescence.The rate of photon
absorption is very rapid. Fluorescence emission occurs at a slower rate.Since the triplet to singlet (or reverse) is a forbidden
transition, meaning it is less likely to occur than the singlet-to-singlet transition, the rate of triplet to singlet is typically slower.
Therefore, phosphorescence emission requires more time than fluorescence.
Table 1: Rates of Absorption and Emission comparison.
Process Transition Timescale (sec)

Light Absorption (Excitation) S0 → Sn ca. 10-15 (instantaneous)

Internal Conversion Sn → S1 10-14 to 10-11


Vibrational Relaxation Sn* → Sn 10-12 to 10-10
Intersystem Crossing S1 → T1 10-11 to 10-6
Fluorescence S1 → S0 10-9 to 10-6
Phosphorescence T1 → S0 10-3 to 100
S1 → S0 10-7 to 10-5
Non-Radiative Decay
T1 → S0 10-3 to 100

Deactivation Processes

7/7/2021 2 https://chem.libretexts.org/@go/page/1765
A molecule that is excited can return to the ground state by several combinations of mechanical steps that will be described
below and shown in Figure 2.The deactivation process of fluorescence and phosphorescence involve an emission of a photon
radiation as shown by the straight arrow in Figure 2. The wiggly arrows in Figure 2 are deactivation processes without the use
of radiation. The favored deactivation process is the route that is most rapid and spends less time in the excited state.If the rate
constant for fluorescence is more favorable in the radiationless path, the fluorescence will be less intense or absent.
Vibrational Relaxation: A molecule maybe to promoted to several vibrational levels during the electronic excitation
process.Collision of molecules with the excited species and solvent leads to rapid energy transfer and a slight increase in
temperature of the solvent. Vibrational relaxation is so rapid that the lifetime of a vibrational excited molecule (<10-12) is
less than the lifetime of the electronically excited state. For this reason, fluorescence from a solution always involves the
transition of the lowest vibrational level of the excited state. Since the space of the emission lines are so close together, the
transition of the vibrational relaxation can terminate in any vibrational level of the ground state.
Internal Conversion: Internal conversion is an intermolecular process of molecule that passes to a lower electronic state
without the emission of radiation.It is a crossover of two states with the same multiplicity meaning singlet-to-singlet or
triplet-to-triplet states.The internal conversion is more efficient when two electronic energy levels are close enough that
two vibrational energy levels can overlap as shown in between S1 and S2. Internal conversion can also occur between S0
and S1 from a loss of energy by fluorescence from a higher excited state, but it is less probable. The mechanism of internal
conversion from S1 to S0 is poorly understood. For some molecules, the vibrational levels of the ground state overlaps with
the first excited electronic state, which leads to fast deactivation.These usually occur with aliphatic compounds (compound
that do not contain ring structure), which would account for the compound is seldom fluorescing. Deactivation by energy
transfer of these molecules occurs so rapidly that the molecule does not have time to fluoresce.
External Conversion: Deactivation of the excited electronic state may also involve the interaction and energy transfer
between the excited state and the solvent or solute in a process called external conversion. Low temperature and high
viscosity leads to enhanced fluorescence because they reduce the number of collision between molecules, thus slowing
down the deactivation process.
Intersystem Crossing: Intersystem crossing is a process where there is a crossover between electronic states of different
multiplicity as demonstrated in the singlet state to a triplet state (S1 to T1) on Figure 1. The probability of intersystem
crossing is enhanced if the vibration levels of the two states overlap. Intersystem crossing is most commonly observed with
molecules that contain heavy atom such as iodine or bromine. The spin and orbital interaction increase and the spin become
more favorable.Paramagnetic species also enhances intersystem crossing, which consequently decreases fluorescence.
Phosphorescence: Deactivation of the electronic excited state is also involved in phosphorescence. After the molecule
transitions through intersystem crossing to the triplet state, further deactivation occurs through internal or external
fluorescence or phosphorescence. A triplet-to-singlet transition is more probable than a singlet-to-singlet internal crossing.
In phosphorescence, the excited state lifetime is inversely proportional to the probability that the molecule will transition
back to the ground state. Since the lifetime of the molecule in the triplet state is large (10-4 to 10 second or more), transition
is less probable which suggest that it will persist for some time even after irradiation has stopped. Since the external and
internal conversion compete so effectively with phosphorescence, the molecule has to be observed at lower temperature in
highly viscous media to protect the triplet state.

Variables that affect Fluorescence


After discussing all the possible deactivation processes, variable that affect the emissions to occur. Molecular structure and its
chemical environment influence whether a substance will fluoresce and the intensities of these emissions. The quantum yield
or quantum efficiency is used to measure the probability that a molecule will fluoresce or phosphoresce. For fluorescence and
phosphorescence is the ratio of the number of molecules that luminescent to the total number of excited molecules. For highly
fluoresce molecules, the quantum efficiency approaches to one.Molecules that do not fluoresce have quantum efficiencies that
approach to zero.
Fluorescence quantum yield (ϕ ) for a compound is determined by the relative rate constants (k) of various deactivation
processes by which the lowest excited singlet state is deactivated to the ground state. The deactivation processes including
fluorescence (kf), intersystem crossing (k ), internal conversion (kic), predissociation (kpd), dissociation (kd), and external
i

conversion (kec) allows one to qualitatively interpret the structural and environmental factors that influence the intensity of the
fluorescence. They are related by the quantum yield equation given below:

7/7/2021 3 https://chem.libretexts.org/@go/page/1765
kf
(1)
kf + ki + kec + kic + kpd + kd

Using this equation as an example to explain fluorescence, a high fluorescence rate (kf) value and low values of the all the
other relative rate constant terms (kf +ki+kec+kic+kpd+kd) will give a large ϕ , which suggest that fluorescence is enhanced. The
magnitudes of kf , kd, and kpd depend on the chemical structure, while the rest of the constants ki, kec, and kic are strongly
influenced by the environment.
Fluorescence rarely results from absorption of ultraviolet radiation of wavelength shorter than 250 nm because radiation at this
wavelength has sufficient energy to deactivate the electron in the excited state by predissociation or dissociation. The bond of
some organic molecules would rupture at 140 kcal/mol, which corresponds to 200-nm of radiation. For this reason, σ → σ ∗

transition in fluorescence are rarely observed. Instead, emissions from the less energetic transition will occur which are either
π → π or π → n transition.
∗ ∗

Molecules that are excited electronically will return to the lowest excited state by rapid vibrational relaxation and internal
conversion, which produces no radiation emission. Fluorescence arises from a transition from the lowest vibrational level of
the first excited electronic state to one of the vibrational levels in the electronic ground state. In most fluorescent compounds,
radiation is produced by a π → π or π → n transition depending on which requires the least energy for the transition to
∗ ∗

occur.
Fluorescence is most commonly found in compounds in which the lowest energy transition is π → π (excited singlet state)

than n → π which suggest that the quantum efficiency is greater for π → π transitions. The reason for this is that the molar
∗ ∗

absorptivity, which measures the probability that a transition will occur, of the π → π transition is 100 to 1000 fold greater

than n → π process. The lifetime of π → π (10-7 to 10-9 s) is shorter than the lifetime of n → π (10-5 to 10-7).
∗ ∗ ∗

Phosphorescent quantum efficiency is the opposite of fluorescence in that it occurs in the n → π excited state which tends to

be short lived and less suceptable to deactivation than the π → π triplet state. Intersystem crossing is also more probable for

π → π

excited state than for the n → π state because the energy difference between the singlet and triplet state is large and

spin-orbit coupling is less likely to occur.

Fluorescence and Structure


The most intense fluorescence is found in compounds containing aromatic group with low-energy π → π transitions. A few∗

aliphatic, alicyclic carbonyl, and highly conjugated double-bond structures also exhibit fluorescence as well. Most
unsubstituted aromatic hydrocarbons fluoresce in solution too. The quantum efficiency increases as the number of rings and
the degree of condensation increases. Simple heterocycles such as the structures listed below do not exhibit fluorescence.

Pyridine Pyrrole Furan Thiophene


With nitrogen heterocyclics, the lowest energy transitions is involved in n → π system that rapidly converts to the triplet

state and prevents fluorescence. Although simple heterocyclics do not fluoresce, fused-ring structures do. For instance, a
fusion of a benzene ring to a hetercyclic structure results in an increase in molar absorptivity of the absorption band. The
lifetime of the excited state in fused structure and fluorescence is observed. Examples of fluorescent compounds is shown
below.

quinoline
Benzene ring substitution causes a shift in the absorption maxima of the wavelength and changes in fluorescence emission.
The table below is used to demonstrate and visually show that as benzene is substituted with increasing methyl addition, the
relative intensity of fluorescence increases.
Table 2. Relative intensity of fluorescence comparison with alkane substituted benzenes.

7/7/2021 4 https://chem.libretexts.org/@go/page/1765
Compound Structure Wavelength of Fluorescence (nm) Relative intensity of Fluorescence

Benzene 270-310 10

Toluene 270-320 17

Propyl Benzene 270-320 17

The relative intensity of fluorescence increases as oxygenated species increases in substitution. The values for such increase is
demonstrated in the table below.
Table 3: Relative intensity of fluorescence comparison with benzene with oxygenated substituted benzene
Compound Structure Wavelength of Fluorescence (nm) Relative intensity of Fluorescence

Phenol 285-365 18

Phenolate ion 310-400 10

Anisole 285-345 20

Influence of a halogen substitution decreases fluorescence as the molar mass of the halogen increases. This is an example of
the “heavy atom effect” which suggest that the probability of intersystem crossing increases as the size of the molecule
increases. As demonstrated in the table below, as the molar mass of the substituted compound increases, the relative intensity
of the fluorescence decreases.
Table 4: Relative intensity fluorescence comparison with halogen substituted compounds
Compound Structure Wavelength of Fluorescence (nm) Relative intensity of Fluorescence

Fluorobenzene 270-320 10

Chlorobenzene 275-345 7

Bromobenzene 290-380 5

In heavy atom substitution such as nitro derivatives or heavy halogen substitution such as iodobenzene, the compounds are
subject to predissociation. These compounds have bonds that easily rupture that can then absorb excitation energy and go
through internal conversion. Therefore, the relative intensity of fluorescence and fluorescent wavelength is not observed and
this is demonstrated in the table below.
Table 5: Relative fluorescent intensities of iodobenzene and nitro derivative compounds
Compound Structure Wavelength of Fluorescence (nm) Relative intensity of Fluorescence

Iodobenzene None 0

Anilinium ion None 0

7/7/2021 5 https://chem.libretexts.org/@go/page/1765
Compound Structure Wavelength of Fluorescence (nm) Relative intensity of Fluorescence

Nitrobenzene None 0

Carboxylic acid or carbonyl group on aromatic ring generally inhibits fluorescence since the energy of the n → π transition ∗

is less than π → π transition. Therefore, the fluorescence yield from n → π transition is low.
∗ ∗

Table 6: Relative fluorescent intensity of benzoic acid


Compound Structure Wavelength of Fluorescence (nm) Relative intensity of Fluorescence

Benzoic Acid 310-390 3

Effect of Structural Rigidity on Fluorescence


Fluorescence is particularly favored in molecules with rigid structures. The table below compares the quantum efficiencies of
fluorine and biphenyl which are both similar in structure that there is a bond between the two benzene group. The difference is
that fluorene is more rigid from the addition methylene bridging group. By looking at the table below, rigid fluorene has a
higher quantum efficiency than unrigid biphenyl which indicates that fluorescence is favored in rigid molecules.
Table 7: Quantum Efficiencies in Rigid vs. Nonrigid structures
Compound Structure Quantum Efficiency

Fluorene 1.0

Biphenyl 0.2

This concept of rigidity was used to explain the increase in fluorescence of organic chelating agent when the compound is
complexed with a metal ion. The fluorescence intensity of 8-hydroxyquinoline is much less than its zinc complex.

vs
8-hydroxyquinoline 8-hydroxyquinoline with Zinc complexed
The explanation for lower quantum efficiency or lack of rigidity in caused by the enhanced internal conversion rate (kic) which
increases the probability that there will be radiationless deactivation. Nonrigid molecules can also undergo low-frequency
vibration which accounts for small energy loss.

Temperature and Solvent Effects


Quantum efficiency of Fluorescence decreases with increasing temperature. As the temperature increases, the frequency of the
collision increases which increases the probability of deactivation by external conversion. Solvents with lower viscosity have
higher possibility of deactivation by external conversion. Fluorescence of a molecule decreases when its solvent contains
heavy atoms such as carbon tetrabromide and ethyl iodide, or when heavy atoms are substituted into the fluorescing
compound. Orbital spin interaction result from an increase in the rate of triplet formation, which decreases the possibility of
fluorescence. Heavy atoms are usually incorporated into solvent to enhance phosphorescence.

Effect of pH on Fluorescence
The fluorescence of aromatic compound with basic or acid substituent rings are usually pH dependent. The wavelength and
emission intensity is different for protonated and unprotonated forms of the compound as illustrated in the table below:

7/7/2021 6 https://chem.libretexts.org/@go/page/1765
Table 8: Quantum efficiency comparison due to protonation
Compound Structure Wavelength of Fluorescence (nm) Relative intensity of Fluorescence

aniline 310-405 20

Anilinium ion None 0

The emission changes of this compound arises from different number of resonance structures associated with the acidic and
basic forms of the molecule.The additional resonance forms provides a more stable first excited state, thus leading to
fluorescence in the ultraviolet region.The resonance structures of basic aniline and acidic anilinium ion is shown below:

basic Aniline Fluorescence of certain compounds have been used a detection of end points in acid-base titrations.An example
of this type of fluorescence seen in compound as a function of pH is the phenolic form of 1-naphthol-4-sulfonic acid.This
compound is not detectable with the eye because it occurs in the ultraviolet region, but with an addition of a base, it becomes
converted to a phenolate ion, the emission band shifts to the visible wavelength where it can be visually seen. Acid
dissociation constant for excited molecules differs for the same species in the ground state.These changes in acid or base
dissociation constant differ in four or five orders of magnitude.

Dissolved oxygen reduces the intensity of fluorescence in solution, which results from a photochemically induced oxidation of
fluorescing species.Quenching takes place from the paramagnetic properties of molecular oxygen that promotes intersystem
crossing and conversion of excited molecules to triplet state.Paramagnetic properties tend to quench fluorescence.

Effects of Concentration on Fluorescence Intensity


The power of fluorescence emission F is proportional to the radiant power is proportional to the radiant power of the
excitation beam that is absorbed by the system. The equation below best describes this relationship.

(1)
Since ϕ f K” is constant in the system, it is represented at K’. The table below defines the variables in this equation.
Table 9: Definitions of all the variables defined in the Fluorescence Emission (F) in Equation 1.
Variable Definition

F Power of fluorescence emission


P0 Power of incident beam on solution
P Power after transversing length b in medium
K” Constant dependent on geometry and other factors

f Quantum efficiency

7/7/2021 7 https://chem.libretexts.org/@go/page/1765
Fluorescence emission (F ) can be related to concentration (c ) using Beer’s Law stating:
F = ϵbc (2)

where ϵ is the molar absorptivity of the molecule that is fluorescing. Rewriting Equation 2 gives:
−ϵbc
P = Po 10 (3)

Plugging this Equation 3 into Equation ??? and factoring out \(P_0\) gives us this equation:
′ −εbc
F = K P0 (1 − 10 ) (4)

The MacLaurin series could be used to solved the exponential term.


2 3 4 n
(2.303εbc) (2.303εbc) (2.303εbc) (2.303εbc)

F = K P0 [2.303εbc − + + +… ] (5)
2! 3! 4! n!
1

Given that (2.303ϵbc = Absorbance < 0.05, all the subsequent terms after the first can be dropped since the maximum error
is 0.13%. Using only the first term, Equation ??? can be rewritten as:

F = K P0 2.303εbc (6)

Equation ??? can be expanded to the equation below and simplified to compare the fluorescence emission F with
concentration. If the equation below were to be plotted with F versus c, a linear relation would be observed.
′′
F = ϕf K P0 2.303εbc (7)

If c becomes so great that the absorbance > 0.05, the higher terms start to become taken into account and the linearity is lost. F
then lies below the extrapolation of the straight-line plot. This excessive absorption is the primary absorption. Another cause
of this negative downfall of linearity is the secondary absorption when the wavelength of emission overlaps the absorption
band. This occurs when the emission transverse the solution and gets reabsorbed by other molecules by analyte or other
species in the solution, which leads to a decrease in fluorescence.

Quenching Methods
Dynamic Quenching is a nonradiative energy transfer between the excited and the quenching agent species (Q).The
requirements for a successful dynamic quenching are that the two collision species the concentration must be high so that there
is a higher possibility of collision between the two species.Temperature and quenching agent viscosity play a role on the rate
of dynamic quenching.Dynamic quenching reduces fluorescence quantum yield and the fluorescence lifetime.
Dissolved oxygen in a solution increases the intensity of the fluorescence by photochemically inducing oxidation of the
fluorescing species.Quenching results from the paramagnetic properties of molecular oxygen that promotes intersystem
crossing and converts the excited molecules to triplet state.Paramagnetic species and dissolved oxygen tend to quench
fluorescence and quench the triplet state.
Static quenching occurs when the quencher and ground state fluorophore forms a dark complex.Fluorescence is usually
observed from unbound fluorophore.Static quenching can be differentiated from dynamic quenching in that the lifetime is not
affected in static quenching.In long range (Förster) quenching, energy transfer occurs without collision between molecules, but
dipole-dipole coupling occurs between excited fluorophore and quencher.

Emission and Excitation Spectra


One of the ways to visually distinguish the difference between each photoluminescence is to compare the relative intensities of
emission/excitation at each wavelength. An example of the three types of photoluminescence (absorption, fluorescence and
phosphorescence) is shown for phenanthrene in the spectrum below.In the spectrum, the luminescent intensity is measure in a
wavelength is fixed while the excitation wavelength is varied. The spectrum in red represents the excitation spectrum, which is
identical to the absorption spectrum because in order for fluorescence emission to occur, radiation needs to be absorbed to
create an excited state.The spectrum in blue represent fluorescence and green spectrum represents the phosphorescence.

7/7/2021 8 https://chem.libretexts.org/@go/page/1765
Figure 3: Wavelength Intensities of Absorption, Fluorescence, and Phosphorescence
Fluorescence and Phosphorescence occur at wavelengths that are longer than their absorption wavelengths.Phosphorescence
bands are found at a longer wavelength than fluorescence band because the excited triplet state is lower in energy than the
singlet state.The difference in wavelength could also be used to measure the energy difference between the singlet and triplet
state of the molecule. The wavelength (λ ) of a molecule is inversely related to the energy (E ) by the equation below:
hc
E = (8)
λ

As the wavelength increases, the energy of the molecule decrease and vice versa.

References
1. D. A. Skoog, et al. "Principles of Instrumental Analysis" 6th Edition, Thomson Brooks/Cole. 2007
2. D. C. Harris and M.D. Bertolucci "Symmetry and Spectroscopy, An Introduction to Vibrational and Electronic
Spectroscopy" Dover Publications, Inc., New York. 1989.

Problems
1. Draw and label the Jablonski Diagram.
2. How do spin states differ in ground singlet state versus excite singlet state and triplet excited state?
3. Describe the rates of deactivation process.
4. What is quantum yield and how is it used to compare the fluorescence of different types of molecule?
5. What roles do solvent play in fluorescence?

Contributors and Attributions


Diana Wong (UCD)

7/7/2021 9 https://chem.libretexts.org/@go/page/1765
Jablonski diagram
Aleksander Jablonski was a Polish academic who devoted his life to the study of molecular absorbance and emission of light.
He developed a written representation that generally shows a portion of the possible consequences of applying photons from
the visible spectrum of light to a particular molecule. These schematics are referred to as Jablonski diagrams.

Introduction
A Jablonski diagram is basically an energy diagram, arranged with energy on a vertical axis. The energy levels can be
quantitatively denoted, but most of these diagrams use energy levels schematically. The rest of the diagram is arranged into
columns. Every column usually represents a specific spin multiplicity for a particular species. However, some diagrams divide
energy levels within the same spin multiplicity into different columns. Within each column, horizontal lines represent
eigenstates for that particular molecule. Bold horizontal lines are representations of the limits of electronic energy states.
Within each electronic energy state are multiple vibronic energy states that may be coupled with the electronic state. Usually
only a portion of these vibrational eigenstates are represented due to the massive number of possible vibrations in a molecule.
Each of these vibrational energy states can be subdivided even further into rotational energy levels; however, typical Jablonski
diagrams omit such intense levels of detail. As electronic energy states increase, the difference in energy becomes continually
less, eventually becoming a continuum that can be approach with classical mechanics. Additionally, as the electronic energy
levels get closer together, the overlap of vibronic energy levels increases.

Figure 1: The Foundation of a typical Jablonski Diagram


Through the use of straight and curved lines, these figures show transitions between eigenstates that occur from the exposure
of a molecule to a particular wavelength of light. Straight lines show the conversion between a photon of light and the energy
of an electron. Curved lines show transitions of electrons without any interaction with light. Within a Jablonski diagram
several different pathways show how an electron may accept and then dissipate the energy from a photon of a particular
wavelength. Thus, most diagrams start with arrows going from the ground electronic state and finish with arrows going to the
ground electronic state.

Absorbance
The first transition in most Jablonski diagrams is the absorbance of a photon of a particular energy by the molecule of interest.
This is indicated by a straight arrow pointing up. Absorbance is the method by which an electron is excited from a lower
energy level to a higher energy level. The energy of the photon is transferred to the particular electron. That electron then
transitions to a different eigenstate corresponding to the amount of energy transferred. Only certain wavelengths of light are
possible for absorbance, that is, wavelengths that have energies that correspond to the energy difference between two different
eigenstates of the particular molecule. Absorbance is a very fast transition, on the order of 10-15 seconds. Most Jablonski
diagrams, however, do not indicate a time scale for the phenomenon being indicated. This transition will usually occur from
the lowest (ground) electronic state due to the statistical mechanical issue of most electrons occupying a low lying state at
reasonable temperatures. There is a Boltzmann distribution of electrons within this low lying levels, based on the the energy
available to the molecules. This energy available is a function of the Boltzmann's constant and the temperature of the system.
These low lying electrons will transition to an excited electronic state as well as some excited vibrational state.

7/4/2021 1 https://chem.libretexts.org/@go/page/1769
Figure 2: Three possible absorption transitions represented.

Vibrational Relaxation and Internal Conversion


Once an electron is excited, there are a multitude of ways that energy may be dissipated. The first is through vibrational
relaxation, a non-radiative process. This is indicated on the Jablonski diagram as a curved arrow between vibrational levels.
Vibrational relaxation is where the energy deposited by the photon into the electron is given away to other vibrational modes
as kinetic energy. This kinetic energy may stay within the same molecule, or it may be transferred to other molecules around
the excited molecule, largely depending on the phase of the probed sample. This process is also very fast, between 10-14 and
10-11 seconds. Since this is a very fast transition, it is extremely likely to occur immediately following absorbance. This
relaxation occurs between vibrational levels, so generally electrons will not change from one electronic level to another
through this method.
However, if vibrational energy levels strongly overlap electronic energy levels, a possibility exists that the excited electron can
transition from a vibration level in one electronic state to another vibration level in a lower electronic state. This process is
called internal conversion and mechanistically is identical to vibrational relaxation. It is also indicated as a curved line on a
Jablonski diagram, between two vibrational levels in different electronic states. Internal Conversion occurs because of the
overlap of vibrational and electronic energy states. As energies increase, the manifold of vibrational and electronic eigenstates
becomes ever closer distributed. At energy levels greater than the first excited state, the manifold of vibrational energy levels
strongly overlap with the electronic levels. This overlap gives a higher degree of probability that the electron can transition
between vibrational levels that will lower the electronic state. Internal conversion occurs in the same time frame as vibrational
relaxation, therefore, is a very likely way for molecules to dissipate energy from light perturbation. However, due to a lack of
vibrational and electronic energy state overlap and a large energy difference between the ground state and first excited state,
internal conversion is very slow for an electron to return to the ground state. This slow return to the ground state lets other
transitive processes compete with internal conversion at the first electronically excited state. Both vibrational relaxation and
internal conversion occur in most perturbations, yet are seldom the final transition.

Figure 3: Possible scenario with absorption, internal conversion, and vibrational relaxation processes shown.

Fluorescence
Another pathway for molecules to deal with energy received from photons is to emit a photon. This is termed fluorescence. It
is indicated on a Jablonski diagram as a straight line going down on the energy axis between electronic states. Fluorescence is
a slow process on the order of 10-9 to 10-7 seconds; therefore, it is not a very likely path for an electron to dissipate energy
especially at electronic energy states higher than the first excited state. While this transition is slow, it is an allowed transition
with the electron staying in the same multiplicity manifold. Fluorescence is most often observed between the first excited

7/4/2021 2 https://chem.libretexts.org/@go/page/1769
electron state and the ground state for any particular molecule because at higher energies it is more likely that energy will be
dissipated through internal conversion and vibrational relaxation. At the first excited state, fluorescence can compete in regard
to timescales with other non-radiative processes. The energy of the photon emitted in fluorescence is the same energy as the
difference between the eigenstates of the transition; however, the energy of fluorescent photons is always less than that of the
exciting photons. This difference is because energy is lost in internal conversion and vibrational relaxation, where it is
transferred away from the electron. Due to the large number of vibrational levels that can be coupled into the transition
between electronic states, measured emission is usually distributed over a range of wavelengths.

Figure 4: Possible scenario with absorption, internal conversion and vibrational relaxation, and fluorescence processes
shown.

Intersystem Crossing
Yet another path a molecule may take in the dissipation of energy is called intersystem crossing. This where the electron
changes spin multiplicity from an excited singlet state to an excited triplet state. It is indicated by a horizontal, curved arrow
from one column to another. This is the slowest process in the Jablonski diagram, several orders of magnitude slower than
fluorescence. This slow transition is a forbidden transition, that is, a transition that based strictly on electronic selection rules
should not happen. However, by coupling vibrational factors into the selection rules, the transition become weakly allowed and
able to compete with the time scale of fluorescence. Intersystem crossing leads to several interesting routes back to the ground
electronic state. One direct transition is phosphorescence, where a radiative transition from an excited triplet state to a singlet
ground state occurs.This is also a very slow, forbidden transition. Another possibility is delayed fluorescence, the transition
back to the first excited singlet level, leading to the emitting transition to the ground electronic state.

Figure 5: Possible scenario with absorption, internal conversion, vibrational relaxation, intersystem crossing, and
phosphorescence processes shown.
Other non-emitting transitions from excited state to ground state exist and account for the majority of molecules not exhibiting
fluorescence or phosphorescent behavior. One process is the energy transfer between molecules through molecular collisions
(e.g., external conversion). Another path is through quenching, energy transfer between molecules through overlap in
absorption and fluorescence spectra. These are non-emitting processes that will compete with fluorescence as the molecule
relaxes back down to the ground electronic state. In a Jablonski diagram, each of these processes are indicated with a curved
line going down to on the energy scale.

Time Scales
It is important to note that a Jablonski diagram shows what sorts of transitions that can possibly happen in a particular
molecule. Each of these possibilities is dependent on the time scales of each transition. The faster the transition, the more

7/4/2021 3 https://chem.libretexts.org/@go/page/1769
likely it is to happen as determined by selection rules. Therefore, understanding the time scales each process can happen is
imperative to understanding if the process may happen. Below is a table of average time scales for basic radiative and non-
radiative processes.
Table 1: Average timescales for radiative and non-radiative processes
Transition Time Scale Radiative Process?

Internal Conversion 10-14 - 10-11 s no

Vibrational Relaxation 10-14 - 10-11 s no


Absorption 10-15 s yes
Phosphorescence 10-4 - 10-1 s yes
Intersystem Crossing 10-8 - 10-3 s no
Fluorescence 10-9 - 10-7 s yes

Each process outlined above can be combined into a single Jablonski diagram for a particular molecule to give a overall
picture of possible results of perturbation of a molecule by light energy. Jablonski diagrams are used to easily visualize the
complex inner workings of how electrons change eigenstates in different conditions. Through this simple model, specific
quantum mechanical phenomena are easily communicated.

References
1. H. H. Jaffe and Albert L. Miller "The fates of electronic excitation energy" J. Chem. Educ., 1966, 43 (9), p 469
DOI:10.1021/ed043p469
2. E. B. Priestley and A. Haug "Phosphorescence Spectrum of Pure Crystalline Naphthalene" J. Chem. Phys. 49, 622 (1968),
DOI:10.1063/1.1670118

Contributors and Attributions


Jordan McEwen (UCD)

7/4/2021 4 https://chem.libretexts.org/@go/page/1769
Jablonski Diagram: Shockwave

Contributors and Attributions


Thomas Chasteen, Department of Chemistry, Sam Houston State University

6/29/2021 Jablonski Diagram.1 https://chem.libretexts.org/@go/page/1767


Metal to Ligand and Ligand to Metal Charge Transfer Bands
In the field of inorganic chemistry, color is commonly associated with d–d transitions. If this is the case, why is it that some
transition metal complexes show intense color in solution, but possess no d electrons? In transition metal complexes a change
in electron distribution between the metal and a ligand gives rise to charge transfer (CT) bands when performing Ultraviolet-
visible spectroscopy experiments. For complete understanding, a brief introduction to electron transfer reactions and Marcus-
Hush theory is necessary.

Outer Sphere Charge Transfer Reactions


Electron transfer reactions(charge transfer) fall into two categories:
Inner- sphere mechanisms– electron transfer occurs via a covalently bound bridging ligand.

Figure 1: Intermediate formed in the reaction between [Fe(CN)6]3- and [Co(CN)5]3-


Outer -sphere mechanisms– electron transfer occurs without a covalent linkage forming between reactants
2+ 3+ 3+ 2+
[M L6 ] + [M L6 ] → [M L6 ] + [M L6 ] (1)

Here, we focus on outer sphere mechanisms.


In a self-exchange reaction the reactant and product side of a reaction are the same. No chemical reaction takes place and only
an electron transfer is witnessed. This reductant-oxidant pair involved in the charge transfer is called the precursor complex.
The Franck-Condon approximation states that a molecular electronic transition occurs much faster than a molecular
vibration.
Let’s look at an example:
2+ 3+ 3+ 2+
[M L6 ] + [M L6 ] → [M L6 ] + [M L6 ] (2)

This process has a Franck-Condon restriction: Electron transfer can only take place when the M–L bond distances in the
ML(II) and ML(III) states are the same. This means that vibrationally excited states with equal bonds lengths must be formed
in order to allow electron transfer to occur. This would mean that the [ ML6 ] 2+ bonds must be compressed and [ML6] 3+
bonds must be elongated in order for the reaction to occur.
Self exchange rate constants vary, because the activation energy required to reach the vibrational states varies according to the
system. The greater the changes in bond length required to reach the precursor complex, the slower the rate of charge transfer.1

A Brief Introduction to Marcus-Hush Theory


Marcus-Hush theory relates kinetic and thermodynamic data for two self-exchange reactions with data for the cross-reaction
between the two self-exchange partners. This theory determines whether an outer sphere mechanism has taken place. This
theory is illustrated in the following reactions
Self exchange 1: [ ML6 ] 2+ + [ML6] 3+ → [ ML6 ] 3+ + [ ML6 ] 2+ ∆GO = 0
Self exchange 2: [ ML6 ] 2+ + [ML6] 3+ → [ ML6 ] 3+ + [ ML6 ] 2+ ∆GO = 0
Cross Reaction: [ ML6 ] 2+ + [ ML6] 3+ → [ ML6 ] 3+ + [ ML6 ] 2+
The Gibbs free energy of activation ∆GŦis represented by the following equation:
∓ ∓ ∓ ∓ ′
ΔG = Δw G + Δo G + Δs G + RT ln(k T /hZ) (3)

6/24/2021 1 https://chem.libretexts.org/@go/page/1773
T = temperature in K
R = molar gas constant
k’ = Boltzman constant
h = Plancks constant
Z = effective frequency collision in solution ~ 1011 dm3 mol-1 s-1
∆wGŦ = the energy associated with bringing the reactants together, includes the work done to counter any repulsion
∆0GŦ = energy associated with bond distance changes
∆s ∆GŦ= energy associated with the rearrangements taking place in the solvent spheres
ln ( k’T / hZ) = accounts for the energy lost in the formation of the encounter complex
The rate constant for the self-exchange is calculated using the following reaction

−Δ G /RT
k = κZe (4)

where κ is the transmission coefficient ~1


The Marcus-Hush equation is given by the following expression
1/2
k12 = (k11 k22 K12 f12 ) (5)

where:
2
(log K12 )
log f12 = (6)
k11 k22
4 log( 2
)
Z

Z is the collision frequency


k11 and ∆GŦ11 correspond to self exchange 1
k22 and ∆GŦ22 correspond to self exchange 2
k12 and ∆GŦ12 correspond to the cross-reaction
K12 = cross reaction equilibrium constant
∆GO12= standard Gibbs free energy of the reaction
The following equation is an approximate from of the Marcus-Hush equation:
log k12 ≈ 0.5 log k11 + 0.5 log log (7)

since f ≈1 and log f .


≈0

How is the Marcus-Hush equation used to determine if an outer sphere mechanism is taking place?
values of k11, k22, K12, and k12 are obtained experimentally
k11 and k22 are theoretically values
K 12is obtained from E cell

If an outer sphere mechanism is taking place the calculated values of k will match or agree with the experimental values. If
12

these values do not agree, this would indicate that another mechanism is taking place.1

The Laporte Selection Rule and Weak d–d Transitions


d- d transitions are forbidden by the Laporte selection rule.
Laporte Selection Rule: ∆ l = + 1
Laporte allowed transitions: a change in parity occurs i.e. s → p and p → d.
Laporte forbidden transitions: the parity remains unchanged i.e. p → p and d → d.
d-d transitions result in weak absorption bands and most d-block metal complexes display low intensity colors in solution
(exceptions d0 and d10complexes). The low intensity colors indicate that there is a low probability of a d-d transition occurring.
Ultraviolet-visible (UV/Vis) spectroscopy is the study of the transitions involved in the rearrangements of valence electrons. In
the field of inorganic chemistry, UV/Vis is usually associated with d – d transitions and colored transition metal complexes.
The color of the transition metal complex solution is dependent on: the metal, the metal oxidation state, and the number of
metal d-electrons. For example iron(II) complexes are green and iron(III) complexes are orange/brown.2

6/24/2021 2 https://chem.libretexts.org/@go/page/1773
Charge Transfer Bands
If color is dependent on d-d transitions, why is it that some transition metal complexes are intensely colored in solution but
possess no d electrons?

Figure 2: Fullerene oxides are intensely colored in solution, but possess no d electrons. Solutions from left to right: C60, C60O,
C60O, and C60O2. Fullerenes, nanometer-sized closed cage molecules, are comprised entirely of carbons arranged in hexagons
and pentagons. Fullerene oxides, with the formula C60On, have epoxide groups directly attached to the fullerene cage.
In transition metal complexes a change in electron distribution between the metal and a ligand give rise to charge transfer (CT)
bands.1 CT absorptions in the UV/Vis region are intense (ε values of 50,000 L mole-1 cm-1 or greater) and selection rule
allowed. The intensity of the color is due to the fact that there is a high probability of these transitions taking place. Selection
rule forbidden d-d transitions result in weak absorptions. For example octahedral complexes give ε values of 20 L mol-1 cm-1
or less.2 A charge transfer transition can be regarded as an internal oxidation-reduction process. 2

Ligand to Metal and Metal to Ligand Charge Transfer Bands


Ligands possess σ, σ*, π, π*, and nonbonding (n) molecular orbitals. If the ligand molecular orbitals are full, charge transfer
may occur from the ligand molecular orbitals to the empty or partially filled metal d-orbitals. The absorptions that arise from
this process are called ligand-to-metal charge-transfer bands (LMCT) (Figure 2).2 LMCT transitions result in intense bands.
Forbidden d-d transitions may also take place giving rise to weak absorptions. Ligand to metal charge transfer results in the
reduction of the metal.

Figure 3: Ligand to Metal Charge Transfer (LMCT ) involving an octahedral d complex.


6

If the metal is in a low oxidation state (electron rich) and the ligand possesses low-lying empty orbitals (e.g., C O or C N ) −

then a metal-to-ligand charge transfer (MLCT) transition may occur. LMCT transitions are common for coordination
compounds having π-acceptor ligands. Upon the absorption of light, electrons in the metal orbitals are excited to the ligand π*
orbitals.2 Figure 3 illustrates the metal to ligand charge transfer in a d5 octahedral complex. MLCT transitions result in intense
bands. Forbidden d – d transitions may also occur. This transition results in the oxidation of the metal.

Figure 4. Metal to Ligand Charge Transfer (MLCT) involving an octahedral d complex.


5

6/24/2021 3 https://chem.libretexts.org/@go/page/1773
Effect of Solvent Polarity on CT Spectra
*This effect only occurs if the species being studied is an ion pair*
The position of the CT band is reported as a transition energy and depends on the solvating ability of the solvent. A shift to
lower wavelength (higher frequency) is observed when the solvent has high solvating ability.
Polar solvent molecules align their dipole moments maximally or perpendicularly with the ground state or excited state
dipoles. If the ground state or excited state is polar an interaction will occur that will lower the energy of the ground state or
excited state by solvation. The effect of solvent polarity on CT spectra is illustrated in the following example.

Example 1
You are preparing a sample for a UV/Vis experiment and you decide to use a polar solvent. Is a shift in wavelength
observed when:
a) Both the ground state and the excited state are neutral
When both the ground state and the excited state are neutral a shift in wavelength is not observed. No change
occurs. Like dissolves like and a polar solvent won’t be able to align its dipole with a neutral ground and excited
state.
b) The excited state is polar, but the ground state is neutral
If the excited state is polar, but the ground state is neutral the solvent will only interact with the excited state. It will
align its dipole with the excited state and lower its energy by solvation. This interaction will lower the energy of the
polar excited state. (increase wavelength, decrease frequency, decrease energy)

c) The ground state and excited state is polar


If the ground state is polar the polar solvent will align its dipole moment with the ground state. Maximum
interaction will occur and the energy of the ground state will be lowered. (increased wavelength, lower frequency,
and lower energy) The dipole moment of the excited state would be perpendicular to the dipole moment of the
ground state, since the polar solvent dipole moment is aligned with the ground state. This interaction will raise the
energy of the polar excited state. (decrease wavelength, increase frequency, increase energy)

d) The ground state is polar and the excited state is neutral

6/24/2021 4 https://chem.libretexts.org/@go/page/1773
If the ground state is polar the polar solvent will align its dipole moment with the ground state. Maximum
interaction will occur and the energy of the ground state will be lowered. (increased wavelength, lower frequency,
and lower energy). If the excited state is neutral no change in energy will occur. Like dissolves like and a polar
solvent won’t be able to align its dipole with a neutral excited state. Overall you would expect an increase in energy
(Illustrated below), because the ground state is lower in energy (decrease wavelength, increase frequency, increase
energy).4

How to Identify Charge Transfer Bands


CT absorptions are selection rule allowed and result in intense (ε values of 50,000 L mole-1 cm-1 or greater) bands in the
UV/Vis region.2 Selection rule forbidden d-d transitions result in weak absorptions. For example octahedral complexes give ε
values of 20 L mol-1 cm-1 or less.2 CT bands are easily identified because they:
Are very intense, i.e. have a large extinction coefficient
Are normally broad
Display very strong absorptions that go above the absorption scale (dilute solutions must be used)

Example 2 : Ligand to Metal Charge Transfer


KMnO4 dissolved in water gives intense CT Bands. The one LMCT band in the visible is observed around 530 nm.

Figure 5: Absorption spectrum of an aqueous solution of potassium permanganate, showing a vibronic progression. (CC
BY-SA 3; Petergans via Wikipedia)
The band at 528 nm gives rise to the deep purple color of the solution. An electron from a “oxygen lone pair” character
orbital is transferred to a low lying Mn orbital.1

Example 3 : Metal to Ligand Charge Transfer


Tris(bipyridine)ruthenium(II) dichloride (\ce{[Ru(bpy)3]Cl2}\)) is a coordination compound that exhbits a CT band is
observed (Figure 6)

6/24/2021 5 https://chem.libretexts.org/@go/page/1773
Figure 6: a) Structure of [Ru(bpy)3]Cl2, b) CT band observed in its V/Vis spectrum. (CC BY-SA 4.0; Albris via
Wikipedia)
A d electron from the ruthenium atom is excited to a bipyridine anti-bonding orbital. The very broad absorption band is
due to the excitation of the electron to various vibrationally excited states of the π* electronic state.6

Practice Problems
1. You perform a UV/Vis on a sample. The sample being studied has the ability to undergo a charge transfer transition. A
charge transfer transitions is observed in the spectra. Why would this be an issue if you want to detect d-d transitions? How
can you solve this problem?
2. What if both types of charge transfer are possible? For example a complex has both σ-donor and π-accepting orbitals? Why
would this be an issue?
3. If the ligand has chromophore functional groups an intraligand band may be observed. Why would this cause a problem if
you want to observe charge transfer bands? How would you identify the intraligand bands? State a scenario in which you
wouldn’t be able to identify the intraligand bands.

Answers to Practice Problems


1. This is an issue when investigating weak d-d transitions, because if the molecule undergoes a charge transfer transitions it
results in an intense CT band. This makes the d – d transitions close to impossible to detect if they occur in the same region
as the charge transfer band. This problem is solved by performing the UV/Vis experiment on a more concentrated solution,
resulting in minor peaks becoming more prominent.
2. Octahedral complexes such as Cr(CO)6, have both σ-donor and π-accepting orbitals. This means that they are able to
undergo both types of charge transfer transitions. This makes it difficult to distinguish between LMCT and MLCT.
3. This would cause a problem because CT bands may overlap intraligand bands. Intraligand bands can be identified by
comparing the complex spectrum to the spectrum of the free ligand. This may be difficult, since upon coordination to the
metal, the ligand orbital energies may change, compared to the orbital energies of the free ligand. It would be very difficult
to identify an intraligand band if the ligand doesn’t exist as a free ligand. If it doesn’t exist as a free ligand you wouldn’t be
able to take a UV/Vis, and thus wouldn’t be able to use this spectrum in comparison to the complex spectrum.

Article References
1. Housecroft, Catherine E., and A. G. Sharpe. "Outer Sphere Mechanism." Inorganic Chemistry. Harlow, England: Pearson
Prentice Hall, 2008. 897-900.
2. Brisdon, Alan K. "UV-Visible Spectroscopy." Inorganic Spectroscopic Methods. Oxford: Oxford UP, 1998. 70-73..
3. Huheey, James E., Ellen A. Keiter, and Richard L. Keiter. "Coordination Chemistry: Bonding, Spectra, and Magnetism."
Inorganic Chemistry: Principles of Structure and Reactivity. New York, NY: HarperCollins College, 1993. 455-59.
4. Drago, Russell S. "Effect of Solvent Polarity on Charge-Transfer Spectra." Physical Methods for Chemists. Ft. Worth:
Saunders College Pub., 1992. 135-37.
5. Miessler, Gary L., and Donald A. Tarr. "Coordination Chemistry III: Electronic Spectra." Inorganic Chemistry. Upper
Saddle River, NJ: Pearson Education, 2004. 407-08.

6/24/2021 6 https://chem.libretexts.org/@go/page/1773
6. Paris, J. P., and Warren W. Brandt. "Charge Transfer Luminescence of a Ruthenium(II) Chelate." Communications to the
Editor 81 (1959): 5001-002.

Literature: Marcus Theory and Charge Transfer Bands


1. Marcus, R. A. "Chemical and Electrochemical Electron-Transfer Theory." Annual Review of Physical Chemistry 15.1
(1964): 155-96.
2. Eberson, Lennart. "Electron Transfer Reactions in Organic Chemistry. II* An Analysis of Alkyl Halide Reductions
by Electron Transfer Reagents on the Basis of Marcus Theory." Acta Chemica Scandinavica 36 (1982): 533-43.
3. Chou, Mei, Carol Creutz, and Norman Sutin. "Rate Constants and Activation Parameters for Outer-sphere Electron-
transfer Reactions and Comparisons with the Predictions of Marcus Theory." Journal of the American Chemical
Society 99.17 (1977): 5615-623.
4. Marcus, R. A. "Relation between Charge Transfer Absorption and Fluorescence Spectra and the Inverted Region."
Journal of Physical Chemistry 93 (1989): 3078-086.

Literature: Examples of Charge Transfer Bands


1. Electron Transfer Reactions of Fullerenes:
Mittal, J. P. "Excited States and Electron Transfer Reactions of Fullerenes." Pure and Applied Chemistry 67.1 (1995):
103-10.
Wang, Y. "Photophysical Properties of Fullerenes/ N,N-diethylanaline Charge Transfer Complexes." Journal of
Physical Chemistry 96 (1992): 764-67.
Vehmanen, Visa, Nicolai V. Tkachenko, Hiroshi Imahori, Shunichi Fukuzumi, and Helge Lemmetyinen. "Charge-
transfer Emission of Compact Porphyrin–fullerene Dyad Analyzed by Marcus Theory of Electron-transfer."
Spectrochimica Acta Part A 57 (2001): 2229-244.
2. Electron Transfer Reactions in RuBpy:
Paris, J. P., and Warren W. Brandt. "Charge Transfer Luminescence of a Ruthenium(II) Chelate." Communications to
the Editor 81 (1959): 5001-002.

Contributor
Melissa A. Rivera (UC Davis)

6/24/2021 7 https://chem.libretexts.org/@go/page/1773
Radiative Decay
Spontaneous emission is the process in which a quantum mechanical system (such as an atom, molecule or subatomic particle)
transitions from an excited energy state to a lower energy state (e.g., its ground state) and emits a quantum in the form of a
photon.

Topic hierarchy

Chemiluminescence
Most chemiluminescence methods involve only a few chemical components to actually generate light.

Fluorescence
Fluorescence, a type of luminescence, occurs in gas, liquid or solid chemical systems. Fluorescence is brought about by
absorption of photons in the singlet ground state promoted to a singlet excited state. The spin of the electron is still paired
with the ground state electron, unlike phosphorescence. As the excited molecule returns to ground state, it involves the
emission of a photon of lower energy, which corresponds to a longer wavelength, than the absorbed photon.

Phosphorescence
Unlike florescence, phosphorescence does not re-emit the light immediately. Instead, phosphorescence releases light very
slowly in the dark due to its energy transition state. When light such as ultraviolet light is shined upon a glow in dark
object, the object emits light, creating phosphorescence.

7/7/2021 1 https://chem.libretexts.org/@go/page/77807
Chemiluminescence
Most chemiluminescence methods involve only a few chemical components to actually generate light. Luminol
chemiluminescence (Nieman, 1989), which has been extensively investigated, and peroxyoxalate chemiluminescence (Given
and Schowen, 1989; Orosz et al., 1996) are both used in bioanalytical methods and will be the subject of this primer on
chemiluminescence. In each system, a "fuel" is chemically oxidized to produced an excited state product. In many luminol
methods it is this excited product that emits the light for the signal. In peroxyoxalate chemiluminescence, the initial excited
state product does not emit light at all and instead it reacts with another compound.

Contributors and Attributions


Dr. Thomas Chasteen (Sam Houston State University)

5/27/2021 1 https://chem.libretexts.org/@go/page/1760
Fluorescence
Fluorescence, a type of luminescence, occurs in gas, liquid or solid chemical systems. Fluorescence is brought about by
absorption of photons in the singlet ground state promoted to a singlet excited state. The spin of the electron is still paired with
the ground state electron, unlike phosphorescence. As the excited molecule returns to ground state, it involves the emission of
a photon of lower energy, which corresponds to a longer wavelength, than the absorbed photon.

Introduction
The energy loss is due to vibrational relaxation while in the excited state. Fluorescent bands center at wavelengths longer than
the resonance line. This shift toward longer wavelengths is called a Stokes shift. Excited states are short-lived with a lifetime
at about 10-8 seconds. Molecular structure and chemical environment affect whether or not a substance luminesces. When
luminescence does occur, molecular structure and chemical environment determine the intensity of emission. Generally
molecules that fluoresce are conjugated systems. Fluorescence occurs when an atom or molecules relaxes through vibrational
relaxation to its ground state after being electrically excited. The specific frequencies of excitation and emission are dependent
on the molecule or atom.

S0 + h νex = S1 (1)

where
hν is a photon energy with
h is Planck's constant and

ν is the frequency of light,

S is the ground state of the fluorophore and


0

S is its first electronically excited state.


1

Figure 1: Jablonski diagram of absorbance, non-radiative decay, and fluorescence. (Public Domain, Jacobkhed)
Figure 1 is a Jablonski energy diagram representing fluorescence. The purple arrow represents the absorption of light. The
green arrow represents vibrational relaxation from singlet excited state, S2 to S1. This process is a non-radiative relaxation in
which the excitation energy is dispersed as vibrations or heat to the solvent, and no photon is emitted. The yellow arrow
represents fluorescence to the singlet ground state, So.
The fluorescence quantum yield ((\Phi\)) gives the efficiency of the fluorescence process. It is the ratio of photons emitted to
photons absorbed.
 # emitted photons 
Φ = (2)
 # absorbed photons 

If every photon absorbed results in a photon emitted. The maximum fluorescence quantum yield is 1.0, and compounds with
quantum yields of 0.10 are still considered fluorescent. Another way to define the fluorescence quantum yield is by the excited
state decay rates:

6/9/2021 1 https://chem.libretexts.org/@go/page/1766
kf
Φ = (3)
∑ ki
i

where k is the rate of spontaneous emission of radiation and the denominator is the sum of all rates of excited state decay for
f

each deactivation process (ie phosphorescence, intersystem crossing, internal conversion…). The fluorescence lifetime is the
average time the molecule remains in its excited state before emitting a photon. Fluorescence typically follows first-order
kinetics:
−t/τ
[ S1 ] = [ S1 ]o e (4)

where
[ S1 ] is the concentration of excited state molecules at time t ,
[S1 ]0 is the initial concentration and τ is the decay rate.
Various radiative and non-radiative processes can de-populate the excited state so the total decay rate is the sum over all rates:
τtot = τrad + τnrad (5)

where τ is the total decay rate, τ the radiative decay rate and τ
tot rad the non-radiative decay rate. If the rate of spontaneous
nrad

emission or any of the other rates are fast, the lifetime is short. The average lifetime of fluorescent compounds that emit
photons with energies from the UV to near infrared are within the range of 0.5 to 20 nanoseconds.
The fluorescence intensity, I is proportional to the amount of light absorbed and the fluorescence quantum yield, Φ
F

−εbc
If = kIo ϕ[1 − (10 )] (6)

where
k is a proportionality constant attributed to the instrument
\(I_o\( is the incident light intensity
ϵ is the molar absorptivity,

b is the path length, and

c is the concentration of the substrate.

If dilute solutions are used so that less than 2% of the excitation energy is absorbed, then an approximation can be made so
that
x
10 ≈ 1 + x+. . . (7)

so Equation 6 can be simplified to


If = kIo Φ[εbc] (8)

This relationship shows that fluorescence intensity is proportional to concentration.


Fluorescence rarely results from absorption of UV-radiation of wavelengths shorter than 250 nm because this type of radiation
is sufficiently energetic to cause deactivation of the excited state by predissociation or dissociation. Most organic molecules
have at least some bonds that can be ruptured by energies of this strength. Consequently, fluorescence due to sigma → σ ∗

transitions is rarely observed. Instead such emission is confined to the less energetic π → π and π → n processes.
∗ ∗

Fluorescence commonly occurs from a transition from the lowest vibrational level of the first excited electronic state to the
one of the vibrational levels of the electronic ground state. Quantum yield (Φ) is greater for π → π transition because these

excited states show short average lifetimes (larger k ) and because deactivation processes that compete with fluorescence is
f

not as likely to happen. Molar absorptivity of π → π* transitions is 100-1000 fold greater. The average lifetime is 10-7 to 10-9
seconds for ?, ?* states and 10-5 to 10-7 seconds for n, π* states.

6/9/2021 2 https://chem.libretexts.org/@go/page/1766
Figure 2: Schematic representation of a fluorescence spectrometer. from OpenStax (CC-BY-3.0)
Figure 2 is a schematic of a typical filter fluorimeter that uses a source beam for fluorescence excitation and a pair of
photomultiplier tubes as transducers. The source beam is split near the source into a reference beam and a sample beam. The
reference beam is attenuated by the aperture disk so that its intensity is roughly the same as the fluorescence intensity. Both
beams pass through the primary filter, with the reference beam being reflected to the reference photomultiplier tube. The
sample beam is focused on the sample by a pair of lenses and causes fluorescence emission. The emitted radiation passes
through a second filter and then is focused on the sample photomultiplier tube. The electrical outputs from the two transducers
are then processed by an analog to digital converter to compute the ratio of the sample to reference intensities, which can then
be used for qualitative and quantitative analysis. To obtain an emission spectrum, the excitation monochromator is fixed and
the emission monochromator varies. To obtain an excitation spectrum, the excitation monochromator varies while the emission
monochromator is fixed.
Fluorescence spectroscopy can be used to measure the concentration of a compound because the fluorescence intensity is
linearly proportional to the concentration of the fluorescent molecule. Fluorescent molecules can also be used as tags. For
example, fluorescence in situ hybridization (FISH) is a method of determining what genes are present in an organism's
genome. Single stranded DNA encoding a gene of interest is covalently bonded to a fluorescent molecule and washed over the
organism's chromosome, binding to its complementary sequence. The presence and placement of the gene in the organism then
fluoresces when shined with ultraviolet light. Green fluorescence protein (GFP) is used in molecular biology to monitor the
activity of proteins. The gene encoding GFP can be inserted next to a gene encoding a protein that will be studied. When the
genes are expressed, the protein will be attached to GFP and can be identified in the cell by its fluorescence.

Contributors and Attributions


Zoe Smith, Christina Roman

6/9/2021 3 https://chem.libretexts.org/@go/page/1766
Phosphorescence
Unlike florescence, phosphorescence does not re-emit the light immediately. Instead, phosphorescence releases light very
slowly in the dark due to its energy transition state. When light such as ultraviolet light is shined upon a glow in dark object,
the object emits light, creating phosphorescence.

Contributors and Attributions


Chern-Yi Tsai (UCD)

6/18/2021 1 https://chem.libretexts.org/@go/page/1771
Selection Rules for Electronic Spectra of Transition Metal Complexes
The Selection Rules governing transitions between electronic energy levels of transition metal complexes are:
1. ΔS = 0 The Spin Rule
2. Δl = +/- 1 The Orbital Rule (or Laporte)
The first rule says that allowed transitions must involve the promotion of electrons without a change in their spin. The second
rule says that if the molecule has a center of symmetry, transitions within a given set of p or d orbitals (i.e. those which only
involve a redistribution of electrons within a given subshell) are forbidden.
Relaxation of these rules can occur through:
Spin-Orbit coupling: this gives rise to weak spin forbidden bands
Vibronic coupling: an octahedral complex may have allowed vibrations where the molecule is asymmetric.
Absorption of light at that moment is then possible.
Mixing: π-acceptor and π-donor ligands can mix with the d-orbitals so transitions are no longer purely d-d.

Transition Types
1. Charge transfer, either ligand to metal or metal to ligand. These are often extremely intense and are generally found in the
UV but they may have a tail into the visible.
2. d-d, these can occur in both the UV and visible region but since they are forbidden transitions have small intensities.
Expected intensities of electronic transitions
Transition type Example Typical values of ε /m2mol-1

Spin forbidden,
[Mn(H2O)6]2+ 0.1
Laporte forbidden
Spin allowed (octahedral complex),
[Ti(H2O)6]3+ 1 - 10
Laporte forbidden
Spin allowed (tetrahedral complex),
Laporte partially allowed [CoCl4]2- 50 - 150
by d-p mixing
Spin allowed,
Laporte allowed [TiCl6]2- or MnO4- 1,000 - 106
e.g. charge transfer bands

Expected Values
The expected values should be compared to the following rough guide.
For M2+ complexes, expect Δ = 7,500 - 12,500 cm-1 or λ = 800 - 1,350 nm.
For M3+ complexes, expect Δ= 14,000 - 25,000 cm-1 or λ = 400 - 720 nm.
For a typical spin-allowed, but Laporte (orbitally) forbidden transition in an octahedral complex, expect ε < 10 m2mol-1.
Extinction coefficients for tetrahedral complexes are expected to be around 50-100 times larger than for octrahedral
complexes. B for first-row transition metal free ions is around 1,000 cm-1. Depending on the position of the ligand in the
nephelauxetic series, this can be reduced to as low as 60% in the complex.

Contributors
Prof. Robert J. Lancashire (The Department of Chemistry, University of the West Indies)

7/6/2021 1 https://chem.libretexts.org/@go/page/1772
Derivation of Laporte Rule
The Laporte Rule is a selection rule in electron absorption spectroscopy that applies to centrosymmetric molecules. It says that
transitions between states of the same symmetry with respect to inversion are forbidden. Using the mathematical concept of even
and odd functions, the Laporte Rule can be derived and summarized as follows: Electronic transitions from waveunfunctions
with g symmetry to wavefunctions with g symmetry are forbidden, as are transitions from wavefunctions with u symmetry to
wavefunctions with u symmetry. Transitions from g to u and u to g, where the symmetry switches, may be (but are not
necessarily) allowed.

Introduction
When an incident photon is absorbed by an electron, the electron is excited from a lower energy state to a higher energy , excited
state. Although there are multiple higher energy excited states, not all are available to the electron, and only certain transitions are
allowed. Electron absorption spectroscopy measures the intensity of transmitted photons as a function of wavelength. From this
the energy of absorbed photons is determined, and transitions are assigned using selection rules. Selection rules are the guidelines
for determining if a transition will be allowed or forbidden. This module will further discuss the orbital selection rule called the
Laporte Rule, but will leave the spin selection rule to be discussed in another module. Specifically, the Laporte selection rule
determines whether an electonic transtion is orbitally allowed or forbidden. The Laporte Rule applies only to centrosymmetric
molecules, or those that contain a center of inversion. An electronic transition is forbidden by the Laporte Rule if the ground and
excited states have the same symmetry with respect to an inversion center. Electronic transitions are described by the transition
moment integral.

^ ex ∣ ^∣ ∣
∫ Ψel M Ψ dτ = ⟨i ∣
∣M ∣ f ⟩∣ (1)
el ∣
−∞

(Ψ is the wavefunction, M is the transition dipole moment operator, i refers to the initial state and f refers to the final state). The
Laporte Rule tells us that if the integrand of the transition moment integral does not contain the totally symmetric representation,
then the transition is forbidden.

Background
In electron absorption spectroscopy, absorption of a photon excites an electron from one energy state to another. Such transitions
are only detectable using optical spectroscopy when there is a corresponding change in the dipole moment of the molecule;
transitions are detectable when the transition moment dipole is nonzero.

ex
μ =∫ Ψμ
^Ψ (2)
−∞

The transition moment dipole given above includes the wave functions of the final and initial states, which contain both
electronic and nuclear wave functions.
Ψ = Ψnuc Ψel (3)

The Born–Oppenheimer approximation tells us that electronic transitions happen on a faster time scale than nuclear transitions so
the two can be separated, and when dealing with electronic transitions we can effectively ignore any nuclear motion. Thus, in
discussing electronic spectroscopy and the "allowedness" of certain electronic transitions, we consider only the electronic
integral.

ex
∫ Ψel H Ψ (4)
el
−∞

Where H is the time-dependent Hamiltonian. We can make this integral easier to work with by invoking Fermi’s Golden Rule,
which allows us to replace H with M, the electric dipole moment operator , which is part of the time-dependent Hamiltonian.

^
M = e ∑ ri (5)

We arrive at the following transition moment integral to describe electronic transitions from a ground state to an excited state.

6/29/2021 1 https://chem.libretexts.org/@go/page/1762

∫ ^ Ψex = ∣⟨i ∣M
Ψel M ^ ∣ f ⟩∣ (6)
el ∣ ∣ ∣ ∣
−∞

The transition moment integral describes the probability of a transition taking place. The integral must be non-zero in order for a
transition to occur. In other words, the two electronic states must overlap in order for a transition to occur.

Derivation
By the Laporte Rule in a centrosymmetric molecule, if the integrand of the transition moment integral contains the totally
symmetric representation, then the transition is Laporte allowed. If the transition does not contain the totally symmetric
representation, then the transition is Laporte forbidden. The totally symmetric representation is the irreducible symmetry
representation that is even with respect to all symmetry operations. On a character table, this representation is usually listed first
and has the designation A, A1, or A1g.
To determine if a transition is allowed, the point group of the molecule is determined, and the corresponding character table gives
the symmetry representation of each electronic wavefunction. The character table is also used to determine the electric dipole
moment operator, which is contained in the transition moment integral and has three parts that transform with the Cartesian x, y,
and z axes.
∣M
^ ∣
x
∣ ∣
^ ^ ∣
M =∣M (7)
y
∣ ∣
^ ∣
∣Mz

For calculating the integrand of the transition moment integral, the symmetry representations corresponding to x, y, and z are
used for the electric dipole moment operator. The integrand of the transition moment integral is then given by the following direct
products:
∣M
^ ∣
x


^ ∣ ⊗ Ψex
Ψel ⊗ ∣ M (8)
y el
∣ ∣
^ ∣
∣Mz

The component of the electric dipole moment operator that gives the totally symmetric representation corresponds to the
direction of polarization of the electronic transition.
For a point group of low symmetry like C2h, computing the integrand for all possible transitions is a relatively short task.
However, when considering a point group with a high degree of symmetry, like Oh, computing the integrand for all possible
transitions in a point group becomes a long and tedious task. If we examine these computations through the lens of mathematics,
this problem can be simplified. The Laporte Rule applies to centrosymmetric molecules, those containing a center of inversion.
All states have a symmetry with respect to the inversion center, and all representations for a given point group have either a g
(gerade) or u (ungerade) designation with respect to the center of inversion. Designations of g and u refer to even and odd
symmetries with respect to the inversion center. Odd and even with respect to orbital symmetry correspond to the mathematical
definitions of odd and even functions. M is an odd function. For the derivation of the Laporte Rule, allow g states to be
represented by the even function cos(x), u states to be represented by the odd function sin(x), and M to be represented by the odd
function x.
A transition between two even states is given by the integral

^ ex
F (x) = ∫ Ψel M Ψ =∫ cos x × x × cos xdx (9)
el
−∞

The integrand is given by


2
G(x) = x × cos x (10)

Evaluate the integrand


π π 2
π π
G(x = ) = × cos ( ) = (11)
4 4 4 8

Evaluate the integrand

6/29/2021 2 https://chem.libretexts.org/@go/page/1762
−π −π −π −π
2
G(x = ) = × cos ( ) = (12)
4 4 4 8

G(x) = −G(−x) (13)

Thus,
2
G(x) = x × cos x (14)

is an odd function, and the transition between two even states is forbidden.
Similarly, a transition between two odd states is given by the integral

^ ex
F (x) = ∫ Ψel M Ψ =∫ sin x × x × sin xdx (15)
el
−∞

The integrand is given by


2
G(x) = x × sin x (16)

Evaluate the integrand


π π π π
2
G(x = ) =( ) × sin ( ) = (17)
4 4 4 8

Evaluate the integrand


−π −π −π −π
2
G(x = ) =( ) × sin ( ) = (18)
4 4 4 8

G(x) = −G(−x) (19)

Thus,
2
G(x) = x × sin x (20)

is an odd function, and the transition between two odd states is forbidden.
A transition between an odd and an even state is given by the integral

F (x) = ∫ ^ Ψex = ∫
Ψel M cos x × x × sin xdx (21)
el
−∞

The integrand is given by


G(x) = cos x × x × sin x (22)

Evaluate the integrand


π π π π π
G(x = ) = sin ( )×( ) × cos ( ) = (23)
4 4 4 4 8

Evaluate the integrand


−π −π −π −π π
G(x = ) = sin ( )×( ) × cos ( ) = (24)
4 4 4 4 8

G(x) = G(−x) (25)

Thus,
G(x) = x × sin x × cos x (26)

is an even function, and the transition between an even and an odd state is not immediately forbidden.
These calculations can be summarized in the following set of rules:

gxg=g Forbidden

uxu=g Forbidden
gxu=u May be Allowed

6/29/2021 3 https://chem.libretexts.org/@go/page/1762
uxg=u May be Allowed

It is shown above that the Laporte Rule expressly states that transitions between two even states or two odd states are forbidden.
It should be emphasized, however, that the Laporte Rule does not state that transitions between an even and an odd state are
allowed. The integrand of the transition moment integral for the transition between an odd and an even state is even, but there are
multiple states with even inversion symmetry that are not the totally symmetric representation. Only an integrand that contains
the totally symmetric representation is allowed. To determine which of the even transitions are actually allowed, the cross product
of the representations of the initial and final electronic states with the electric dipole moment operator must be computed. All
resulting cross products will be even; only those which contain the totally symmetric representation will be allowed.

Example 1: Octohedral Group


Oh (octahedral) point group

Mx , My , Mz = T1u (27)

A1g ⊗ T1u ⊗ A1g = T1u (28) Equation (1)

A2u ⊗ T1u ⊗ A2u = T1u (29) Equation (2)

A2g ⊗ T1u ⊗ A1g = A1g (30) Equation (3)

A2u ⊗ T1u ⊗ T1g = A2g + Eg + T1g + T2g (31) Equation (4)

Equation (1) shows a forbidden even-even transition. Equation (2) shows a forbidden odd-odd transition. Equations (3) and
(4) demonstrate the subtlety of the Laporte Rule. Both are even transitions between an odd and an even state. However, (3)
contains the totally symmetric representation and is an allowed transition, while (4) does not contain the totally symmetric
representation and is thereby forbidden.

Problems
1. For the C4h point group, which transitions are forbidden by the Laporte Rule? Which transitions are orbitally allowed?
2. For the C2h point group, what are the polarizations of the Laporte allowed transitions?
3. Using the Laporte Rule, are transitions between d orbitals in a molecule of octahedral (Oh) symmetry allowed? Are transitions
between p orbitals allowed? Are transitions between p and d orbitals allowed?
4. [Cu(H2O)6]2+ and [Cu(NH3)4]2+ both appear blue in solution because of the presence of copper ions. However, the two
solutions are not identical. How would the appearance of these solutions differ? If given an unlabeled sample of each, how
could the two solutions be distinguished without collecting any spectra?

Answers
1. The C4h point group contains the representations Ag, Bg, Eg, Au, Bu, and Eu. Via the Laporte rule: g to g and u to u transitions
are orbitally forbidden. For g to u and u to g transitions, the integrand of the transition moment integral must contain the totally
symmetric representation, which is Ag in this case.

Ag ⊗ Ag (32) Laporte Forbidden

Ag ⊗ Bg (33) Laporte Forbidden

Ag ⊗ Eg (34) Laporte Forbidden

Ag ⊗ Au (35)

∣ Eu ∣ ∣ Ag ∣ Laporte Allowed
Ag ∣ ∣ Au = ∣ ∣ (36)
∣A ∣ ∣ Eu ∣
u

6/29/2021 4 https://chem.libretexts.org/@go/page/1762
Ag ⊗ Bu (37) Laporte Forbidden
∣ Eu ∣ ∣ Eg ∣
Ag ∣ ∣ Bu = ∣ ∣ (38)
∣ Au ∣ ∣ Bg ∣

Ag ⊗ Eu (39)

∣ Eu ∣ ∣ Ag + Au + 2 Bg ∣ Laporte Allowed
Ag ∣ ∣ Eu = ∣ ∣ (40)
∣ Au ∣ ∣ Eg ∣

Bg ⊗ Bg (41) Laporte Forbidden

Bg ⊗ Eg (42) Laporte Forbidden

Bg ⊗ Au (43)

∣ Eu ∣ ∣ Eg ∣ Laporte Forbidden
Bg ∣ ∣ Au = ∣ ∣ (44)
∣ Au ∣ ∣ Bg ∣

Bg ⊗ Bu (45)

∣ Eu ∣ ∣ Eg ∣ Laporte Forbidden
Bg ∣ ∣ Bu = ∣ ∣ (46)
∣ Au ∣ ∣ Bg ∣

Bg ⊗ Eu (47)

∣ Eu ∣ ∣ Ag + Au + 2 Bg ∣ Laporte Allowed
Bg ∣ ∣ Eu = ∣ ∣ (48)
∣ Au ∣ ∣ Eg ∣

Eg ⊗ Eg (49) Laporte Forbidden

Eg ⊗ Au (50)

∣ Eu ∣ ∣ Ag + Au + 2 Bg ∣ Laporte Allowed
Eg ∣ ∣ Au = ∣ ∣ (51)
∣ Au ∣ ∣ Eg ∣

Eg ⊗ Bu (52)

∣ Eu ∣ ∣ Bu + Bg + 2 Ag ∣ Laporte Allowed
Eg ∣ ∣ Bu = ∣ ∣ (53)
∣ Au ∣ ∣ Eg ∣

Eg ⊗ Eu (54)

∣ Eu ∣ ∣ Eg + Eu + 2 Eg ∣ Laporte Allowed
Eg ∣ ∣ Eu = ∣ ∣ (55)
∣ Au ∣ ∣ Ag + Au + 2 Bg ∣

Au ⊗ Au (56) Laporte Forbidden

Au ⊗ Bu (57) Laporte Forbidden

Au ⊗ Eu (58) Laporte Forbidden

Bu ⊗ Bu (59) Laporte Forbidden

Bu ⊗ Eu (60) Laporte Forbidden

Eu ⊗ Eu (61) Laporte Forbidden

6/29/2021 5 https://chem.libretexts.org/@go/page/1762
2. The C2h point group contains Ag, Bg, Au, and Burepresentations. By the Laporte rule g to g and u to u transitions are orbitally
forbidden. For g to u and u to g transitions, the integrand of the transition moment integral must contain the totally symmetric
representation, which is Ag in this case. The component of the transition moment operator that gives the totally symmetric
representation dictates the polarization of the transition.

∣A ∣ ∣A ∣
u u
∣ ∣ ∣ ∣
Ag Bu Ag = Bu (62) Laporte Forbidden
∣ ∣ ∣ ∣
∣ Bu ∣ ∣ Bu ∣

∣A ∣ ∣B ∣
u u
∣ ∣ ∣ ∣
Ag Bu Bg = Au (63) Laporte Forbidden
∣ ∣ ∣ ∣
∣ Bu ∣ ∣ Au ∣

∣A ∣ ∣A ∣
u g
∣ ∣
∣ ∣
Ag

Bu

Au = ∣ Bg ∣ (64) Laporte Allowed: x-polarized
∣ ∣
∣ Bg ∣ ∣ Bg ∣

∣A ∣ ∣B ∣
u g
∣ ∣
∣ ∣
Ag

Bu

Bu = ∣ Ag ∣ (65) Laporte Allowed: y-, z-polarized
∣ ∣
∣ Bg ∣ ∣ Ag ∣

∣A ∣ ∣ Au ∣
u
∣ ∣ ∣ ∣
Bg

Bu

Bg =

Bu

(66) Laporte Forbidden
∣ Bg ∣ ∣ Bu ∣

∣A ∣ ∣B ∣
u g
∣ ∣
∣ ∣
Bg

Bu

Au = ∣ Ag ∣ (67) Laporte Allowed: y-, z-polarized
∣ ∣
∣ Bg ∣ ∣ Ag ∣

∣A ∣ ∣B ∣
u g
∣ ∣
∣ ∣
Bg

Bu

Bu = ∣ Ag ∣ (68) Laporte Allowed: y-, z-polarized
∣ ∣
∣ Bg ∣ ∣ Ag ∣

∣A ∣ ∣ Au ∣
u
∣ ∣ ∣ ∣
Au

Bu

Au =

Bu

(69) Laporte Forbidden
∣ Bg ∣ ∣ Bu ∣

∣A ∣ ∣ Bu ∣
u
∣ ∣ ∣ ∣
Au

Bu

Bu =

Au

(70) Laporte Forbidden
∣ Bg ∣ ∣ Au ∣

∣A ∣ ∣ Au ∣
u
∣ ∣ ∣ ∣
Bu

Bu

Bu =

Bu

(71) Laporte Forbidden
∣ Bg ∣ ∣ Bu ∣

3. Using the character table, we see that d-orbitals in octahedral have the following symmetries:

2 2 2
z ,x −y (72) Eg (73)

xy, xz, yz (74) T2g (75)

Possible transitions are

6/29/2021 6 https://chem.libretexts.org/@go/page/1762
Eg × Eg (76)

Eg × T2g (77)

T2g × T2g (78)

All possible transitions are g to g, and are thus forbidden by the Laporte Rule.
P-orbitals have the following symmetries:

x, y, z (79) T1u (80)

All possible transitions are u to u, and are thus forbidden by the Laporte Rule.
P-orbital to d-orbital transitions would all be u to g, and are not forbidden by the Laporte Rule.
4. [Cu(NH3)4]2+ is a tetrahedral complex and is therefore non-centrosymmetric. Since it is non-centrosymmetric, it is not Laporte
forbidden. [Cu(H2O)6]2+ is an octahedral complex whose d-d transitions are Laporte forbidden. [Cu(NH3)4]2+ will be a darker
shade of blue in solution because its d-d transitions are not forbidden, and [Cu(H2O)6]2+ will be a paler shade of blue in solution
because its d-d transitions are Laporte forbidden. Coloration and absorption of [Cu(H2O)6]2+ in the ultraviolet-visible range is
attributed to vibronic coupling.

References
1. Harris, Daniel C, and Michael D. Bertolucci. Symmetry and Spectroscopy: An Introduction to Vibrational and Electronic
Spectroscopy. New York: Oxford University Press, 1978.
2. Laporte, O.; Meggers, W. F. J. Opt. Soc. Am. 1925, 11, 459. http://dx.doi.org/10.1364/JOSA.11.000459.
3. Rao, C N. R. Ultra-violet and Visible Spectroscopy: Chemical Applications. London: Butterworth, 1975.

Contributors and Attributions


Kathryn A. Newton

6/29/2021 7 https://chem.libretexts.org/@go/page/1762
Spin-orbit Coupling
Spin-orbit coupling refers to the interaction of a particle's "spin" motion with its "orbital" motion.

The Spin-orbit coupling Hamiltonian


The magnitude of spin-orbit coupling splitting is measured spectroscopically as
1
Hso = hcA ((l + s)(l + s + 1) − l(l + 1) − s(s + 1))
2

1
2 2 2 2
= hcA (l +s + ls + sl + l + s − l −l−s − s))
2

= hcAl ⋅ s

The expression can be modified by realizing that j = l + s .


1
Hso = hcA(j(j + 1) − l(l + 1) − s(s + 1)) (1)
2

where A is the magnitude of the spin-orbit coupling in wave numbers. The magnitude of the spin orbit coupling can be
calculated in terms of molecule parameters by the substitution
2
Zα 1
ˆ ˆ ˆ ˆ
hcA L ⋅ S = L⋅S (2)
3
2 r

where a is the fine structure constant (a = 1/137.037) and the carrots indicate that L and S are operators. The fine structure
2
e
constant is a dimensionless constant, a = . Z is an effective atomic number. The spin orbit coupling splitting can be
ác
calculated from
ˆ ˆ

Z ∗
L⋅S
Eso = ∫ Ψ HSO Ψ dτ = ∫ Ψ Ψ dτ (3)
2 3
2(137) r

This expression can be recast to give an spin-orbit coupling energy in terms of molecular parameters
1 Z 1
Eso = (j(j + 1) − l(l + 1) − s(s + 1)) = ⟨ ⟩ (4)
2 3
2 2(137) r

where
1 1

⟨ ⟩ =∫ Ψ ( )Ψ dτ (5)
3 3
r r

We can evaluate this integral explicitly for a given atomic orbital.


For example for Y210 we have
3

1 Z Zr
2 −Zr/2a0
Ψ210 = ( ) e cos θ (6)
−−
4 √2π a0 a0

so that the integral is


5 2z z ∞
1 1 Z 1
2 2 Zr/a0 2
⟨ ⟩ = ( ) ∫ dϕ ∫ cos θ sin θ dθcosθ ∫ r e ( )r dr (7)
3 3
r 32π a0 0 0 0 r

which integrates to

7/3/2021 1 https://chem.libretexts.org/@go/page/5558
5 2 3
1 1 Z 2 a0 1 Z
⟨ ⟩ = ( ) 2π( )( ) = ( ) (8)
r3 32π a0 3 Z2 24 a0

Or Z 3
/24 in atomic units.
Therefore in atomic units we have
3
1 Z
⟨ ⟩ = (9)
3 3
r n l(l + 1/2)(l + 1)

Therefore, in general the spin-orbit splitting is given by


4
Z j(j + 1) − l(l + 1) − s(s + 1)
Eso = ( ) (10)
2 3
2(137) n 2l(l + 1/2)(l + 1)

Note that the spin-orbit coupling increases as the fourth power of the effective nuclear charge Z, but only as the third power of
the principal quantum number n. This indicates that spin orbit-coupling interactions are significantly larger for atoms that are
further down a particular column of the periodic table.

Contributors and Attributions


Stefan Franzen (North Carolina State University)

7/3/2021 2 https://chem.libretexts.org/@go/page/5558
Atomic Term Symbols
In electronic spectroscopy, an atomic term symbol specifies a certain electronic state of an atom (usually a multi-electron one),
by briefing the quantum numbers for the angular momenta of that atom. The form of an atomic term symbol implies Russell-
Saunders coupling. Transitions between two different atomic states may be represented using their term symbols, to which
certain rules apply.

History
At the beginning, the spectroscopic notation for term symbols was derived from an obsolete system of categorizing spectral
lines. In 1885, Johann Balmer, a Swiss mathematician, discovered the Balmer formula for a series of hydrogen emission lines.
2
m
λ = B( ) (1)
2
m −4

where
B is constant, and
m is an integer greater than 2.
Later it was extended by Johannes Rydberg and Walter Ritz.
Yet this principle could hardly explain the discovery of fine structure, the splitting of spectral lines. In spectroscopy, spectral
lines of alkali metals used to be divided into categories: sharp, principal, diffuse and fundamental, based on their fine
structures. These categories, or "term series," then became associated with atomic energy levels along with the birth of the old
quantum theory. The initials of those categories were employed to mark the atomic orbitals with respect to their azimuthal
quantum numbers. The sequence of "s, p, d, f, g, h, i, k..." is known as the spectroscopic notation for atomic orbitals.
By introducing spin as a nature of electrons, the fine structure of alkali spectra became further understood. The term "spin"
was first used to describe the rotation of electrons. Later, although electrons have been proved unable to rotate, the word "spin"
is reserved and used to describe the property of an electron that involves its intrinsic magnetism. LS coupling was first
proposed by Henry Russell and Frederick Saunders in 1923. It perfectly explained the fine structures of hydrogen-like atomic
spectra. The format of term symbols was developed in the Russell-Saunders coupling scheme.

Term Symbols
In the Russell-Saunders coupling scheme, term symbols are in the form of 2S+1LJ, where S represents the total spin angular
momentum, L specifies the total orbital angular momentum, and J refers to the total angular momentum. In a term symbol, L is
always an upper-case from the sequence "s, p, d, f, g, h, i, k...", wherein the first four letters stand for sharp, principal, diffuse
and fundamental, and the rest follow in an alphabetical pattern. Note that the letter j is omitted.

Angular momenta of an electron


In today's physics, an electron in a spherically symmetric potential field can be described by four quantum numbers all
together, which applies to hydrogen-like atoms only. Yet other atoms may undergo trivial approximations in order to fit in this
description. Those quantum numbers each present a conserved property, such as the orbital angular momentum. They are
sufficient to distinguish a particular electron within one atom. The term "angular momentum" describes the phenomenon that
an electron distributes its position around the nucleus. Yet the underlying quantum mechanics is much more complicated than
mere mechanical movement.
The azimuthal angular momentum: The azimuthal angular momentum (or the orbital angular momentum when
describing an electron in an atom) specifies the azimuthal component of the total angular momentum for a particular
electron in an atom. The orbital quantum number, l, one of the four quantum numbers of an electron, has been used to
represent the azimuthal angular momentum. The value of l is an integer ranging from 0 to n-1, while n is the principal
quantum number of the electron. The limits of l came from the solutions of the Schrödinger Equation.
The intrinsic angular momentum: The intrinsic angular momentum (or the spin) represents the intrinsic property of
elementry particles, and the particles made of them. The inherent magnetic momentum of an electron may be explained by

6/30/2021 1 https://chem.libretexts.org/@go/page/1775
its intrinsic angular momentum. In LS-coupling, the spin of an electron can couple with its azimuthal angular momentum.
The spin quantum number of an electron has a value of either ½ or -½, which reflects the nature of the electron.

Coupling of the electronic angular momenta


Coupling of angular momenta was first introduced to explain the fine structures of atomic spectra. As for LS coupling, S, L, J
and MJ are the four "good" quantum numbers to describe electronic states in lighter atoms. For heavier atoms, jj coupling is
more applicable, where J, MJ, ML and Ms are "good" quantum numbers.

LS coupling
LS coupling, also known as Russell-Saunders coupling, assumes that the interaction between an electron's intrinsic angular
momentum s and its orbital angular momentum L is small enough to be considered as an perturbation to the electronic
Hamiltonian. Such interactions can be derived in a classical way. Let's suppose that the electron goes around the nucleus in a
circular orbit, as in Bohr model. Set the electron's velocity to be Ve. The electron experiences a magnetic field B due to the
relative movement of the nucleus
1 Ze
B = (E × p) = L (2)
me c2 4πϵ0 me c2 r3

while E is the electric field at the electron due to the nucleus, p the classical monumentum of the electron, and r the distance
between the electron and the nucleus.
The electron's spin s brings a magnetic dipole moment μs
−gse s
^
μs = (L ⋅ ^
s) (3)
2me

where gs is the gyromagnetic ratio of an electron. Since the potential energy of the coulumbic attraction between the electron
and nucleus is6
2
−Ze
V (r) = (4)
4π ε0 r

the interaction between μs and B is


2
gs Ze gs ∂V
^
H so = (L ⋅ s) = (L ⋅ s) (5)
2 2 2 2 2
2me c 4πϵ r 2me c ∂r
0

After a correction due to centripetal acceleration10, the interaction has a format of


1 ∂V
^
H so = (L ⋅ s) (6)
2 2
2μ c ∂r

Thus the coupling energy is


^
⟨ψnlm | H so | ψnlm ⟩ (7)

In lighter atoms, the coupling energy is low enough be treated as a first-order perturbation to the total electronic Hamiltonian,
hence LS coupling is applicable to them. For a single electron, the spin-orbit coupling angular momentum quantum number j
has the following possible values
j = |l-s|, ..., l+s
if the total angular momentum J is defined as J = L + s. The azimuthal counterpart of j is mj, which can be a whole number in
the range of [-j, j].
The first-order perturbation to the electronic energy can be deduced so6
2 ∞
ℏ [j(j + 1) − l(l + 1) − s(s + 1)] 1 ∂V
2 2
∫ r dr R (r) (8)
2 2 nl
4μ c 0
r ∂r

6/30/2021 2 https://chem.libretexts.org/@go/page/1775
Above is about the spin-orbit coupling of one electron. For many-electron atoms, the idea is similar. The coupling of angular
momenta is
J = L+S (9)

thereby the total angular quantum number


J = |L-S|, ..., L+S
where the total orbital quantum number

L = ∑ li (10)

and the total spin quantum number

S = ∑ si (11)

While J is still the total angular momentum, L and S are the total orbital angular momentum and the total spin, respectively.
The magnetic momentum due to J is
gJ eJ
μ =− (12)
J
2me

wherein the Landé g factor is


J(J + 1) + S(S + 1) − L(L + 1)
gJ = 1 + (13)
2J(J + 1)

supposing the gyromagnetic ratio of an electron is 2.

jj coupling
For heavier atoms, the coupling between the total angular momenta of different electrons is more significant, causing the fine
structures not to be "fine" any more. Therefore the coupling term can no more be considered as a perturbation to the electronic
Hamiltonian, so that jj coupling is a better way to quantize the electron energy states and levels.
For each electron, the quantum number j = l + s. For the whole atom, the total angular momentum quantum number

J = ∑ ji (14)

Term symbols for an Electron Configuration


Term symbols usually represent electronic states in the Russell-Saunders coupling scheme, where a typical atomic term
symbol consists of the spin multiplicity, the symmetry label and the total angular momentum of the atom. They have the
format of
2S+1
LJ (15)

such as 3D2, where S = 1, L = 2, and J = 2.


Here is a commonly used method to determine term symbols for an electron configuration. It requires a table of possibilities of
different "micro states," which happened to be called "Slater's table".6 Each row of the table represents a total magnetic
quantum number, while each column does a total spin. Using this table we can pick out the possible electronic states easily
since all terms are concentric rectangles on the table.
The method of using a table to count possible "microstates" has been developed so long ago and honed by so many scientists
and educators that it is hard to accredit a single person. Let's take the electronic configuration of d3 as an example. In the
Slater's table, each cell contains the number of ways to assign the three electrons quantum numbers according to the MS and
ML values. These assignments follow Pauli's exclusion law. The figure below shows an example to find out how many ways to
assign quantum numbers to d3 electrons when ML = 3 and MS = -1/2.

6/30/2021 3 https://chem.libretexts.org/@go/page/1775
MS
-3/2 -1/2 1/2 3/2
5 0 1 1 0
4 0 2 2 0
3 1 4 4 1
2 1 6 6 1
1 2 8 8 2
ML 0 2 8 8 2
-1 2 8 8 2
-2 1 6 6 1
-3 1 4 4 1
-4 0 2 2 0
-5 0 1 1 0

Now we start to subtract term symbols from this table. First there is a 2H state. And now it is subtracted from the table.

MS
-3/2 -1/2 1/2 3/2
5 0 0 0 0
4 0 1 1 0
3 1 3 3 1
2 1 5 5 1
1 2 7 7 2
ML 0 2 7 7 2
-1 2 7 7 2
-2 1 5 5 1
-3 1 3 3 1
-4 0 1 1 0
-5 0 0 0 0

And now is a 4F state. After being subtracted by 4F, the table becomes

MS
-3/2 -1/2 1/2 3/2
ML 5 0 0 0 0
4 0 1 1 0
3 0 2 2 0
2 0 4 4 0
1 1 6 6 1

6/30/2021 4 https://chem.libretexts.org/@go/page/1775
0 1 6 6 1
-1 1 6 6 1
-2 0 4 4 0
-3 0 2 2 0
-4 0 1 1 0
-5 0 0 0 0

And now 2G.


MS
-3/2 -1/2 1/2 3/2
5 0 0 0 0
4 0 0 0 0
3 0 1 1 0
2 0 3 3 0
1 1 5 5 1
ML 0 1 5 5 1
-1 1 5 5 1
-2 0 3 3 0
-3 0 1 1 0
-4 0 0 0 0
-5 0 0 0 0

Now 2F.
MS
-3/2 -1/2 1/2 3/2
5 0 0 0 0
4 0 0 0 0
3 0 0 0 0
2 0 2 2 0
1 1 4 4 1
ML 0 1 4 4 1
-1 1 4 4 1
-2 0 2 2 0
-3 0 0 0 0
-4 0 0 0 0
-5 0 0 0 0

Here in the table are two 2D states.


MS
-3/2 -1/2 1/2 3/2
ML 5 0 0 0 0
4 0 0 0 0
3 0 0 0 0

6/30/2021 5 https://chem.libretexts.org/@go/page/1775
2 0 0 0 0
1 1 2 2 1
0 1 2 2 1
-1 1 2 2 1
-2 0 0 0 0
-3 0 0 0 0
-4 0 0 0 0
-5 0 0 0 0

4
P.
MS
-3/2 -1/2 1/2 3/2
5 0 0 0 0
4 0 0 0 0
3 0 0 0 0
2 0 0 0 0
1 0 1 1 0
ML 0 0 1 1 0
-1 0 1 1 0
-2 0 0 0 0
-3 0 0 0 0
-4 0 0 0 0
-5 0 0 0 0

And the final deducted state is 2P. So in total the possible states for a d3 configuration are 4F, 4P, 2H, 2G, 2F, 2D, 2D and 2P.
Taken J into consideration, the possible states are:
4 4 4 4 4 4 2 2 2 2 2 2 2 2 2 2 2 2
F2 F3 F4 P0 P1 P2 H 9 H 11 G 9 G 7 F 7 F 5 D 5 D 5 D 3 D 3 P 3 P 3 (16)
2 2 2 2 2 2 2 2 2 2 2 2

For lighter atoms before or among the first-row transition metals, this method works well.

Using group theory to determine term symbols


Another method is to use direct products in group theory to quickly work out possible term symbols for a certain electronic
configuration. Basically, both electrons and holes are taken into consideration, which naturally results in the same term
symbols for complementary configurations like p2 vs p4. Electrons are categorized by spin, therefore divided into two
categories, α and β, as are holes: α stands for +1, and βstands for -1, or vice versa. Term symbols of different possible
configurations within one category are given. The term symbols for the total electronic configuration are derived from direct
products of term symbols for different categories of electrons. For the p3 configuration, for example, the possible combinations
of different categories are eα3, eβ3, eα2eβ and eαeβ2. The first two combinations were assigned the partial term of S.6 As eα2 and
eβ were given an P symbol, the combination of them gives their direct product
P × P = S + [P] + D.
The direct product for eα and eβ2 is also
P × P = S + [P] + D.
Considering the degeneracy, eventually the term symbols for p3 configuration are 4S, 2D and 2P.
There is a specially modified version of this method for atoms with 2 unpaired electrons. The only step gives the direct product
of the symmetries of the two orbitals. The degeneracy is still determined by Pauli's exclusion.

6/30/2021 6 https://chem.libretexts.org/@go/page/1775
Determining the ground state
In general, states with a greater degeneracy have a lower energy. For one configuration, the level with the largest S, which has
the largest spin degeneracy, has the lowest energy. If two levels have the same S value, then the one with the larger L (and also
the larger orbital degeneracy) have the lower energy. If the electrons in the subshell are fewer than half-filled, the ground state
should have the smallest value of J, otherwise the ground state has the greatest value of J.

Electronic transitions
Electrons of an atom may undergo certain transitions which may have strong or weak intensities. There are rules about which
transitions should be strong and which should be weak. Usually an electronic transition is excited by heat or radiation.
Electronic states can be interpreted by solutions of Schrödinger's equation. Those solutions have certain symmetries, which are
a factor of whether transitions will be allowed or not. The transition may be triggered by an electric dipole momentum, a
magnetic dipole momentum, and so on. These triggers are transition operators. The most common and usually most intense
transitions occur in an electric dipolar field, so the selection rules are
1. ΔL = 0, ±1 except L = 0 ‡ L' = 0
2. ΔS = 0
3. ΔJ = 0; ±1 except J = 0 ‡ J' = 0
where a double dagger means not combinable. For jj coupling, only the third rule applies with an addition rule: Δj = 0; ±1.

References
1. W. Ritz (1908), "On a New Law of Series Spectra", Astrophysical Journal 28(10): 237. doi: 10.1086/141591
2. §7.12, Jeremy B. Tatum, Stellar Atomsphere, online book. Accessed on line February 6, 2011.
3. Herzberg, Gerhard (1945). Atomic Spectra and Atomic Structure. New York: Dover. pp. 54–5. ISBN 0-486-60115-3.
4. Levine, Ira (2000). Quantum Chemistry (5 ed.). Prentice Hall. pp. 144–145. ISBN 0-13-685512-1
5. George Kean Sweetnam, The Command of Light - google books. P 182.
6. W. S. Struve (1989), Fundamentals of Molecular Spectroscopy, John Wiley & Sons, Inc. pp. 44-46, 61-62. ISBN 0-471-
85424-7
7. Niels Bohr (1913). "On the Constitution of Atoms and Molecules, Part I". Philosophical Magazine 26: 1–24
8. M. H. Nayfeh, M. K (1985). Brussel, Electricity and Magnetism, Wiley, New York. ISBN 0-471-87681-X
9. E. Merzbacher (1966), Quantum Mechanics, Wiley, New York.
10. W. H. Furry (1955), "Lorentz transformation and the Thomas precession", American Journal of Physics 23(8); 517-525.
doi: 10.1119/1.1934085
11. Darl H. McDaniel, "Spin factoring as an aid in the determination of spectroscopic terms", Journal of Chemical Education
54(3):147 (1977). doi: 10.1021/ed054p147
12. Fermi, Enrico (1926). "Sulla quantizzazione del gas perfetto monoatomico" (in Italian). Rend. Lincei 3: 145–9., translated
as On the Quantization of the Monoatomic Ideal Gas. 1999-12-14.

Problems
1. Potassium has the electronic configuration of [Ar]4s1, what are the possible term symbols of a neutral K atom?
2. What is the ground state of Cr2+? Specify the value of J.
3. Why does not the spin selection rule apply for electronic transitions under jj coupling?
4. Why do different term symbols appear as rectangles on the Slater's table?
5. Tanabe-Sugano diagrams tell us the order of energy levels in a complex, usually the order of d electron levels. Look up for
the possible term symbols for the d2 configuration, and determine how many fundamentals (transition from the ground
state to an excited state) may occur when Δ = 0 i.e. minor field due to ligands.

Answers
1. 2S.
2. 5D.
3. Under jj coupling the spin quantum number is not a "good" quantum number any more, which means it cannot describe
and differentiate electronic states properly.

6/30/2021 7 https://chem.libretexts.org/@go/page/1775
4. A certain term symbol represents a certain value of L as well as of S, which implies limited possible values of ML and Ms.
Following Pauli's exclusion rule, there must be only one possible way to assign the configuration a certain value of ML and
Ms.
5. No appropriate answer.

Contributors and Attributions


Anda Zheng (UCD)

6/30/2021 8 https://chem.libretexts.org/@go/page/1775
Molecular Term Symbols
Molecular term symbols specify molecular electronic energy levels. Term symbols for diatomic molecules are based on
irreducible representations in linear symmetry groups, derived from spectroscopic notations. They usually consist of four
parts: spin multiplicity, azimuthal angular momentum, total angular momentum and symmetry. All molecular term symbols
discussed here are based on Russel-Saunders coupling.

Introduction
Molecular term symbols mark different electronic energy levels of a diatomic molecule. These symbols are similar to atomic
term symbols, since both follow the Russell-Saunders coupling scheme. Molecular term symbols employ symmetry labels
from group theory. The possibility of an electronic transition can be deducted from molecular term symbols following
selection rules. For multi-atomic molecules, symmetry labels play most of term symbols' roles.
For homonuclear diatomics, the term symbol has the following form:
2S+1 (+/−)
Λ (1)
Ω,(g/u)

whereas Λ is the projection of the orbital angular momentum along the internuclear axis: Ω is the projection of the total
angular momentum along the internuclear axis; g/u is the parity; and +/− is the reflection symmetry along an arbitrary plane
containing the internuclear axis. Λ may be one of the greek letters in the sequence: Σ Π Δ Φ... when Λ = 0, 1, 2, 3...,
respectively. For heteronuclear diatomics, the term symbol does not include the g/u part, for there is not inversion center in the
molecule.

Determining term symbols of diatomics


Let's start with CO again. As we have seen before, the molecule has a close-shell configuration. Its ground state is a totally
symmetric singlet, 1Σ+, since the only possible values of (S, Λ) are (0, 0). If one of the HOMO electrons on the 5σ+ orbital has
jumped to the LUMO, this molecule will be in an excited state as follows.

Suppose a CO molecule is in the excited state shown above. In order to know the term symbol of this state, a direct product of
the labels is required for the two MO's with unpaired electrons. The multiplication is such as Π × Σ = Π . According to
+

Pauli's exclusion rule, these two unpaired electrons can never share the same set of quantum numbers, therefore the spin
degeneracy S can reach its maximum 3. The resulting term symbols are 1Π and 3Π.
Now if we look at O2, it does not have a close-shell configuration at its ground state. There are two unpaired electrons each
occupying one of the two degenerate 2π orbitals, which can be seen in the diagram below.

7/7/2021 1 https://chem.libretexts.org/@go/page/1770
The term symbol for oxygen molecule at its ground state is therefore derived such as Π x Π = Σ+ + Σ- + [Δ], as the symbol in
brackets does not allow the oxygen atoms to commute.

Transition between electronic states of diatomics


We'll focus on selection rules. Like atomic electronic states, different selection rules apply when differently incurred
transitions occur. Usually for electric dipole field induced transitions, the selection rules are the same as for atoms.
1. ΔΛ = 0, ±1 except Λ = 0 ‡ Λ' = 0
2. ΔS = 0
3. ΔΩ = 0; ±1 except Ω = 0 ‡ Ω' = 0

References
1. Harris, Daniel C. (republished in 1989). Symmetry and Spectroscopy: An Introduction to Vibrational and Electronic
Spectroscopy. Dover Publications. pp. 421-478. ISBN 0-486-66144-X
2. D. J. Willock (2009). Molecular Symmetry. John Wiley & Sons Ltd. ISBN 0-470-85348-4

Contributors and Attributions


Anda Zheng (UCD) and Sudarson S Sinha (TAU)

7/7/2021 2 https://chem.libretexts.org/@go/page/1770
The Russell Saunders Coupling Scheme
Review of Quantum Numbers
Electrons in an atom reside in shells characterized by a particular value of n, the Principal Quantum Number. Within each shell
an electron can occupy an orbital which is further characterized by an Orbital Quantum Number, l, where l can take all values
in the range:
l = 0, 1, 2, 3, . . . , (n − 1), (1)

traditionally termed s, p, d, f, etc. orbitals.


Each orbital has a characteristic shape reflecting the motion of the electron in that particular orbital, this motion being
characterized by an angular momentum that reflects the angular velocity of the electron moving in its orbital. A quantum
mechanics approach to determining the energy of electrons in an element or ion is based on the results obtained by solving the
Schrödinger Wave Equation for the H-atom. The various solutions for the different energy states are characterized by the three
quantum numbers, n, l and ml.
ml is a subset of l, where the allowable values are: ml = l, l-1, l-2, ..... 1, 0, -1, ....... , -(l-2), -(l-1), -l.
There are thus (2l +1) values of ml for each l value, i.e. one s orbital (l = 0), three p orbitals (l = 1), five d orbitals (l = 2),
etc.
There is a fourth quantum number, ms, that identifies the orientation of the spin of one electron relative to those of other
electrons in the system. A single electron in free space has a fundamental property associated with it called spin, arising
from the spinning of an asymmetrical charge distribution about its own axis. Like an electron moving in its orbital around a
nucleus, the electron spinning about its axis has associated with its motion a well defined angular momentum. The value of
ms is either + ½ or - ½.
In summary then, each electron in an orbital is characterized by four quantum numbers (Table 1).
Table 1: Quantum Numbers
symbol description range of values

Principal Quantum Number - largely governs


n 1,2,3 etc
size of orbital and its energy
Azimuthal/Orbital Quantum Number - largely
l determines shape of subshell (0 ≤ l ≤ n-1) for n = 3 then l = 0, 1, 2 (s, p, d)
0 for s orbital, 1 for p orbital etc
Magnetic Quantum Number - orientation of
ml l ≥ ml ≥ -l for l = 2, then ml = 2, 1, 0, -1, -2
subshell's shape for example px with py and pz
ms Spin Quantum Number either + ½ or - ½ for single electron

Russell Saunders coupling


The ways in which the angular momenta associated with the orbital and spin motions in many-electron-atoms can be combined
together are many and varied. In spite of this seeming complexity, the results are frequently readily determined for simple
atom systems and are used to characterize the electronic states of atoms. The interactions that can occur are of three types.
spin-spin coupling
orbit-orbit coupling
spin-orbit coupling
There are two principal coupling schemes used:
Russell-Saunders (or L - S) coupling
and jj coupling.
In the Russell Saunders scheme (named after Henry Norris Russell, 1877-1957 a Princeton Astronomer and Frederick Albert
Saunders, 1875-1963 a Harvard Physicist and published in Astrophysics Journal, 61, 38, 1925) it is assumed that:

7/2/2021 1 https://chem.libretexts.org/@go/page/7432
spin-spin coupling > orbit-orbit coupling > spin-orbit coupling.
This is found to give a good approximation for first row transition series where spin-orbit (J) coupling can generally be
ignored, however for elements with atomic number greater than thirty, spin-orbit coupling becomes more significant and the j-j
coupling scheme is used.

Spin-Spin Coupling
S - the resultant spin quantum number for a system of electrons. The overall spin S arises from adding the individual ms
together and is as a result of coupling of spin quantum numbers for the separate electrons.

Orbit-Orbit Coupling
L - the total orbital angular momentum quantum number defines the energy state for a system of electrons. These states or
term letters are represented as follows:
Total Orbital Momentum
L 0 1 2 3 4 5

S P D F G H

Spin-Orbit Coupling
Coupling occurs between the resultant spin and orbital momenta of an electron which gives rise to J the total angular
momentum quantum number. Multiplicity occurs when several levels are close together and is given by the formula (2S+1).
The Russell Saunders term symbol that results from these considerations is given by:
(2S+1)
L (2)

Configuration
S= + ½, hence (2S+1) = 2

L=2 and the Ground Term is written as 2D


The Russell Saunders term symbols for the other free ion configurations are given in the Table below.
Terms for 3dn free ion configurations
Configuration # of quantum states # of energy levels Ground Term Excited Terms
2D
d1,d9 10 1 -

d2,d8 45 5 3F 3P, 1G,1D,1S

d3,d7 120 8 4F 4P, 2H, 2G, 2F, 2 x 2D, 2P


3H, 3G, 2 x 3F, 3D, 2 x 3P,
d4,d6 210 16 5D 1I, 2 x 1G, 1F, 2 x 1D, 2 x
1S

4G, 4F, 4D, 4P, 2I, 2H, 2x


d5 252 16 6S
2G, 2 x 2F, 3 x 2D, 2P, 2S

Note that dn gives the same terms as d10-n

Hund's Rules
The Ground Terms are deduced by using Hund's Rules. The two rules are:
1. The Ground Term will have the maximum multiplicity
2. If there is more than 1 Term with maximum multipicity, then the Ground Term will have the largest value of L.
A simple graphical method for determining just the ground term alone for the free-ions uses a "fill in the boxes" arrangement.

dn 2 1 0 -1 -2 L S Ground Term
2D
d1 ↑ 2 1/2

7/2/2021 2 https://chem.libretexts.org/@go/page/7432
dn 2 1 0 -1 -2 L S Ground Term

d2 ↑ ↑ 3 1 3F

d3 ↑ ↑ ↑ 3 3/2 4F

d4 ↑ ↑ ↑ ↑ 2 2 5D

d5 ↑ ↑ ↑ ↑ ↑ 0 5/2 6S

d6 ↑↓ ↑ ↑ ↑ ↑ 2 2 5D

d7 ↑↓ ↑↓ ↑ ↑ ↑ 3 3/2 4F

d8 ↑↓ ↑↓ ↑↓ ↑ ↑ 3 1 3F

d9 ↑↓ ↑↓ ↑↓ ↑↓ ↑ 2 1/2 2D

To calculate S, simply sum the unpaired electrons using a value of ½ for each. To calculate L, use the labels for each column
to determine the value of L for that box, then add all the individual box values together.

Configuration
For a d7 configuration, then:
in the +2 box are 2 electrons, so L for that box is 2*2= 4
in the +1 box are 2 electrons, so L for that box is 1*2= 2
in the 0 box is 1 electron, L is 0
in the -1 box is 1 electron, L is -1*1= -1
in the -2 box is 1 electron, L is -2*1= -2
Total value of L is therefore +4 +2 +0 -1 -2 or L=3.
Note that for 5 electrons with 1 electron in each box then the total value of L is 0. This is why L for a d1 configuration is
the same as for a d6.

The other thing to note is the idea of the "hole" approach. A d1 configuration can be treated as similar to a d9 configuration. In
the first case there is 1 electron and in the latter there is an absence of an electron i.e., a hole.
The overall result shown in the Table above is that:
4 configurations (d1, d4, d6, d9) give rise to D ground terms,
4 configurations (d2, d3, d7, d8) give rise to F ground terms
and the d5 configuration gives an S ground term.

The Crystal Field Splitting of Russell-Saunders terms


The effect of a crystal field on the different orbitals (s, p, d, etc.) will result in splitting into subsets of different energies,
depending on whether they are in an octahedral or tetrahedral environment. The magnitude of the d orbital splitting is
generally represented as a fraction of Δoct or 10Dq.
The ground term energies for free ions are also affected by the influence of a crystal field and an analogy is made between
orbitals and ground terms that are related due to the angular parts of their electron distribution. The effect of a crystal field on
different orbitals in an octahedral field environment will cause the d orbitals to split to give t2g and eg subsets and the D ground
term states into T2g and Eg, (where upper case is used to denote states and lower case orbitals). f orbitals are split to give
subsets known as t1g, t2g and a2g. By analogy, the F ground term when split by a crystal field will give states known as T1g,
T2g, and A2g.
Note that it is important to recognize that the F ground term here refers to states arising from d orbitals and not f orbitals and
depending on whether it is in an octahedral or tetrahedral environment the lowest term can be either A2g or T1g.
The Crystal Field Splitting of Russell-Saunders terms
in high spin octahedral crystal fields.
Russell-Saunders Terms Crystal Field Components

7/2/2021 3 https://chem.libretexts.org/@go/page/7432
Russell-Saunders Terms Crystal Field Components

S (1) A1g

P (3) T1g
D (5) Eg , T2g
F (7) A2g , T1g , T2g
G (9) A1g , Eg , T1g , T2g
H (11) Eg , 2 x T1g , T2g
I (13) A1g , A2g , Eg , T1g , 2 x T2g

Note that, for simplicity, spin multiplicities are not included in the table since they remain the same for each term. The table
above shows that the Mulliken symmetry labels, developed for atomic and molecular orbitals, have been applied to these states
but for this purpose they are written in CAPITAL LETTERS.
Mulliken Symbols
Mulliken Symbol
Explanation
for atomic and molecular orbitals

a Non-degenerate orbital; symmetric to principal Cn

b Non-degenerate orbital; unsymmetric to principal Cn


e Doubly degenerate orbital
t Triply degenerate orbital
(subscript) g Symmetric with respect to center of inversion
(subscript) u Unsymmetric with respect to center of inversion
(subscript) 1 Symmetric with respect to C2 perp. to principal Cn
(subscript) 2 Unsymmetric with respect to C2 perp. to principal Cn
(superscript) ' Symmetric with respect to sh
(superscript) " Unsymmetric with respect to sh

For splitting in a tetrahedral crystal field the components are similar, except that the symmetry label g (gerade) is absent. The
ground term for first-row transition metal ions is either D, F or S which in high spin octahedral fields gives rise to A, E or T
states. This means that the states are either non-degenerate, doubly degenerate or triply degenerate.

Reference
H. N. Russell and F. A. Saunders, New Regularities in the Spectra of the Alkaline Earths, Astrophysical Journal, vol. 61, p.
38 (1925)

Contributors and Attributions


Prof. Robert J. Lancashire (The Department of Chemistry, University of the West Indies)

7/2/2021 4 https://chem.libretexts.org/@go/page/7432
Two-photon absorption
Two-photon absorption is one of a variety of two-photon processes. In this specific process, two photons are absorbed by a sample
simultaneously. Neither photon is at resonance with the available energy states of the system, however, the combined frequency of the
photons is at resonance with an energy state.

Introduction
Two-photon absorption may seem very similar to non-resonance Raman, and was in fact predicted due to this phenomena. In both
cases, a non-resonant photon is used for excitation. However, in the absorption case a secondary non-resonant photon is used for
excitation as well, while in Raman a second non-resonant photon is emitted. For Raman this results in occupation of an energy state at
the difference of the frequencies of the absorbed and emitted photon. However, in two-photon absorption this results in the occupation
of an energy state at the sum of the frequencies of the absorbed photons. The basic process is illustrated below in Figure 1.

Figure 1. Schematic of a two-photon absorption process. Both photons that are absorbed are done so simultaneously, not
sequentially. They constructively interfere in order to gain access to a higher lying eigenstate.
Two-photon absorption is not a feature of a specific type of spectroscopy. It can be utilized in any type of spectroscopy including IR,
NMR, XAS, and UV-VIS. However, it will result in a change of selection rules for many of these spectroscopies. This means that the
peak shapes should remain the same, but the intensities should be significantly smaller due to the result of the second order
perturbation.

History
In her 1931 dissertation, Maria Goeppert-Mayer postulated the existence of two-photon absorption for the first time. She says that she
based the derivation of this phenomena off of Kramer and Heisenberg's derivation of the probability of two-photon emission which
uses Dirac's dispersion theory for calculation.
The invention of the laser had a great impact on the field of two-photon spectroscopy because of the necessity of a high-intensity
electromagnetic field to induce transitions. Although, lasers were not formally invented until 1969, Bell Labs was testing masers in
1958, which were only capable of short pulses of intense electromagnetic radiation. In 1961 Kaiser and Garret reported the first two-
photon absorption of a compound. They used the new laser technology to excite CaF Eu with both red and blue light to induce a
2
2+

two photon transition.

General Perturbation Method


The following constants, variables, and operators are used in this section:

ℏ Reduced Planck's constant = 1.058x10-34 Js

e Elementary electric charge = 1.602x10-19 Coulombs


t time
ω angular frequency
2
^
p
^
H0
2m

p
^ −iℏ∇

First we must take a look at how the atomic system interacts with any perturbation
^ ^ ^
H = H0 + H1 (1)

If we are only going to look at the effects of the electric dipole then:

6/18/2021 1 https://chem.libretexts.org/@go/page/1776
^ ^ ^
H = H0 + Hμ (2)
e

Now we must consider how our states will be evolving in time because the Electric Dipole operator is time-dependent. According to
the time-dependent Schrödinger equation:
^
i H (t − t0 )
Ψi (t) = exp (− ) Ψi (t0 ) (3)

We have described our ground state as it is moving in time so now we can look at our final state which we can only describe as being
at a set energy value:
^
H 0 Ψf = ℏωf Ψf (4)

Now that we have described both the final and the initial states.
We can find the probability of such a transition by finding the magnitude of the projection of the intial state upon the final state,
otherwise known as the magnitude of the inner product.
2
| ⟨Ψf | Ψi (t)⟩ | = (5)

2
∣ ∣ ^ ∣ ∣
i H (t − t0 )
∣ ∣ ∣ ∣
⟨ Ψf exp (− ) Ψi (t0 )⟩ (6)
∣ ∣ ∣ ∣

∣ ∣ ∣ ∣

Now that we've found the general formula for the probability of our transitions in a time-dependent system, we must apply the
perturbation from earlier. To do this we use the following identity:
^ ^ ^ ^
iH 0 t iH t d iH 0 t iH t
^
exp ( ) H μ exp (− ) = iℏ [exp ( ) exp (− )] (7)
e
ℏ ℏ dt ℏ ℏ

The issue now is that the electric dipole Hamiltonian is not in the exponential form, this prevents us from simply solving for the total
Hamiltonian exponential. To solve this we integrate with respect to time, which nullifies the derivative.
t ^ ^
iH 0 t iH t
^
∫ exp ( ) H μ exp (− ) dt1 (8)
e

t0
ℏ ℏ

^ ^ ^ ^
iH 0 t iH t iH 0 t0 iH t0
= iℏ [exp ( ) exp (− ) − exp ( ) exp (− )] (9)
ℏ ℏ ℏ ℏ

This equation can be simplified and solved for using the approximation that the initial time was well before the incident of interaction
with the perturbation. This allows us to assume that the electric dipole Hamiltonian would yield 0 at the initial time. This allows for
the following iterative equation.
^ ^ t ^ ^
−i H t iH 0 t i iH 0 t iH t
^
exp ( ) = exp (− ) × [1 − ∫ exp ( ) H μ exp(ϵt1 )exp (− ) dt1 ] (10)
e
ℏ ℏ ℏ −∞
ℏ ℏ

Now for the purposes of two-photon absorption we will need to find the 2nd order perturbation. Thus we must find the second
iteration of this equation. Where epsilon is a very small number, we come to the following equation for both the first and second order
contribution:

^ ∣ ^
⎛ ⟨Ψf ∣ ∣ ∣
∣H μe ∣ Ψl ⟩ ⟨Ψl ∣H μe ∣ Ψi ⟩ ⎞
1 exp(−i ωi t) 1
^∣ ^
⟨Ψf ∣
∣H ∣ Ψi ⟩ = ⎜⟨Ψf ∣ ∣
∣H μe ∣ Ψi ⟩ + ∑ ⎟ (11)
ℏ ωi − ωf ℏ ωi − ωl
⎝ l ⎠

Electric Dipole Contribution


The following constants, variables and operators are introduced in this section:

ξ Amplitude of Electric field


^
μ Dipole operator = e ∑ N

i=1
qi r
^i

6/18/2021 2 https://chem.libretexts.org/@go/page/1776
For the perturbation method described in the previous section the use of a single iteration accurately describes the transitions between
a two levels. This describes the results of Fermi's Golden Rule. However, in order to describe these two-photon processes which
transgress through more than one transition, we must look at the second iteration within the perturbation.
^ ^
^ (E ^
Hμ = eμ 1 + E2 ) (12)
e

Where
^
En = ξn cos(ωn t) (13)

Thus
^ ^ (ξ1 cos(ω1 t) + ξ2 cos(ω2 t))
Hμ = eμ (14)
e

Substituting this into the second iteration gives us:

^∣
⟨Ψf ∣
∣H ∣ Ψi ⟩ = (15)

⎛ ⟨Ψf ∣ ^ ^ ^ ∣ ∣ ^ ^ ^ ∣
1 exp(−i ωi t) 1 ∣eμ(E 1 + E 2 )∣ Ψl ⟩ ⟨Ψl ∣eμ(E 1 + E 2 )∣ Ψi ⟩ ⎞

⎜⟨Ψf ∣ ^ ^ ^ ∣
∣eμ(E 1 + E 2 )∣ Ψi ⟩ + ∑ ⎟ (16)
ℏ ωi − ωf ℏ ωi − ωl
⎝ l ⎠

Looking at this equation we see that the first term is that corresponding to single photon transitions. However, because we are looking
for two-photon transitions we can ignore this term and focus on the second term. This second term has a summation to distinguish
between the different permutations of arranging the perturbations and achieve the same final state but it includes the interaction of
two photons from the same light source which we do not need to consider at the moment. If we eliminate the terms of the sum
responsible for these guys we should find the expression for the two permutations of two-photon transitions from our two light
sources. Now let us implement these changes.

∣ ^∣
⟨Ψf ∣H ∣ Ψi ⟩ = (17)

1 exp(−i ωi t) ⟨Ψf |eμ


^ξ1 cos(ω1 t)| Ψl ⟩ ⟨Ψl |eμ
^ξ2 cos(ω2 t)| Ψi ⟩ + ⟨Ψf |eμ
^ξ2 cos(ω2 t)| Ψl ⟩ ⟨Ψl |eμ
^ξ1 cos(ω1 t)| Ψi ⟩
( ) (18)
2
ℏ ωi − ωf ωi − ωl

This equation describes the electric dipole contribution towards two-photon transitions.

Selection Rules
The following constants, variables and operators are introduced in this section:

l Angular Momentum Quantum Number

ml Magnetic Quantum Number


Y Spherical Harmonic
^
Z Position operator = z
^
Pz Momentum Operator in the Z direction −iℏ d

dz

In order to take a thorough look at the selection rules for the two-photon system induced by the electric field contribution we must
look qualitatively at the results of the equation for the probability of a transition. The Hamiltonian presented in the previous section
for the electric dipole contribution is the result of a gauge transformation of the original electric dipole contribution. For the purposes
of finding the electric dipole selection rules we shall revert back to the original form.
qE
^ ^
H μ (t) = P z sin ωt (19)
e

We will consider the effect of the sinusoid part of this equation later in this section. For now let us track the selection rules due only to
the operator P^ . z

^ ^
∂H 0 Pz
^ ^
[Z , H 0 ] = iℏ = iℏ (20)
^ m
∂P z

This can be shown to be true by simply working out the commutator and leads to the following statement

6/18/2021 3 https://chem.libretexts.org/@go/page/1776
∣ ^ ^ ∣ ^ ^ ^ ^∣
⟨Ψf [Z , H 0 ] Ψi ⟩ = ⟨Ψf ∣
∣Z H 0 − H 0 Z ∣ Ψi ⟩ (21)
∣ ∣

iℏ
^∣ ^ ∣
= −(Ef − Ei ) ⟨Ψf ∣
∣Z ∣ Ψi ⟩ = ⟨Ψf ∣
∣P z ∣ Ψi ⟩ (22)
m

This shows the result of an electric dipole induced transition for a two level system. However, we need to slightly change things in
order to incorporate the three levels that will exist as a result of the two-photon transition as well as the fact that we will have two
different perturbations. Thus the Hamiltonian for the two-photon system will be:
qE1 qE2
^ ^ ^
H μ (t) = P z sin ω1 t + P z sin ω2 t (23)
e
mω1 mω2

Similar to before we can ignore the sinusoid part for later in this section. If we do this we instead come to the result:

^∣ ∣ ^∣
(Ef − El )(El − Ei ) ⟨Ψf ∣
∣Z ∣ Ψl ⟩ ⟨Ψl ∣Z ∣ Ψi ⟩ (24)

Whereas the momentum operator is difficult to find the selection rules for, the position operator is fairly straight forward. This is
because the variable z , which is the result of the operator Z^ , is equal to the following wavefunction.
−−


0
z =√ rY (θ) (25)
1
3

This works under the assumption that the light is polarized in the z direction but we can also polarize the light in multiple directions
as there are two photons. If this is the case we need to expand to define both x and y in terms of an angular function. We'll only
examine the x case here because it gives the same results.
−−


−1 1
x =√ r (Y −Y ) (26)
1 1
3

While for our case we have two position operators acting on different kets, however, that has no effect on the ket because
multiplication is commutative. This means the angular part of the inner product will result in the integral for the double z and double
x case respectively. Because all of the results retain the same l values that will not be addressed until after we first look at the m

values.

mf ∗ 0 ml ml ∗ 0 mi
∫ dΩ Y (θ, ϕ)Y (θ)Y (θ, ϕ)Y (θ, ϕ)Y (θ)Y (θ, ϕ) (27)
lf 1 ll ll 1 li

mf ∗ −1 ml ml ∗ −1 mi
1 1
∫ dΩ Y (θ, ϕ) (Y −Y )Y (θ, ϕ)Y (θ, ϕ) (Y −Y )Y (θ, ϕ) (28)
lf 1 1 ll ll 1 1 li

The intermediate or virtual state angular functions have no effect since they are complex conjugates of each other, their l and m

values will annihilate each other. Thus the integral we need to be concerned with is:

mf ∗ 0 0 mi
∫ dΩ Y (θ, ϕ)Y (θ)Y (θ)Y (θ, ϕ) ) (29)
lf 1 1 li

mf ∗ −1 1 −1 1 mi
∫ dΩ Y (θ, ϕ) (Y −Y ) (Y −Y )Y (θ, ϕ) (30)
lf 1 1 1 1 li

Because the basis to the selection rule is that only a sum of m values that results in 0 will result in an even function to integrate over,
we can now determine a range for m values. For purely Z polarized two-photon absorption the sum of the m values without the
initial and final states is still 0, thus the Δm must remain 0. However, for the inclusion of Z and X or Y . the sum without the initial
and final states will be ±1. Thus in order to retain a total sum of 0 the Δm must be ±1. Finally for the pure X, pure Y , and XY
cases the sum from the operators can results in ±2 or 0. Therefore, the Δm = ±2, 0 to negate the operators.
Now to talk about the l values. The selection rule that governs them is that we need the sum of all the l values to be even. Complex
conjugate l values are negative, the same holds true for the m values. Thus because each operator is has an l = 1 and we now have
two operators thus we can have the following possible changes to result in even sums.

Δl = 0, ±2 (31)

Now the rationality behind the change in the selection rules will be discussed. Each photon has a unit of momentum with it. During a
single photon excitation that momentum must be transferred due to conservation of momentum. Thus if it destructively interferes with
the electron it can lower the total momentum by one unit and if it constructively interferes with the electron is can increase the total

6/18/2021 4 https://chem.libretexts.org/@go/page/1776
momentum by one unit. However, for the two photon case, we now have to consider not only the ways in which the photon and
electron will interfere but also how the photons will interfere with each other. This results in a doubly constructive interference state
with an increase of 2 units, one constructive and one destructive interference state that results in no change and a doubly destructive
interference state that results in a decrease of 2 units.
Inversion Selection Rules
If we take a look at the equation presented at the end of the electric dipole section, it is a simple analysis to derive the basic selection
rules for electric dipole induced two-photon transitions. However, we do not need to consider the constant values to determine
selection rules as they only determine the intensity of the signal, not the existence of the peak. Where C is a constant:

^∣
⟨Ψf ∣
∣H ∣ Ψi ⟩ = (32)

^ ξ1 cos(ω1 t)| Ψl ⟩ ⟨Ψl |eμ


C (⟨Ψf |eμ ^ ξ2 cos(ω2 t)| Ψi ⟩ + ⟨Ψf |eμ
^ ξ2 cos(ω2 t)| Ψl ⟩ ⟨Ψl |eμ
^ ξ1 cos(ω1 t)| Ψi ⟩) (33)

Although for one-photon transitions the dipole moment is of considerable importance. We notice that because both terms contain two
dipole moments, regardless of whether they are odd or even, their combination will still become even. Furthermore, we can see that
because the wave part of the electric field is an even function, that this too will preserve parity of the transition in both terms, even if
it were an odd function as each term contains two of such functions. Thus the result is that as long as the final and initial state
conserve parity (inversion symmetry or antisymmetry is maintained) the transition will be allowed. This is different from one-photon
processes where parity conservation is forbidden and makes two-photon spectroscopy especially useful.

Magnetic Dipole Contribution


The following constants, variables, and operators are introduced in this section:

^
Lx Angular momentum Operator iℏ (sin ϕ ∂

∂θ
+ cot θcos ϕ

∂ϕ
)

^
Sx Spin momentum Operator ℏ

2
σx

σx Pauli Matrix
B Strength of Magnetic Field

The magnetic dipole contribution is determined by the Magnetic dipole Hamiltonian:


q
^ ^ ^
H DM = − (Lx + 2 Sx )Bcos(ωt) (34)
2m

where B is the strength of the magnetic field. So for the two photon case we have:
q
^ ^ ^ ^ ^
H DM = − [(Lx + 2 Sx )B1 cos(ω1 t) + (Lx + 2 Sx )B2 cos(ω2 t)] (35)
2m

Be aware that the photons do not need to be from the same polarization as they are here. One could be polarized in the X direction
while the other is polarized in the Z direction. We will cover these situations in the derivation of the selection rules.
If we similarly substitute this Hamiltonian into the second iteration for perturbation we get:

1 exp(−i ωi t) ∣ q ∣
^∣ ^ ^
⟨Ψf ∣
∣H ∣ Ψi ⟩ = ( ⟨Ψf − (Lx + 2 Sx ) [ B1 cos(ω1 t) + B2 cos(ω2 t)] Ψi ⟩ + (36)
ℏ ωi − ωf ∣ 2m ∣

q q
∣ ^ ^ ∣ ∣ ^ ^ ∣
⟨Ψf − (Lx + 2 S x ) [B1 cos(ω1 t) + B2 cos(ω2 t)] Ψl ⟩ ⟨Ψl − (Lx + 2 S x ) [B1 cos(ω1 t) + B2 cos(ω2 t)] Ψi ⟩
∣ 2m ∣ ∣ 2m ∣
∑ ) (37)
ℏ(ωi − ωl )
l

We can ignore the first term of this equation as it corresponds to the magnetic dipole contribution of the single photon transitions.
From the summation we can ignore the squared terms as well because they represent the two photon transitions due to the beam
interfering with itself. While that does exist and may be relevant for some spectroscopic work, here we are looking at the two-photon
absorption from two light sources.

^∣
⟨Ψf ∣
∣H ∣ Ψi ⟩ = (38)

∣ ^ ^ ∣ ∣ ^ ^ ∣ ∣ ^ ^ ∣ ∣ ^ ^ ∣
2 ⎛ ⟨Ψf (Lx + 2 S x ) Ψl ⟩ ⟨Ψl (Lx + 2 S x ) Ψi ⟩ + ⟨Ψf (Lx + 2 S x ) Ψl ⟩ ⟨Ψl (Lx + 2 S x ) Ψi ⟩ ⎞
q cos (ω1 t)cos (ω2 t) exp(−iωi t) ∣ ∣ ∣ ∣ ∣ ∣ ∣ ∣
⎜ ⎟ (39)
2
2
4m ℏ ωi − ωf ωi − ωl
⎝ ⎠

6/18/2021 5 https://chem.libretexts.org/@go/page/1776
Selection Rules
Just as before, our selection rules stem from the magnetic dipole Hamiltonian:
q
^ ^ ^
H DM = − (Lx + 2 Sx )Bcos(ωt) (40)
2m

Here, as before, the sinusoid part is even and thus will have no effect on the selection rules. We must look at the operators to
determine the selection rules.
Because neither L^
or S^ have any effect on the value of l the selection rule for Δl is 0. However both L
x x
^
xand S^ can change their
x

corresponding m and m values by one unit. For m this corresponds to changing the spin of the electron. If the magnetic field is
l s s

perpendicular to the spin of the system, then the Δm can also have a value of 0. If we extrapolate this to a two photon system then
s

the same m values should still apply given that even if it is changed by 1 by the first photon, it can change back by -1 or remain in
s

the + state as those are the only two options. Therefore, we should expect the same selection rules for Δm = 0, ±1 in a two-
1

2
s

photon system.

Modern Research Interests


Two-photon absorption has traditionally been a large part of the spectroscopy field. However, it was only within the past decade that
biomedical applications have returned it to the spotlight. Two-photon absorption, and potentially all nonlinear absorptions, allow for
unprecedented depth in medical imaging technology. Traditional single photon methods result in an enormous amount of scattering
from the biological tissue samples. However, these nonlinear optics allow for the assignment of the scattered photons to their origins.
Much of the research in this area is focused on maximizing the depth and clarity of the signal for this potentially non-intrusive
medical procedure. Several parameters have been looked at including excitation wavelength, beam size, pulse width, and pulse
frequency. As of now the largest depths achievable are around 1 mm.

References
1. Cohen-Tannoudji, Claude & Diu, Bernard & Laloe, Franck; 2005; Quantum Mechanics; Paris, France; Hermann
2. Eberly, Joseph & Lambropoulos, Peter; 1977; Multiphoton Processes; Rochester, New York; John Wiley & Sons
3. Garrett, C. G. B., Laboratories, B. T., & Hill, M. 1961. Two-Photon Excitation in CaF Eu , 7(6), 229–232.
2
2+

4. Helmchen and Denk, “Deep Tissue Two-photon Microscopy.”, Nature Methods, 2005, 2, 932-940
5. Loudon, Rodney; (1983); The quantum theory of light; New York; Oxford Publishing
6. Masters, Barry R.; English translation of: Göppert, M. (1929). Über die Wahrscheinlichkeit des Zusammenwirkens zweier
Lichtquanten in einem Elementarakt. (2010)., 1929, 273–294.
7. Tsukerblat, Boris; Group Theory in Chemistry and Spectroscopy: A Simple Guide to Advanced Usage; 2006; Mineola, New York;
Dover Publications

Problems
1. Derive the expression for the quadrupole contribution to the probability of a two-photon transition.
2. Find the third iteration of the Hamiltonian from the Perturbation adjustment using the iterative equation.
3. Use this third iteration to find the quantum mechanical selection rules for the electric dipole contribution to a three-photon
process.

Contributors and Attributions


Clifton Wagner

6/18/2021 6 https://chem.libretexts.org/@go/page/1776
CHAPTER OVERVIEW
PHOTOELECTRON SPECTROSCOPY
Photoelectron spectroscopy involves the measurement of kinetic energy of photoelectrons to
determine the bonding energy,intensity and angular distributions of these electrons and use the
information obtained to examine the electronic structure of molecules.

PHOTOELECTRON SPECTROSCOPY: APPLICATION


Photoelectron spectroscopy (PES) is a technique used for determining the ionization potentials of
molecules. Underneath the banner of PES are two separate techniques for quantitative and
qualitative measurements. They are ultraviolet photoeclectron spectroscopy (UPS) and X-ray
photoelectron spectroscopy (XPS). XPS is also known under its former name of electron
spectroscopy for chemical analysis (ESCA). UPS focuses on inoization of valence electrons, while
XPS involves ionizing core electrons.

PHOTOELECTRON SPECTROSCOPY: THEORY


Photoelectron spectroscopy involves the measurement of kinetic energy of photoelectrons to determine the binding energy, intensity
and angular distributions of these electrons and use the information obtained to examine the electronic structure of molecules. It
differs from the conventional methods of spectroscopy in that it detects electrons rather than photons to study electronic structures of
a material.

1 7/7/2021
Photoelectron Spectroscopy: Application
Photoelectron spectroscopy (PES) is a technique used for determining the ionization potentials of molecules. Underneath the
banner of PES are two separate techniques for quantitative and qualitative measurements. They are ultraviolet photoeclectron
spectroscopy (UPS) and X-ray photoelectron spectroscopy (XPS). XPS is also known under its former name of electron
spectroscopy for chemical analysis (ESCA). UPS focuses on inoization of valence electrons while XPS is able to go a step
further and ionize core electrons and pry them away.

Photoelectron Instrumentation
The main goal in either UPS or XPS is to gain information about the composition, electronic state, chemical state, binding
energy, and more of the surface region of solids. The key point in PES is that a lot of qualitative and quantitative information
can be learned about the surface region of solids. Specifics about what can be studied using XPS or UPS will be discussed in
detail below in separate sections for each technique following a discussion on instrumentation for PES experiments. The focus
here will be on how the instrumentation for PES is constructed and what types of systems are studied using XPS and UPS. The
goal is to understand how to go about constructing or diagramming a PES instrument, how to choose an appropriate analyzer
for a given system, and when to use either XPS or UPS to study a system.
There are a few basics common to both techniques that must always be present in the instrumental setup.
1. A radiation source: The radiation sources used in PES are fixed-energy radiation sources. XPS sources from x-rays while
UPS sources from a gas discharge lamp.
2. An analyzer: PES analyzers are various types of electron energy analyzers
3. A high vacuum environment: PES is rather picky when it comes to keeping the surface of the sample clean and keeping
the rest of the environment free of interferences from things like gas molecules. The high vacuum is almost always an ultra
high vacuum (UHV) environment.

Figure 1. Diagram of a basic, typical PES instrument used in XPS, where the radiation source is an X-ray source. When the
sample is irradiated, the released photoelectrons pass through the lens system which slows them down before they enter the
energy analyzer. The analyzer shown is a spherical deflection analyzer which the photoelectrons pass through before they are
collected at the collector slit.

Radiation sources
While many components of instruments used in PES are common to both UPS and XPS, the radiation sources are one area of
distinct differentiation. The radiation source for UPS is a gas discharge lamp, with the typical one being an He discharge lamp
operating at 58.4 nm which corresponds to 21.2 eV of kinetic energy. XPS has a choice between a monocrhomatic beam of a
few microns or an unfocused non-monochromatic beam of a couple centimeters. These beams originate from X-Ray sources of
either Mg or Al K-? sources giving off 1486 eV and 1258 eV of kinetic energy respectively. For a more versitile light source,
synchrotron radiation sources are also used. Synchrotron radiation is especially useful in studying valence levels as it provides
continuous, polarized radiation with high energies of > 350 eV.
The main thing to consider when choosing a radiation source is the kinetic energy involved. The source is what sets the kinetic
energy of the photoelectrons, so there needs to not only be enough energy present to cause the ionizations, but there must also
be an analyzer capable of measuring the kinetic energy of the released photoelectrons.

6/8/2021 Photoelectron Spectroscopy.1 https://chem.libretexts.org/@go/page/1836


In XPS experiments, electron guns can also be used in conjunction with x-rays to eject photoelectrons. There are a couple of
advantages and disadvantages to doing this, however. With an electron gun, the electron beam is easily focused and the
excitation of photoelectrons can be constantly varied. Unfortunately, the background radiation is increased significantly due to
the scattering of falling electrons. Also, a good portion of substances that are of any experimental interest are actually
decomposed by heavy electron bombardment such as that coming from an electron gun.

Analyzers
There are two main classes of analyzers well-suited for PES - kinetic energy analyzers and deflection or electrostatic
analyzers. Kinetic energy analyzers have a resolving power of E/δE, which means the higher the kinetic energy of the
photoelectrons, the lower the resolution of the spectra. Deflection analyzers are able to separate out photoelectrons through an
electric field by forcing electrons to follow different paths according to their velocities, giving a resolving power, E/δE, that
is greater than 1,000.
Since the resolving power of both types of analyzer is E/δE, the resolution is directly dependent on the kinetic energy of the
photoelectrons. The intensity of the spectra produced is also dependent on the kinetic energy. The faster the electrons are
moving, the lower the resolution and intensity is. In order to actually get well resolved, useful data other components must be
introduced into the instrument.
Adding a system of optics (lenses) to a PES instrument helps with this problem immensely. Electron optics are capable of
decelerating the photoelectrons through retardation of the electric field. The energy the photoelectrons decelerate to is known
as the "pass energy." This has the benefit of significantly raising the resolution, however this does, unfortunately, lower the
sensitivity. Optics are also capable of accelerating the electrons as well. The design of any lens system greatly effects the
photoelectron counts. These lenses are also capable of focusing on a small area of a particular sample.
Specific Analyzers
Within the broad picture of two main analyzer classes, there are a variety of specific analyzers in existence that are used in
PES. The list below goes over several well-used analyzers, though this list is, by no means, exhaustive. The most common
type of analyzer is a hemispherical analyzer, which will be explained in more depth under the spherical deflection analyzer
topic.
Plane Mirror Analyzer (PMA)
PMAs, the simplest type of electric analyzer are also known as parallel-plate mirror analyzers. These analyzers are condensers
made from two parallel plates with a distance, d, across them. Parabolic trajectories of electrons are obtained due to the
constant potential difference, V, between the two plates.

Figure 2. Schematic of a PMA where the angle between the bottom plate and the electrons entering is 45 degrees and the angle
between the bottom plate and the electrons exiting is also 45 degrees.
In order for transmission to occur, the potential must be: V=Eod/eLo. Eo=kinetic energy of electron in eV and e=charge of the
electron. To obtain better focus, the electron entrance and exit angle is capable of being shifted to 30 degrees, but this is not
necessarily a good idea as it sacrifices transmission instead.
Cylindrical Mirror Analyzer (CMA)
CMAs are advantageous over PMAs. They employ 2? geometry to overcome the low transmission with a PMA. A CMA
consists of two cylinders having a potential difference, V, between them. The entrance and exit slits are all contained on the
inner cylinder.

6/8/2021 Photoelectron Spectroscopy.2 https://chem.libretexts.org/@go/page/1836


Figure 3. Schematic of a CMA where the angle between the center of the cylinders and the electrons is 42.3 degrees. Rin is the
radius of the inner cylindar and Rout is the radius of the outer cylinder. The electron path should be more parabolic than the
overly elliptical shape shown here.

Here: V=1.3E0 ln(Rout/Rin) where L0=6.1(Rin) and E0 is in volts.


They are good for applications that require a high sensitivity with only a moderate resolution.
Cylindrical Deflection Analyzer (CDA)
CDAs consist of two cylinders spanning a 127 degree angle. It is this reason that CDAs are sometimes called "127 degree
analyzers."

Figure 4. Schematic of a CDA where the angle the cylinders span is 127 degrees.
The potential difference in a CDA is: 2V=E0(Rin/Rout) where E0 is the energy of incoming photoelectrons, in eV, that are
focused.
These analyzers have high resolution, however their transmission is low.
Spherical Deflection Analyzer (SDA)
SDAs are similar to CDAs, but they consist of two concentric hemispheres instead. In an SDA, the transmission of
photoelectrons with initial energy, E0, occurs along a path where R0=(Rin/Rout)/2. Since SDAs are the most common, prevalent
type of PES analyzer, they will be discussed in more depth than any of the previous analyzers as a thorough understanding of
how they apply to PES is, theoretically, of greater importance.

Figure 5. Schematic of an SDA. Only photoelectrons of the correct energy are able to pass through the detector with the right
arc and exit instead of colliding with the side walls of the hemispheres and becoming lost.
Here, the potential is different for both the inner and the outer hemisphere:
Vin = E0 [3 − 2(R0 / Rin )] (Photoelectron Spectroscopy.1)

6/8/2021 Photoelectron Spectroscopy.3 https://chem.libretexts.org/@go/page/1836


and
Vout = E0 [3 − 2(R0 / Rout )] (Photoelectron Spectroscopy.2)

The resolving power of these analyzers is proportional to the radius of the inner and outer hemispheres. These analyzers are
also capable of running in two separate modes when coupled with an optical system - fixed analyzer transmission mode (FAT)
and fixed retardation energy mode (FRR). In FAT mode, the lens either retards or accelerates the electrons so that all
photoelectrons enter the analyzer with the same kinetic energy. For this to occur, the analyzer is also arranged so that only
photoelectrons of a specific, fixed kinetic energy will pass through and reach the detector. In this case, the lens is scanned for
different energies. In FRR mode, the lens only retards the photoelectrons, and it does so in a uniform manner causing all
photoelectrons to be reduced in energy to a fixed value such as 15 eV, 30 eV, or whatever energy is desired. The hemispheres
of the analyzer here have a potential difference between them that is varied so that photoelectrons of different kinetic energies
can reach the detector. The more common of these two modes is FAT because it provides a greater signal intensity at low
electron kinetic energy and is also makes quantification of the spectra simpler.
These analyzers are a particularly good class of deflection analyzer. The slits in an SDA define the acceptable range of
entrance and exit trajectories a photoelectron may have when entering or leaving the analyzer. The photoelectrons that do
make it through the entrance slits will then only exit if they follow a specific, curved path down the middle of the two
hemispheres. The path they follow has the "correct energy" for exit to occur, and is determined by the selection of Vin and
Vout. Photoelectrons that are of higher or lower kinetic energy than what is defined by the hemispheres will be lost through
collisions with the walls.

Detection & Spectra


Detection relies on the ability of the instrument to measure energy and photoelectron output. One type of energy measured is
the binding energy, which is calculated through the following equation:

Ke = hν − BE − ϕ (Photoelectron Spectroscopy.3)

where:
Ke= Kinetic energy, this is measured
hν = Photon energy from the radiation source, this is controlled by the source

ϕ = Work function of the spectrometer, this is found through calibration

BE= Binding energy, this is the unknown of interest and can be calculated from the other three variables
Another part of PES detection is in the use of electron multipliers. These devices act as electron amplifiers because they are
coated with a material that produce secondary photoelectrons when they are struck by an electron. Typically, they are able to
produce two to three photoelectrons per every electron they are hit with. Since the signals in PES are low, the huge
amplification, up to 107 and higher when run in series so the secondary electrons from one multiplier strike the next, they
greately improve the signal strength from these instruments.
One type of spectra in these experiments is recorded by varying the potential difference between the plates or hemispheres of
the analyzer. The output is known as an electron kinetic energy spectrum and is obtained by measuring the photoelectron
current at the detector as a function of the voltage applied to the hemispheres or plates. The voltage is then used in the
calculation of kinetic energy.
Further detail on the spectra produced in PES experiments and the analysis of said spectra is planned for a future module on
the interpreation of photoelectron spectroscopy.

Limitations
The main limitation in a PES instrument is the resolution. The problems in resolution come from four main areas: the
dimensions of the analyzer, the widths of the entrance and exit slits, other charges such as outside electronic fields or outside
magnetic fields, and local charges inside the instrument itself arising from things such as contamination in the analyzer. Steps
can be taken to improve the resolution, but some methods then sacrifice other factors such as the sensitivity. Obtaining high
spatial resolution and high energy resolution always comes at the expense of the signal intensity.
One resolution improving technique that, then, messes with the sensitivity is changing the width of the entrance and/or exit
slits. For example, in an SDA these slits are what define the range of trajectories photoelectrons may have when entering or

6/8/2021 Photoelectron Spectroscopy.4 https://chem.libretexts.org/@go/page/1836


exiting the analyzer. Decreasing the widths will certainly cause the resolution to go up, but the smaller slit size will decrease
the number of photoelectrons allowed in and out of the analyzer, therefore lowering the sensitivity.
Another technique which was discussed above given it's relevance to the discussion on analyzers, is the addition of electron
optics to the instrument.
A third method of improving resolution is specific to XPS and is the addition of an x-ray monochromator to the system. These
monochromators eliminate satellite radiation from x-rays and are capable of narrowing the x-ray line width from ~1eV to
~0.2eV. The use of monochromatic x-rays also serves to simplify the spectrum.

Ultraviolet Photoelectron Spectroscopy - UPS


In early UPS, the sample was a gas or a vapor that is irradiated with a narrow beam of UV radiation. More modern UPS
instruments are now capable of studying solids as well. The photoelectrons produced are passed through a slit into a vacuum
region where they are then deflected by magnetic or electrostatic fields to give an energy spectrum. UPS is sensitive to the
very near surface region, up to around 10 nm in depth.

Figure 6. Schematic of a basic UPS instrument. As in figure 1 at the start of this module, the analyzer in use is an SDA. In this
instrument, there are no optics in use, nor is there an electron multiplier. This schematic shows separate chambers for the
sample and the analyzer, both of which are under UHV.
There are two main areas UPS is used to study:
1. Electronic structure of solids
2. Adsorbed molecules on metals
Specific examples of UPS studies include:
1. The measurement of molecular orbital energies that can be compared to theortical values calculated from quantum
chemistry
2. Determination and assignment of bonding, nonbonding, and/or antibonding molecular orbitals
3. The binding and orientation of adsorbed species on the surface of solids
4. Band structure mapping in k-space with angle-resolved techniques

Spectral output
Briefly, the spectrum produced from a UPS experiment has peaks that correspond to the ionization potentials of the molecule.
These also correspond to the orbital energies. Because of this, UPS can also give information on the vibrational energy levels
of the ions formed.

Limitations
UPS is capable only of ionizing valence electrons, which limits the range and depth of UPS surface experiments. Conventional
UPS has relatively poor resolution.

Advantages
Ultraviolet radiation has a very narrow line width and a high flux of photons available from simple discharge sources. Higher
resolution UPS scans allow for the observation of the fine structures that are due to vibrational levels of the molecular ion
which, then, allows molecular orbital assignment of specific peaks.

6/8/2021 Photoelectron Spectroscopy.5 https://chem.libretexts.org/@go/page/1836


X-Ray Photoelectron Spectroscopy - XPS
A diagram of a typical XPS instrument was shown at the beginning in figure 1. XPS is extremely good for surfaces. This is
because the kinetic energy of the escaping photoelectrons limits the depth able to be probed. The samples studied are all solids
of some type ranging from metals to frozen liquids. When the sample is irradiated, the electrons ejected are from the inner
shells of the atoms.
There are several areas suited to measurement by XPS:
1. Elemental composition
2. Empirical formula determination
3. Chemical state
4. Electronic state
5. Binding energy
6. Layer thickness in the upper portion of surfaces
Some specific examples of systems studied by XPS are:
1. Analysis of stains and residues on surfaces
2. Reactive frictional wear of solid-solid reactions
3. Silicon oxynitried thickness and measurements of dosage
4. Depth profiling: In depth profiling, a sputter source is used. This removes successive layers from the surface of a sample
and allows for the quanitation of element depth profiles to be recorded in the near-surface region. This is useful in the
composition of thin films.
5. Angle dependence measurements: When the angle of measurement is changed, the depth of the information gathered can
be varied by 1-10 nm. The usefulness here is in determining the concentration of additives in the surface region.
6. Imaging of surfaces
7. Utilizing a special imaging mode, the distribution of elements in surface structures can be determined. This technique is
useful in dimensions up to about 3 um.

Spectral output
Briefly, the spectrum from an XPS experiment is a graph of emission intensity vs binding energy. This allows elements on the
surface to be identified based on the unique binding energy each element has. The peak areas on these spectra can also be used
to obtain the concentration of the elements on the surface as well. Detailed information on the interpretation of XPS spectra is
planned for a future module.

Limitations
Despite the many benefits to XPS, nothing is foolproof, nor is anything without limitations. The smallest analytical area XPS
can measure is ~10 um. Samples for XPS must be compatible with the ultra high vacuum environment. Because XPS is a
surface technique, there is a limited amount of organic information XPS can provide. XPS is limited to measurements of
elements having atomic numbers of 3 or greater, making it unable to detect hydrogen or helium. XPS spectra also take a long
time to obtain. The use of a monochromator can also reduce the time per experiment.

Advantages
XPS has a greater range of potential application than UPS since it can probe down to core electrons. XPS is good for
identifying all but two elements, identifying the chemical state on surfaces, and is good with quantitative analysis. XPS is
capable of detecting the difference in chemical state between samples. XPS is also able to differentiate between oxidations
states of molecules.

UPS vs. XPS


One question that is always of consideration when multiple techniques are available for use is which technique is the best for
the system or sample of interest.
Here is a brief table of some of the experimental applications of PES and the suitability to either XPS, UPS, or both.

Application XPS UPS

6/8/2021 Photoelectron Spectroscopy.6 https://chem.libretexts.org/@go/page/1836


Application XPS UPS

Materials on surfaces X X

Depth profiling X
Angle dependent studies X X
Binding energy X X
Valence band fine structure X
Elemental composition X
Empirical formulas X
Electron energy levels X X

References
1. Hufner, Stefan. Photoelectron Spectroscopy: Principles and Applications. Springer 2003
2. Skoog. Principles of Instrumental Analysis. Brooks/Cole
3. Oxford Dictionary of Chemistry, 6th ed.. Oxford University Press
4. Somasundaran, P. ed. Encyclopedia of Surface and Colloid Science. Taylor and Francis
5. Ellis, Andrew M.; Feher, Miklos; Wright, Timothy G. Electronic and Photoelectron Spectroscopy: Fundamentals and Case
Studies. Cambridge University Press. 2005

Problems
1. What are the limitations involved with PES analyzers?
2. Is it possible to obtain both high sensitivity and high resolution with XPS? Why or why not?
3. Name three methods for improving the signal output from a PES instrument.
4. Can you study a system using both UPS and XPS? What are the advantages to using both techniques? Are there any
disadvantages?
5. Why is the SDA the most widely used analyzer for PES experiments?

Contributors and Attributions


Jessica Woods

6/8/2021 Photoelectron Spectroscopy.7 https://chem.libretexts.org/@go/page/1836


Photoelectron Spectroscopy: Theory
Photoelectron spectroscopy involves the measurement of kinetic energy of photoelectrons to determine the binding energy,
intensity and angular distributions of these electrons and use the information obtained to examine the electronic structure of
molecules. It differs from the conventional methods of spectroscopy in that it detects electrons rather than photons to study
electronic structures of a material.

Introduction
Photoelectron spectroscopy (PES) is the energy measurements of photoelectrons emitted from solids, gases, or liquids by the
photoelectric effect. Depending on the source of ionization energy, PES can be divided accordingly into Ultraviolet
Photoelectron Spectroscopy (UPS) and X-ray Photoelectron Spectroscopy (XPS). The source of radiation for UPS is a noble
gas discharge lamp, usually a He discharge lamp. For XPS, also referred to as Electron Spectroscopy for Chemical Analysis
(ESCA), the source is high energy X-rays (1000-1500 eV). Furthermore, depending on the source of the ionization energy,
PES can probe either valence or core electrons. UPS, which uses the energy of ultraviolet rays (<41 eV) will only be sufficient
to eject electrons from the valence orbitals, while the high energy X-rays used in XPS can eject electrons from the core and
atomic orbitals (Figure 1).

Figure 1. Scheme for UPS and XPS. Illustrates the ejection of electrons from respectively the valence electrons and core
electrons.1
Further information, about XPS and UPS, is discussed within the module on Photoelectron Spectroscopy: Application, which
includes a discussion on methods of study for both these spectroscopic techniques and a comparison.

Photoelectric Effect
To understand the principles of photoelectron spectroscopy, the photoelectric effect must be applied. The photoelectric effect
states that electrons can be pushed off the surface of a solid by electromagnetic radiation. The ejected electrons are called
photoelectrons.

Figure 2. Scheme of photoelectric effect. Incoming light hits the surface of a solid causing the ejection of a photoelectron.
Originally, known as the Hertz effect, the photoelectric effect was first observed by Heinrich Hertz in 1887, when Hertz
noticed that sparks would more readily jump between two charged surfaces that were illuminated with light. Hertz’s
observation would ultimately lead to Einstein’s photoelectric law; the kinetic energy of an emitted photoelectron is given by
Ek = hν − EI (Photoelectron Spectroscopy.1)

where h is Planck’s constant, ν is the frequency of the ionizing light, and EI is the ionization energy, which is synonymous with
electron binding energy, of the electron. The term photoelectric effect is only considered when discussing solids, exclusively.
As PES can be used for energy measurements of solids, liquids, and gases the term photoionization or photoemission better

6/29/2021 Photoelectron Spectroscopy.1 https://chem.libretexts.org/@go/page/1837


represents the principles of PES. Photoionization is the process in which molecule (M) is ionized by a beam of photons, in
which the molecule will lose an electron:
+

M + hν → M (Eint ) + e (Photoelectron Spectroscopy.2)

This process of photoionization follows the three-step model. The three-step model breaks down the process of
photoionization into three independent steps:
1. The molecule will absorb the photon, causing the energy of the photon to be transferred to the molecule's electrons, which
will become excited.
2. The excited electron will travel to the surface of the molecule. During this step, the excited electron travels it may or may
not collide with other particles. Any excited electrons which do collide with a particle will loss energy.
3. The excited electron will escape the surface of the molecule into the vacuum where it will be detected.
The process of photoionization can only occur if the photon has energy greater than the energy which is holding the electron to
the molecule, which is the lowest ionization potential. If there is excess energy after the ionization has occurred then the
excess energy will be in the form of kinetic energy. By rearranging equation 1, the ionization energy will be the difference
between the photon energy (hv) and the photoelectron's kinetic energy (Ek). These two variables, photon energy and kinetic
energy, are both measured by the PE spectrometer. Thus, by using PES it is possible to measure the energies of the ground and
excited states, after the loss of an electron from a neutral molecule, given by the above chemical formula.

Figure 3. A basic photoionization process, showing the different ionization potentials. The measurement of kinetic energy is
equal to hv-(Ip + Eint), as it represents the excess energy after photoionization, where Eint is the internal energy.4

Ionization Energy
Ionization energy, also known as electron binding energy, determined by photoelectron spectroscopy provides some of the
most detailed quantitative information about electronic structure of organic and inorganic molecules. Ionization is defined by
transitions from the ground state of a neutral molecule to the ion states (equation 2). There are two types of ionization energy:
adiabatic and vertical ionization energy. Adiabatic ionization energy of a molecule is defined as the minimum amount of
energy needed to eject an electron from the neutral molecule. Additionally, can be referred to as the difference between the
energy of the vibrational ground state of the neutral molecule and positive ion. The second type: vertical ionization energy
accounts for any additional transitions between the ground and excited vibrational state of the neutral molecule. The vertical
ionization energy is the most probable transition. The Frank-Condon principle explains the relative intensity of the vibrational
bands for photoionization transitions.
Koopman's theorem, which states that the negative of the eigenvalue of an occupied orbital from a Hartree-Fock calculation is
equal to the vertical ionization energy of the ion state formed by the photoionization of the molecule. Due to Kooperman's
theorem, ionization energies are shown to be directly related to the energies of molecular orbitals; however, there are
limitations to Koopman's theorem.
During the process of photoionization, the ejection of an electron will result in the formation of a positive ion (M+). The
energy required to cause the ejection of an electron is known as ionization energy or electron binding energy. Overall,
ionization energy will depend on the location of the electrons in preference to the nucleus of the molecule. As electrons are
arranged in orbitals surrounding the atomic nucleus, the ionization energy will be higher or lower depending on whether the
electrons are located in the core or valence shell. Obviously, core electrons, which are closer to the nucleus, will require more
energy to be ejected. Furthermore, each chemical element has a different number of protons in the nucleus, resulting in a
unique set of ionization energies for every element. By using photoelectron spectroscopy, the ionization energy is determined
by subtracting the energy of the incoming photon from the measured kinetic energy of the ejected electron. Thus, it is possible

6/29/2021 Photoelectron Spectroscopy.2 https://chem.libretexts.org/@go/page/1837


to use PES to determine the chemical elements within an unknown sample based on the observed ionization energies in a PE
spectrum.
The location of the ejected electron will factor greatly into which type of photoelectron spectroscopy is used. X-ray
photoelectron spectroscopy (XPS) is used to eject electrons from the core or valence shell. The sample used in XPS will first
be placed in an ultra-high vacuum chamber to prevent photons and emitted electrons from being absorbed by gases. Then the
sample will be bombarded with x-rays, causing the ejection of electrons. The ejected electrons energies will be measured by
their dispersal within an electric field. Due to the vacuum environment of the sample, XPS cannot be used for liquids. In
addition, XPS will provide information about oxidation states for any elements present in the sample, as the ionization
energies of core electrons are slightly higher when an oxidation state is present.
UPS works in a similar fashion as XPS, but uses photons, produced by a noble gas discharge lamp, in the ultraviolet range of
the spectrum. Originally, UPS was used only to determine the ionization energies of gaseous molecules; however, over the
years it is also attributed information to the electronic structure of molecules.

Splitting
Various types of splitting occur in the photoelectron spectrum due to the removal of an electron from an orbital. The Russell-
Saunders term symbol notation: L 2S+1
J
is used to determine the differences in the initial and final states for the spectral
transitions. The first type, spin orbit splitting is purely an initial state effect, which occurs during photoionization if an electron
is removed from a degenerate subshell. In addition, spin orbit splitting will never occur for s orbitals, as it depends on an
electron being removed from a degenerate subshell. The PE spectrum will represents spin orbit splitting for p, d, and f orbitals
as doublets for XPS. The intensity of the peaks for the doublets will depend on the J value in the Russell-Saunders term. For
example, the binding energy for the doublet with the lower J value will result in the highest intensity. Furthermore, due to
nuclear shielding the magnitude of spin-orbit splitting will decrease the further the way from the nucleus. Another type of
splitting is multiplet splitting which arises when there is interaction between an unpaired electron formed by photoelectron
ejected and an already pre-existing unpaired electron. This can result in the formation of multiple final states being formed
during the photoionization. For example, consider the three electron atom lithium. The ground state is 1s22s; 2S, which when it
undergoes photoionization can yield two 1s final states with different angular momenta:
+ 1 1 −
Li (1 s 2s; S )+e (Photoelectron Spectroscopy.3)

and
+ 1 3 −
Li (1 s 2s; S )+e (Photoelectron Spectroscopy.4)

Overall, the energy difference which occurs is known as multiplet splitting, which will result in a multi-peak envelope on the
PE spectrum. Lastly, Jahn-Teller splitting will occur when the symmetry of a molecule is destroyed by photoionization.

Photoelectron Instrumentation
All photoelectron spectrometers must have three components. The first is an excitation source used to irradiate the sample into
releasing electrons. The second is an electron energy analyzer which will disperse the emitted photoelectrons according to
their respective kinetic energy. Lastly, a detector. In addition, the spectrometer needs to have a high vacuum environment,
which will prevent the electrons from being scattered by gas particles. These various components in photoelectron
spectrometers are available in many different forms, which are discussed within the module on Photoelectron Spectroscopy:
Application. A block diagram of a basic PE spectrometer is listed below:

6/29/2021 Photoelectron Spectroscopy.3 https://chem.libretexts.org/@go/page/1837


Figure 4: A block diagram of PE spectrometer.
An example of a photoelectron spectrum obtained by a PE spectrometer is shown in Figure 5. This plot shows the kinetic
energy distribution of emitted photon obtained by the electron energy analyzer, resulting in a plot of of the number of electrons
detected versus the binding energy of electrons obtained.

References
1. Rabalais J. W., Principles of Ultraviolet Photoelectron Spectroscopy, Wiley, New York, 1977
2. Hüfner S., Photoelectron spectroscopy: principles and applications, Springer, Berlin; New York, 2003
3. Harris, D., Symmetry and spectroscopy: an introduction to vibrational and electronic spectroscopy, Dover Publications,
New York, 1989
4. Ghosh P. K., Intrdocution to Photoelectron Spectroscopy, Wiley, New York, 1983
5. Briggs D., Handbook of X-Ray and Ultraviolet Photoelectron Spectroscopy, Heyden, London, 1978
6. Bock, H.; Mollere, P. J.Chem. Educ. 1974, 51, 506-514
7. James, T. J. Chem. Educ. 1971, 48, 712-718

Problems
1. Which radiation source is used to eject core electrons?
2. Describe how PES can be used to calculate the ionization energy of a molecule.
3. Describe the photoelectric effect.

Answers
1. An X-ray radiation source.
2. PES involves a given energy of photon to ionize a molecule. As the excess energy, will be in the form of kinetic energy, is
calculated by the photoelectron spectrometer it is possible to calculate ionization energy of a molecule, by rearranging the
following equation: E = hν − E , to solve for E , ionization energy.
k I I

3. The photoelectric effect occurs when light hits a metal surface; thus, causing the ejection of electrons from the surface of
the metal.

Contributors and Attributions


Kristin Peck (UC Davis)

6/29/2021 Photoelectron Spectroscopy.4 https://chem.libretexts.org/@go/page/1837


CHAPTER OVERVIEW
X-RAY SPECTROSCOPY
X-ray Spectroscopy is a broadly used method to investigate atomic local structure as well as
electronic states. Very generally, an X-ray strikes an atom and excites a core electron that can either
be promoted to an unoccupied level, or ejected from the atom; both of these processes will create a
core hole.

EXAFS: THEORY
X-RAYS
Like light, X-rays are electromagnetic radiation with very short wavelengths. Thus, X-ray photons
have high energy, and they penetrate opaque material, but are absorbed by materials containing
heavy elements.

XANES: APPLICATION
XANES, short for X-ray Absorption Near-Edge Structure, is a subset of X-ray Absorption Spectroscopy. The absorption edge
corresponding to the liberation of a core electron from an element will exhibit several identifiable features which change depending
on the chemical environment of the element being probed. The study and modelling of the characteristics of near-edge features helps
answer questions about the oxidation state, coordination, and spin state of the probed element.

XANES - THEORY
X-ray Absorption Near Edge Structure (XANES), also known as Near edge X-ray Absorption Fine Structure (NEXAFS), is loosely
defined as the analysis of the spectra obtained in X-ray absorption spectroscopy experiments. It is an element-specific and local
bonding-sensitive spectroscopic analysis that determines the partial density of the empty states of a molecule.

XANES - THEORY II
XANES, x-ray absorption near edge structure, is sometimes also called NEXAFS, near edge x-ray absorption fine structure. It is
defined as partial analysis of x-ray absorption, and the range of XANES is between the threshold and where EXAFS begins. It can be
used to determine the oxidation state and coordination number of the metal center in a complex; as well as the covalency and site
symmetry of the molecule.

XAS - THEORY
XAS, or X-ray Absorption Spectroscopy, is a broadly used method to investigate atomic local structure as well as electronic states.
Very generally, an X-ray strikes an atom and excites a core electron that can either be promoted to an unoccupied level, or ejected
from the atom. Both of these processes will create a core hole. If the electron dissociates, this produces an excited ion as well as
photoelectron and is studied by X-ray Photoelectron Spectroscopy (XPS).

BACK MATTER

INDEX

1 7/7/2021
EXAFS: Theory
EXAFS (Extended X-ray Absorption Fine Structure) and XANES (X-Ray Absorption Near Edge structure) are regions of the
spectrum obtained from XAS(X-ray Absorption Spectroscopy). EXAFS corresponds to the oscillating part of the spectrum to
the right of the absorption edge(appearing as a sudden, sharp peak), starting at roughly 50 eV and extending to about 1000 eV
above the edge (shown in Figure 1). Through mathematical analysis of this region, one can obtain local structural information
for the atom in question.

Introduction
The ability of EXAFS to provide information about an atom's local environment has widespread application, particularly to the
geometric analysis of amorphous crystalline solids. Over time, EXAFS has become more applicable to quantitative analysis of
noncrystalline materials. When analyzing a single atom within a material, properties analyzed include coordination number,
disorder of neighboring atoms, and distance of neighboring atoms. Of these three properties, radial distance is the only
property reliably measured. Theoretically obtained structural information becomes more accurate the further down the EXAFS
region, ~0.02 Angstroms or better. Experimentally it has also been shown that application of the EXAFS technique is most
accurate for systems of low thermal or static disorder.

Qualitative understanding of the EXAFS region


X-Ray spectroscopy involves a process in which an X-ray beam is applied to an atom and causes the ejection of an electron,
usually a core electron. This leaves a vacancy in the core shell, and results in relaxation of an outer electron to fill that
vacancy. This phenomenon is only observed when the energy of the X-ray exceeds the ionization energy of the electrons in
that shell. We can relate this occurrence to the X-ray Absorption coefficient, which becomes the basis of EXAFS theory. The
X-ray Absorption coefficient, or μ, describes the relationship between the intensity of an X-ray beam going into a sample and
its intensity leaving the sample after traveling a distance x within the sample. The absorption coefficient is given by
dlnI
μ =− (EXAFS.1)
dx

In this expression, dx is the traversed distance, and I is the intensity. In a typical EXAFS spectrum, various sharp peaks will be
observed when energy(usually in eV) is plotted against absorbance. These peaks(called edges), which vary by atom,
correspond to the ionization of a core orbital, K-edges describing the excitation of the innermost 1s electron, and L-edges and
M-edges referring to the same for higher energy orbitals. A qualitative diagram of these energies, as well as their dependence
on atomic number, is shown in Figure 2 below.

6/29/2021 EXAFS.1 https://chem.libretexts.org/@go/page/1867


After each edge, a series of downward oscillations will be observed. These oscillations correspond to wave interactions
between the ejected photoelectron and electrons surrounding the absorbing atom. The neighboring atoms are called
backscattering atoms, since the waves emitted from the absorbing atom change paths when they hit these neighboring atoms,
returning to the original atom. Maxima in the oscillations result from constructive interference between these waves, and
minima result from destructive interference. These oscillations are also characteristic of the surrounding atoms and their
distances from the central atom, since varying distances will result in different backscattering paths, and as a result different
wave interactions. The EXAFS fine structure begins at roughly 30 eV past each edge, where oscillations begin to decay. In
addition, this wave interaction will depend on the mechanism of scattering, since the path taken by a wave sometimes involves
collision with an intermediate atom, or even multiple atoms, before it reverts to the absorbing atom.

Figure 3 effectively displays the phase shift that should occur (opposed to the diagram where E=E1) (the red arcs represent the
original diagram and the blue arcs represent the phase shift).

Extraction of Data using the EXAFS Equation


To represent the EXAFS region from the whole spectrum, a function χ can be roughly defined in terms of the absorption
coefficient, as a function of energy:
μ(E) − μo (E)
χ(E) = (EXAFS.2)
Δμo

Here, the subtracted term represents removal of the background and the division by Δμo represents the normalization of the
function, for which the normalization function is approximated to be the sudden increase in the absorption coefficient at the
edge. When interpreting data for the EXAFS, it is general practice to use the photoelectron wave vector, k, which is an
independent variable that is proportional to momentum rather than energy.
We can solve for k by first assuming that the photon energy E will be greater than E0 (the initial X-ray absorption energy at the
edge). Since energy is conserved, excess energy given by E - E0 is conserved by being converted into the kinetic energy of the
photoelectron wave. Since wavelengths are dependent on kinetic energies, the photoelectron wave (de Broglie wavelength)
will propagate through the EXAFS region with a velocity of ν where the wavelength of the photoelectron will be scanned. This
gives the relation, (E - E0) = meν2/2. One of the identities for the de Broglie wavelength is that it is inversely proportional to
the photoelectrons momentum (meν): λ=h/meν. Using simple algebraic manipulations, we are able to obtain the following:

UndefinedNameError: reference to undefined name 'math' (click for details)

To amplify the oscillations graphically, k is often plotted as k3


Now that we have an expression for k, we begin to develop the EXAFS equation by using Fermi's Golden Rule. This rule
states that the absorption coefficient is proportional to the square of the transition moment integral, or |<i|H|f>|2 , where i is the
unaffected core energy level before it interferes with the neighboring atoms, H is the interaction, and f is the final state in
which the core energy level has been affected and a photoelectron has been ejected. This integral can be related to the total
wavefunction Ψk, which represents the sum of all interacting waves, from the backscattering atoms and the absorbing atom.

6/29/2021 EXAFS.2 https://chem.libretexts.org/@go/page/1867


The integral is proportional to the total wavefunction squared, which refers to the probability that the photoelectron is found at
the atom where the photon is absorbed, as a function of radius. This wavefunction describes the constructive/destructive nature
of the wave interactions within it, and varies depending on the phase difference of the waves interacting. This phase difference
can be easily expressed in terms of our photoelectron wave number and R, the distance to the innershell from the wave, as
2k/R. In addition, another characteristic of this wave interaction is the amplitude of waves backscattering to the central atom.
This amplitude is can provide coordination number as well since it is thought to be directly proportional to the number of
scatters.
The physical description of all these properties is given in a final function for χ(k), called the EXAFS equation:
2 2
Nj fj (k)exp[−2 k σ ]exp[−2 Rj /λ]
j
χ(k) = ∑ sin[2kRj + δj (k)] (EXAFS.3)
2
j
kR
j

Looking at the qualitative relationships between the contributions to this equation gives an understanding of how certain
factors can be extracted from this equation. In this equation, f(k) represents the amplitude and δ(k) represents the phase shift.
Since these two parameters can be calculated for a certain k value, R, N, and σ are our remaining unknowns. These unknown
values represent the information we can obtain from this equation: radius, coordination number(number of scattering atoms),
and the measure of disorder in neighboring atoms, respectively. These are all also properties of the scattering atom.
We can also see the terms in this equation represented in a typical EXAFS spectrum: the sine term gives the origin of the
sinusoid shape, with a greater phase shift making the oscillations greater. In addition, this oscillation depends on the energy
and on the radial distance. The Debye-Waller Factor explains the decay with increasing energy, as well as increasing disorder.
This factor is partially due to thermal effects. In addition, we deduce the reason why EXAFS does not work over long
distances (up to 4-5 Å): the term Rj-2 causes the expression to decrease exponentially over large values of Rj (larger distances
between absorbing and scattering atoms), making EXAFS much weaker over long-distances as opposed to short ranged
neighboring atoms.
The last step in order to complete the data extraction is to take a Fourier transform of this expression into frequency space,
which results in a radial distribution function where peaks correspond to the most likely distances of the nearest-neighbors.

Multiple Scattering
The cases thus far have only dealt with single scattering pathways. Figure 4 shows a scenario in which two scattering atoms
(denoted by s) are present with one photoelectron wave source. In this case, what would happen? There is a clear dependence
on the distance and angles of the scattering atoms. Thus, it is no longer a simple matter of 1-D space distance relations, but
rather a 2-D polar coordinate system. In the case of Cu-Cu (shown by Kroneck et al) the copper is not only the absorbing
atom, it is also the scattering atom. Given this paradox, what would happen to the amplitude, phase, and frequency with
respect to the two coppers, and with respect to its neighboring atoms?

The effect of multiple scattering is a powerful tool, which is used to determine a variety of information on the local structure.
Each individual atom will produce more scattering or absorbing atoms depending on the type of atom and what type of wave
hits it. The angle at which the wave scatters back and its distance from the scattering atom will also affect the overall intensity
of the EXAFS spectrum. In scenario (a), shown in figure 4, the absorbing atom is hit by an X-ray where it produces an
absorbing wave which will scatter off S1 and hit S2. It will then send a scattering wave back to (a), which allows us to obtain
further information on the local structure of this molecule. However, this will only work if the absorbing atom is within 4-5 Å

6/29/2021 EXAFS.3 https://chem.libretexts.org/@go/page/1867


away from the scattering atoms. Otherwise, the wave becomes very dampened and the information retrieved becomes very
unreliable. In short, multiple scattering effects can be observed, but for the most part, they will only affect the EXAFS
spectrum a miniscule (but noticeable) amount. When an absorbing wave goes through another absorbing atom as mentioned in
Kroneck et al, it will produce a dampened absorbing wave (due to destructive interference) and give a lower signal than
normal. kroneck et al. tested this on two types of systems, the first was a copper single scattering onto a neighboring atom, the
second was two coppers scattering onto a scattering atom (all lined up linearly). When both of the radial distances were
collected, the second test showed only a small difference (0.02 Å) compared to the first test, suggesting experimentally that
multiple scattering waves do have an effect on the determination of a molecular system.

References
1. Que, L (2000). Physical Methods in Bioinorganic Chemistry. Sasuolito, California: University Science Books.
2. Kroneck, P. M. H., Antholine; W. E.; Kastrau, D. H. W.; Buse. G.; Steffens, G. C. M.; Zumft, W. G., "Multifrequency EPR
Evidence for a Bimetallic Centre at the CuA site in Cytochrome c Oxidase." Eur. J. Biochem. 1992, 209, 875-881
3. Farrar, J. A.; Lappalainen, P.; Zumft, W. G.; Saraste, M.; Thomson, A. J., "Spectroscopic and Mutagenesis Studies on the
CuA centre from the Cytochrome-c Oxidase complex of Paracoccus denitrificans." Eur. J. Biochem. 1996, 232, 294-303
4. J.J. Rehr and R.C. Albers, "Theoretical approaches to X-ray absorption fine structure", Reviews of Modern Physics 72
(2000), 621-654
5. C.A. Ashley and S. Doniach, "Theory of Extended X-ray Absorption Fine Structure (EXAFS) in Crystalline Solids",
Physical Review B. 11, 1279-1288.

Outside Links
en.Wikipedia.org/wiki/Extende...fine_structure
cars9.uchicago.edu/xafs_schoo...ille_Intro.pdf

Problems
Consider the crystal lattice zinc sulfide, whose unit cell is pictured below.

Zn = Purple, S = Blue.
1. An X-Ray Beam is applied to a sample of ZnS, and excites a core sulfur electron. Sketch the raw spectrum you might
expect to see, and indicate where the EXAFS/XANES regions may occur. Be sure to indicate where the K-edge is, and
include units.
2. For the phenomenon described, explain what happens to the energy associated with the ejection of the photoelectron?
3. Upon completing data extraction, and transforming the final EXAFS equation into frequency space, describe what you
would expect the radial distribution to look like for ZnS.
4. How would the amplitude for the oscillations of a Zn connected to a single sulfur compare to the amplitude for the
oscillations of a Zn from ZnS?
5. When a photoelectron is emitted, spherical waves are produced. Redraw this unit cell, and sketch two possible paths these
waves may take, taking into account nearest-neighbor atoms. Label the absorbing atom, backscattering atom, and

6/29/2021 EXAFS.4 https://chem.libretexts.org/@go/page/1867


intermediate atom, if applicable, in each diagram.

Answers
1. Sulfur's K-edge occurs at about 2.472 keV(this can be found using online sources). If energy is plotted against the
absorption coefficient, we would expect to see a constant flat line, then a sudden, sharp increase at roughly 2.472 keV.
After this peak the spectrum would begin to decay, and there would be a few slight peaks(XANES), then steady downward
oscillations(EXAFS).
2. When the core sulfur electron is excited, this means the energy of the photon from the X-ray source exceeds the bonding
energy of the electron. Its energy is consumed, and the electron becomes excited. This causes the ejection of the energy as a
photoelectron wave.
3. There would be expected to be a series of peaks, with four being prominent. These correspond to the four radial distances
more likely to be occupied around the sulfur atom. Since in this unit cell each sulfur atom's coordination number is four, we
would expect to see four main peaks, though in practice other peaks would probably interfere.
4. Amplitude of backscattering is believed to be proportional to coordination number. Therefore, we would expect the
amplitude of the oscillations to be about four times as high for this structure than for a single Zn-S bond.
5.

Contributors and Attributions


Ray Wong (UC Davis), Josh Alamillo (UC Davis)

6/29/2021 EXAFS.5 https://chem.libretexts.org/@go/page/1867


X-Rays
Learning Objectives
Explain X-rays.
Interpret the symbols used in the Bragg equation.

Like light, X-rays are electromagnetic radiation with very short wavelengths. Thus, X-ray photons have high energy, and they
penetrate opaque material, but are absorbed by materials containing heavy elements.

X-ray Diffraction
When light passes through a series of equal-spaced pinholes, it gives rise to a pattern due to wave interference, and such a
phenomenon is known as diffraction. X-rays have wavelengths comparable to the interatomic distances of crystals, and the
interference patterns are developed by crystals when a beam of X-rays passes a crystal or a sample of crystal powder. The
phenomena are known as diffraction of X-rays by crystals. More theory is given in Introduction to X-ray Diffraction.
X-ray diffraction, discovered by von Laue in 1912, is a well established technique for material analysis. This link is the home
page of Lambda Research, which provide various services using X-ray diffractions. For example:
Residual Stress Measurement
Qualitative Phase Analysis
Quantitative Phase Analysis
Precise Lattice Parameter Determination
In 1913, the father and son team of W.H. Bragg and W.L. Bragg gave the equation for the interpretation of X-ray diffraction,
and this is known as the Bragg equation.
2 d sin θ = n λ

where d is the distance between crystallographic planes, θ is half the angle of diffraction, n is an integer, and λ is the
wavelength of the X-ray. A set of planes gives several diffraction beams; each is known as the nth order.

Example 1
The X-ray wavelength from a copper X-ray is 154.2 pm. If the inter-planar distance from NaCl is 286 pm, what is the
angle θ ?
Solution
λ
sin θ = (1)
2d
154
= (2)
2 × 282

= 0.273 (3)


θ = 15.8

Example 2
An X-ray of unknown wavelength is used. If the inter-planar distance from NaCl is 286 pm, and the angle θ is found to
be 7.23°, what is λ ?
Solution
λ = 2 d sin θ (4)

= 2 × 282 × sin(7.23 ) (5)

= 71 pm (6)

7/7/2021 1 https://chem.libretexts.org/@go/page/35135
Example 3
The X-ray of wavelength 71 pm is used. If the inter-planar distance from KI is 353 pm, what is the angle θ for the second
order diffracted beam?
Solution
The calculation is shown below:
λ
sin θ = (7)
2d
71
= (8)
2 × 353

= 0.100 (9)

θ = 5.8 (10)

These examples illustrate some example of the applications of X-ray diffraction for the study of solids.

Questions
1. Hint: 30 degrees
2. If the wavelength is 150 pm and the interplanar distance d is 300 pm, what is the angle θ in the Bragg equation, for
n = 2?
Hint: 30 degrees
3. Hint: NaCl
Discussion -
The larger the interplanar distance d, the smaller the angle.

Contributors and Attributions


Chung (Peter) Chieh (Professor Emeritus, Chemistry @ University of Waterloo)

7/7/2021 2 https://chem.libretexts.org/@go/page/35135
XANES: Application
XANES, short for X-ray Absorption Near-Edge Structure, is a subset of X-ray Absorption Spectroscopy. The absorption edge
corresponding to the liberation of a core electron from an element will exhibit several identifiable features which change
depending on the chemical environment of the element being probed. The study and modelling of the characteristics of near-
edge features helps answer questions about the oxidation state, coordination, and spin state of the probed element.

Introduction
X-ray Absorption Near-Edge Structure (XANES), though less-developed or practiced than Extended X-ray Absorption Fine
Structure (EXAFS), may provide valuable information about the oxidation state, coordination environment, and bonding
characteristics of specific elements in a sample. The less common term Near-Edge X-ray Absorption Fine Structure
(NEXAFS) is used generally in the context of solid-state studies and is synonymous with XANES. The technique in practice
requires a mix of qualitative and quantitative analysis to interpret the data and draw conclusions.
The mechanisms and terminology in XANES are a subset of X-ray Absorption spectroscopy (XAS) and will only be
summarized here. The absorption edges of a material are the element-specific sudden increase in the absorption coefficient due
to the promotion of a core-level electron to unoccupied orbitals or unbound state. The core electron may be a member of the s,
p, d, or higher-order orbitals, provided it is not a valence (outer shell) electron. "XAS" refers to the nature of the electron that
is excited rather than the energy range of the exciting photon. Edges resulting from the excitation of an n=1 (1s) electron are
termed "K-edges", n=2 are "L-edges", n=3 "M-edges", and so on. The figure below labels these shells and corresponding
allowed bound-state absorption levels that are seen as pre- and on-edge features in XANES. Dotted lines correspond to
observed emission lines. When an orbital at the top of a dotted line is not full, the dotted line also indicates an absorption from
a lower shell (a bound-state transition). For example, the K-edge of Fe has an on-edge bound-state feature corresponding to 1s
-> 4p absorption, which would be the dotted line in the figure between K1 and N2,3. These follow selection rules ∆L = +/- 1.

XANES is the study of the features immediately before and after the edge, within approximately 1% on either side of the main
edge energy. Features include the edge position (a primary indicator of oxidation state), presence/shape of small features just
before the main edge ("pre-edge" features), and intensity, number, position and shape of peaks at the top of the main edge. The
Fe K-edge shown below from Carpenter (2010) labels some of these features. The pre-edge feature is weak as it involves a
forbidden transition (which is partially allowed due to mixing of ligand p-character), while the first line on the main edge is
due to the allowed 1s to 4p bound-state transition. In some compounds this transition is much stronger than the rest of the edge
(see Sulfur K-edge in next figure) and was historically called the "white line" due to its saturated appearance on the
photographic film used to record the spectrum.

6/14/2021 XANES.1 https://chem.libretexts.org/@go/page/1868


The power of XANES lies in the sensitivity of edge features to the chemical environment. The sensitivity varies among
elements, from just-detectable to pronounced. George et. al provide an excellent illustration of the range of the sulfur K edge
among various organic compounds. [3]

Instrumentation
Unlike visible, infrared, or microwave spectroscopy, no single off-the-shelf commercial instrument exists for XAS. The
equipment must be assembled from multiple components to suit the needs of the experiment and may involve a significant
amount of custom engineering. Laboratory X-ray sources were used before the advent of synchrotron radiation sources, but
today they are rarely considered. The typical XAS experiment at a synchrotron may be broken down into two general parts:
1. The "beam line": the set of components (bend magnet or undulator/wiggler, mirrors, monochromator(s), slits, diagnostics)
that produce and deliver a controllable high-intensity monochromatic X-ray beam.
2. The "end station": the set of instruments specific to the type of measurement, including sample handling (gas cells, fluid
cells, cryostats, positioning stages), detectors, diagnostics and support systems, and radiation protection enclosure.
XANES measurements use essentially the same equipment and setup as EXAFS, though as the energy range covered is much
smaller the emphasis may be on high-resolution over wide-range abilities. A typical beam line schematic is shown below.

Measurement Methods
Experimental considerations for different measurement techniques
Transmission Electron Yield Fluorescence Yield

Sample Thickness Thin Thick/Any Thick/Any

Background High Moderate Low


Sensitivity Bulk Surface Bulk
Sample Concentration High High Low

Transmission

6/14/2021 XANES.2 https://chem.libretexts.org/@go/page/1868


As the name implies, this method involves passing X-rays through the sample and comparing the incident to the transmitted
intensities. The sample thickness must be considered before measurement, as absorption coefficients vary greatly across the X-
ray energy range and among materials. For too-thick samples the beam may be totally attenuated at the edge, while too-thin
results in poor signal-to-noise ratio. For ultra-soft measurements the sample may need to be less than 1 micron thick, while
high-Z K-edges may need many centimeters of sample. Reference databases such as the LBNL Center for X-ray Optics
database give attenuation data and calculation tools to help estimate the proper thickness, which will depend on the
concentration of the measured species and the attenuation of other species present.
Measurement of the intensity may be made with gas ionization chambers, photodiodes, PN-junctions, metal grids, or from
scattering off optics or windows. The incident beam (I0) must be only partially sampled, typically by an ion chamber or grid,
while the transmitted beam (I1) may be blocked and absorbed completely. A helpful calibration technique is to place a
reference standard and detector after the I1 detector, so that the beam path looks as in the following diagram so the reference
spectrum is measured simultaneous with the sample.

Transmission measurements may be performed on any type of sample (gas, liquid, solid) provided the thickness and density is
controllable. For dilute measurements the signal-to-noise ratio is typically poor.
Electron Yield
Absorption of X-rays results in emission of electrons from the sample proportional to the absorption coefficient, from both
photoelectrons and Auger electrons. Photoelectrons are those that are ejected from core orbitals and will have a kinetic energy
that is the difference of the X-ray energy and their binding energy. Auger electrons are emitted as part of the relaxation process
as a higher-orbital electron fills the hole left behind by the photoelectron. The Auger electron energy is characteristic of the
element and core level being occupied and is analogous to a fluorescence photon. Collecting all produced electrons is known
as Total Electron Yield (TEY) and is measured with electron multipliers electrical current through a lead (in vacuum), or via
gas ionization collected by a grid close to the sample surface (non-vacuum). Alternately, an electron energy analyzer may be
used to discriminate between photo and Auger electrons, which helps reject background from other species at the expense of
throughput. This method is Partial Electron Yield (PEY). This method is only sensitive to the first ~10s of nm of the sample
surface due to electron scattering, and only works for solid samples. This may be an advantage for thin films or monolayers on
a substrate where other methods would suffer high background issues. Auger electrons and fluorescence are competing
processes and the ratio between the two changes depending on the element. Auger yield predominates for light elements like C
and N, but gradually decreases in favor of fluorescence as Z increases. The crossover where fluorescence is greater than Auger
yield is Z=30.[27]
Fluorescence Yield
As with electron yield, X-ray fluorescence is proportional to the absorption coefficient as valence electrons emit photons to fill
the core holes left by absorption. As with electron yield, fluorescence yield may be measured in total or partial modes, the later
requiring an energy-discriminating detector. For partial-yield, solid-state detectors such as Germanium and drifted Silicon
detectors are often used due to their high efficiency and moderate energy resolution (enough to separate emission from
different species). Dispersive spectrometers with gratings or Bragg crystals may also be used when higher resolution is needed
to separate species emissions that are close-together. The detector is typically oriented perpendicular to the incident beam
polarization to suppress the elastic scattering peak and improve signal-to-noise. Total yield may be measured with any X-ray-
sensitive detector not immediately in the incident or transmitted beam.
Fluorescence yield is especially well-suited for dilute samples due to its selectivity and bulk-sensitivity. However, when used
for concentrated samples, a phenomenon called "self-absorption" can lower the apparent absorption coefficient at high levels
of absorption. This is partially due to non-negligible reduction in the penetration depth as the absorptivity increases (directly
impacting the coefficient due to Beer's law), and re-absorption of the fluorescence photons by the same species before the
photons can leave the sample.

Hard X-ray Instrumentation

6/14/2021 XANES.3 https://chem.libretexts.org/@go/page/1868


Where "Hard" X-rays begin varies widely, typically between 2 and 10 keV photon energy. They are the X-rays that are high
enough in energy to penetrate significant distances through materials. This penetrating capacity is a blessing and a curse: many
materials may be used as windows, the sample can be in a variety of environments, and measurements are less sensitive to
thickness tolerances; however, optics must take this penetration into account, and complete shielding is needed to protect the
user from radiation. The K-edges of first-row transition metals and L-edges of rare earth elements fall in this range. Solid
samples may be contained in a metal or plastic holder with low-Z material windows. Liquids and gases may require custom
sample alignment which is usually automated with motorized stages. The endstation is shielded from the user by a protective
hutch. Radiation-sensitive samples may be cooled to low temperatures (liquid Nitrogen or Helium temperature) in a cryostat
modified with X-ray windows. The X-ray beam often passes from the high-vacuum beamline through a Beryllium window and
travels moderate distances through air or an exchange gas such as Helium before impinging the sample. Transmission and
fluorescence yields are the primary measurement methods.

Soft X-ray Instrumentation


"Soft" X-rays are those that are readily absorbed by most materials and air with very short attenuation lengths, roughly in the
range of ~100eV to 2-5 keV. The entire experiment (sample, detector, diagnostics) must be contained in vacuum up to Ultra-
High Vacuum (UHV). Window materials are few and must be thin. Sample containment must take into account in the vacuum
and window materials. Most often, the sample is inserted directly into vacuum as a solid (as a powder, crystal or amorphous
solid), or dried a substrate. Gas cells are also used for appropriate samples and for reaction/catalysis experiments. Electron and
fluorescence yields predominate in Soft X-rays.

Data handling and Analysis


The treatment of data depends on the compound complexity and nature of the problem solved. Very complicated molecules
which are difficult to simulate with software may be compared to simpler model compounds to determine coordination
electronic structure.
Recent advances in theoretical calculations and software allow general users to fit the spectrum based on ionization state,
coordination group, and various molecular parameters such as crystal field splitting and degree of orbital hybridization. The
post-edge region dominated by multiple-scattering may provide structural information directly similar to EXAFS, though by a
different approach. While simple-scattering in EXAFS is derived from the Fourier transform of the post-edge oscillations,
multiple-scattering involves fully modeling the quantum-mechanical spherical wave scattering from local neighbors via
Green's functions.[16]
Other software packages utilizing Density Functional Theory (DFT) or Charge-Transfer Multiplet Theory (CTM) are used to
derive the manifold of atomic or molecular orbital energies from bound-state transitions.[11]
Figure 1 from Metzler et. al[9] illustrates the assignment of components from simulation which may be correlated with the
observed spectrum. The assignments of the features by Metzler et al. are as follows: "peak 1 at 285 eV, corresponds to the C ls
→ ∏* transition in C=C double bonds; peak 2 at 287 eV is the C-H C ls → ∑*, mostly due to PLL side chains; peaks 3 and 4
at ~288 eV are both associated with C=0 C ls → ∏*; peak 5 at ~290 eV, corresponds to the C ls → ∏* transition in C=0
double bonds in carbonates." They include a dotted feature for the absorption edge due to ionization (core electron -> unbound
continuum states).
Investigation of multiple edges of the same element may also offer insight, as different features are emphasized at different
edges. The lower-energy edges are usually more highly-resolved than the K-edge due to longer core-hole lifetimes (by the
Heisenberg uncertainty principle, longer-lived states are less broad in energy) and allow the ionization state to be determined
more accurately. A comparison of the K, L, and M-edges of Molybdenum is shown below in which the analogous feature
(bound state transitions to d orbitals) is lined up at 0 eV in each spectrum.

6/14/2021 XANES.4 https://chem.libretexts.org/@go/page/1868


Applications
Biomolecules
The understanding of large molecules such as proteins can benefit greatly from XANES. The active-site clusters often contain
one or several metal ions embedded in bulk proteins. The presence of large amounts of organic backbone may interfere with
UV-Vis spectroscopy of the active site, while XANES, being element-specific, allows single atoms to be measured in even
large proteins. Preparation of the enzyme in resting and active states before the measurement and careful handling during the
measurement may elucidate the oxidation and coordination state. Depending on the substrate, the bound-state may be
measured from both the metal and ligand XAS.
A literature example of the sort of information which can be derived from XANES is the paper by Ralston et. al [20] to
determine the spin state of Ni in the active site of CO-Dehydrogenase (CODH). Through a careful study of multiple model
compounds of known oxidation states and spin configurations, from Ni(I), to low- and high-spin Ni(III) up to Ni(IV), a
relationship is derived between the position of the L3 edge and the ratio of the integrals of the L3 and L2 edges. A large ratio
between the two edges indicates a high-spin complex while lower ratio indicates low-spin, with the turning point occurring at a
ratio of 0.71 L3:L2 intensity. This scheme is applied to CODH prepared as a film in its resting state, CO-bound state, and
dithionate-reduced state to determine the Ni oxidation and spin states.

Solid State physics


Edge features corresponding to bound-state transitions will have the symmetries of the immediate coordination environment
(that surrounding the probed element only). For solid state samples in a single-crystal form, this may be combined with the
fact that synchrotron radiation is typically linearly polarized in the plane of the storage ring to probe specific orbitals of the
element. Electronic structure and band gap measurements are therefore possible. [21,25]

Materials science
Materials scientists often care about the behavior of materials under extreme conditions. High-pressure diamond cells
combined with powerful lasers allow the probing of materials in the regions of 5000K and 100s of GPa.[24] Small X-ray beam
spot sizes on the order of microns allow so-called "Micro-XANES" measurements to probe changes along pressure gradients
of materials in these cells. Temperature-dependent XANES measurements can capture phase transitions in materials and give
insight into structural changes.[22]

6/14/2021 XANES.5 https://chem.libretexts.org/@go/page/1868


Development of novel materials such as high-Tc superconductors and advanced scintillators can also benefit from XANES.[23]
The low-concentration of dopants in scintillators is perfectly suited to fluorescence-yield XAS to determine concentration and
ionization state. The plot below depicts the M4 and M5 edges of europium doped at 1% measured by partial fluorescence yield.

Catalysis
Specially-designed cells combined with XANES may provide insight into reaction dynamics and changes in electronic
configuration and oxidation state. Performing the reaction under different temperatures and conditions gives insight into
dynamics. With the development of fast, continuous-scan beamlines with the capability to scan an entire edge in less than a
minute, time-resolved studies are now possible as well. Leoferti et. al followed the Cu oxidation state during oxychlorination
of ethylene catalyzed by copper chloride,[26] which was built-upon later by Lamberti et. al with time-resolved XANES with
30-second resolution.[13] Further experiments are possible with the development of reaction cells which allow multiple
simultaneous measurements, such as XANES, UV-Vis and electro-chemistry.[14]

Surface Science
The shallow penetration depth of soft X-rays combined with the probing depth of electron yield measurement allows for very
sensitive studies on films as thin as a monolayer.[15] When deposited on a single-crystal surface in controlled conditions, the
monolayer species can be made to bond with a particular orientation to that surface allowing for polarization-dependent
studies. As the X-rays emitted by bend magnets and standard wigglers or undulators is linearly polarized in the plane of the
ring, changing the orientation of the crystal to the incident beam can suppress or enhance features in the absorption edge. The
monolayer bonding of the sample to substrate may also alter the characters of the edge independent of the polarization[17]
which opens up further possibilities for understanding the system.

Sample Diagnostics
Experiments which benefit from synchrotron radiation, such as X-ray crystallography, Scanning Transmission X-ray
Microstopy (STXM), and X-ray Photo-emission Electron Microscopy (XPEEM), typically need large doses of radiation to be
effective. For radiation-sensitive samples the dose may be an important consideration as to the validity of the data: for
crystallography, the high-energy electrons liberated from atoms after ionization may distort the very structure one is trying to
measure, while STXM and PEEM may damage or destroy the large bio-structures often probed with these methods. XANES,
combined with an appropriate model for the absorption of radiation and its affects on the edge features, may be used to both
set limits and parameters for the experiment[18,19] and monitor the sample condition as the experiment progresses.

References
1. Frank de Groot and Akio Kotani, "Core Level Spectroscopy of Solids", CRC Press, Boca Raton, FL (2008).
2. Joachim Sthöhr, "NEXAFS Spectroscopy", Second Printing, Springer-Verlag, Heidelberg, Germany (2003).
3. Graham N. George, Martin L. Gorbaty, "Sulfur K-edge x-ray absorption spectroscopy of petroleum asphaltenes and model
compounds"; J. Am. Chem. Soc. 111 (9), pp 3182–3186 (1988).
4. C. R. Natoli, D. K. Misemer, S. Doniach, and F. W. Kutzler, "First-principles calculation of x-ray absorption-edge structure
in molecular clusters"; Phys. Rev. A 22, 1104–1108 (1980).
5. Owen B. Drury, "Development of High Resolution X-Ray Spectrometers for the Investigation of Bioinorganic Chemistry
in Metalloproteins"; Ph.D. Thesis, University of California, Davis (2007).

6/14/2021 XANES.6 https://chem.libretexts.org/@go/page/1868


6. H. Oyanagi, Z. H. Sun, Y. Jiang, M. Uehara, H. Nakamura, K. Yamashita, L. Zhang, C. Lee, A. Fukanoa, and H. Maeda,
"In situ XAFS experiments using a microfluidic cell: application to initial growth of CdSe nanocrystals"; J. Synchrotron
Radiaion 18, 272–279 (2011).
7. Simon J. George, Owen B. Drury, Juxia Fu, Stephan Friedrich, Christian J. Doonan, Graham N. George, Jonathan M.
White, Charles G. Young, Stephen P. Cramer, "Molybdenum X-ray absorption edges from 200 to 20,000 eV: The benefits
of soft X-ray spectroscopy for chemical speciation"; J. Inog. Biochem. 103, 157–167 (2009).
8. Matthew H. Carpenter, "Helium Atmosphere Chamber for Soft X-ray Spectroscopy of Biomolecules", MS Thesis,
University of California, Davis (2010).
9. Rebecca A. Metzler, Ronke M. Olabisi, Mike Abrecht, Daniel Ariosa, Christopher J. Johnson, Benjamin Gilbert, Bradley
H. Frazer, Susan N. Coppersmith, and P.U.P.A Gilbert, "XANES in Nanobiology"; Proceedings of X-ray Absorption Fine
Structure—XAFSU 13, American Institute of Physics, 51 (2007).
10. S. Della Longa, A. Arcovito, M. Girasole, J. L. Hazemann, and M. Benfatto, "Quantitative Analysis of X-Ray Absorption
Near Edge Structure Data by a Full Multiple Scattering Procedure: The Fe-CO Geometry in Photolyzed Carbonmonoxy-
Myoglobin Single Crystal"; Phys. Rev. Lett. 87, 15, 155501 (2001).
11. E. Stavitski and F.M.F. de Groot, "The CTM4XAS program for EELS and XAS spectral shape analysis of transition metal
L edges"; Micron 41, 687 (2010).
12. Takashi Fujikawa, "Basic Features of the Shrort-Range-Order Multiple Scattering XANES Theory"; J. Phys. Soc. Japan 62,
6, pp. 2115 (1993).
13. Carlo Lamberti, Carmelo Prestipino, Francesca Bonino, Luciana Capello, Silvia Bordiga, Giuseppe Spoto, Adriano
Zecchina, Sofia Diaz Moreno, Barbara Cremaschi, Marco Garilli, Andrea Marsella, Diego Carmello, Sandro Vidotto, and
Giuseppe Leofanti, "The Chemistry of the Oxychlorination Catalyst: an In Situ, Time-Resolved XANES Study"; Angew.
Chem. Int. Ed. 41, No. 13, pp 2341 (2002).
14. Walkiria S. Schlindwein, Aristea Kavvada, Roger G. Linford, Roger J. Latham and J. Günter Grossmann, "Combined
XAS/SAXS/Electrochemical studies on the conformation of poly(vinylferrocene) under redox conditions"; Ionics 8, 1-2,
85-91 (2002).
15. A. Biancoli, "Surface X-ray absorption spectroscopy: Surface EXAFS and surface XANES"; Applications of Surface
Science 6, 3-4, 392-418 (1980).
16. M. Benfatto, C. R. Natoli, A. Bianconi, J. Garcia, A. Marcelli, M. Fanfoni, and I. Davoli, "Multiple-scattering regime and
higher-order correlations in x-ray-absorption spectra of liquid solutions"; Phys. Rev. B 34, 5774–5781 (1986).
17. E. E. Doomes, P. N. Floriano, R. W. Tittsworth, R. L. McCarley, and E. D. Poliakoff, "Anomalous XANES Spectra of
Octadecanethiol Adsorbed on Ag(111)"; Phys. Chem. B 107 (37), 10193–10197 (2003).
18. J.Wang, C. Morin, L. Li, A.P. Hitchcock, A. Scholl, A. Doran, "Radiation damage in soft X-ray microscopy"; Journal of
Electron Spectroscopy and Related Phenomena 170, 25–36 (2009).
19. James W. Murray, Enrique Rudiño-Piñera, Robin Leslie Owen, Martin Grininger, Raimond B. G. Ravelli, and Elspeth F.
Garmana, "Parameters affecting the X-ray dose absorbed by macromolecular crystals"; J. Synchrotron Rad. 12, 268-275
(2005).
20. C. Y. Ralston, Hongxin Wang, S. W. Ragsdale, M. Kumar, N. J. Spangler, P. W. Ludden, W. Gu, R. M. Jones, D. S. Patil,
and S. P. Cramer, "Characterization of Heterogeneous Nickel Sites in CO Dehydrogenases from Clostridium
thermoaceticum and Rhodospirillum rubrum by Nickel L-Edge X-ray Spectroscopy"; J. Am. Chem. Soc. 122, 10553-
10560 (2000).
21. O. Seifarth, J. Dabrowski, and P. Zaumseil, S. Müller and D. Schmeißer, H.-J. Müssig, and T. Schroeder."On the band gaps
and electronic structure of thin single crystalline praseodymium oxide layers on Si(111)"; J. Vac. Sci. Technol. B 27, 271
(2009).
22. B. Ravel and E.A. Stern, "Temperature and Polarization Dependent XANES Measurements on Single Crystal PbTiO3"; J.
Phys IV France 7, 1223 (1997).
23. Stephan Friedrich, Owen B. Drury, Shaopang Yuan, Piotr Szupryczynski, Merry A. Spurrier, and Charles L. Melcher, "A
36-Pixel Tunnel Junction Soft X-Ray Spectrometer for Scintillator Material Science"; IEEE Transactions on Applied
Superconductivity 17, 2, 351 (2007).
24. Giuliana Aquilanti, Sakura Pascarelli, Olivier Mathon, Manuel Muñoz, Olga Naryginac, and Leonid Dubrovinsky,
"Development of micro-XANES mapping in the
diamond anvil cell"; J. Synchrotron Rad. 16, 376–379 (2009).

6/14/2021 XANES.7 https://chem.libretexts.org/@go/page/1868


25. V. L. Mazalova and A. V. Soldatov, "Geomtrical and Electronic Structure of Small Copper Nanoclusters: XANES and DFT
Analysis"; Journal of Structural Chemistry 49, Supplement, S107-S115 (2008).
26. G. Leofanti, A. Marsella, B. Cremaschi, M. Garilli, A. Zecchina, G. Spoto, S. Bordiga, P. Fisicaro, C. Prestipino, F. Villain,
and C. Lamberti, "Alumina-Supported Copper Chloride: 4. Effect of Exposure to O2 and HCl"; Journal of Catalysis 205,
375–381 (2002).
27. Krause, M.O., and Oliver J.H., "Natural Widths of Atomic K and L Levels, Ka XRay Lines and Several KLL Auger
Lines"; Journal of Chemical and Physical
Reference Data 8(2), 329-337 (1979).

Simulation Software
CTM4XAS (ionization and coordination chemistry)
FEFF (multiple scattering)
FitIt graphical front-end for suite of fitting software.

Problems
1. As the oxidation state of an atom increases, which direction does the absorption edge shift (higher or lower in energy)?
2. Name 2 characteristcs which may be determined by XANES.
3. What measurement method is most appropriate for measuring the Cu K-edge of the protein Stellacyanin? What would be
most appropriate for the Ni L-edge of NiO?
4. List the energies of the K- and L-edges of Si, Fe, and Zn, and the K-edges of C, N, P and S.
5. Why are pre-edge features corresponding to forbidden transitions sometimes observed on metal K-edges?

Solutions
1. The edge shifts higher in energy. The lesser amount of shielding of the nucleus by surrounding electrons increases its
effective charge Zeff; the more tightly-bound electrons require more energy to liberate.
2. Possible answers: oxidation state, valence (number of ligands), coordination geometry, near-neighbor distances (through
multiple-scattering).
3. Stellacyanin is a 20 kDa protein with 1 Cu, putting it in the "dilute" regime; it should be measured by Partial Fluorescence
Yield. NiO is stochiometrically 1/2 Ni; due to the concentration, the high fluorescence-yield of Ni, and the fact that the Ni
L-edge lies in the soft X-ray region means it should be measured by Total or Partial Electron Yield.
4. Si: K=1849eV; L1=149.7eV; L2=99.8eV; L3=99.4eV; Fe: K=7112eV; L1=844.6eV; L2=719.9eV; L3=706.8eV; Zn:
K=9659eV; L1=1196.2eV; L2=1044.9eV; L3=1021.8eV; C: K= 284.2eV; N: K=409.9eV; P: K=2145.5eV; S: K=2472eV.
5. The 1s → 3d transition is quantum mechanically forbidden; however, in the view of molecular orbital theory, the metal 3d
mix with ligand 2p or 3p orbitals and gain some "p orbital character", weakly alloing the transition.

6/14/2021 XANES.8 https://chem.libretexts.org/@go/page/1868


XANES - Theory
X-ray Absorption Near Edge Structure (XANES), also known as Near edge X-ray Absorption Fine Structure (NEXAFS), is
loosely defined as the analysis of the spectra obtained in X-ray absorption spectroscopy experiments. It is an element-specific
and local bonding-sensitive spectroscopic analysis that determines the partial density of the empty states of a molecule.

Introduction
X-rays are ionizing electromagnetic radiation that have sufficient energy to excite a core electron of an atom to an empty
below the ionization threshold called an excitonic state, or to the continuum which is above the ionization threshold. Different
core electrons have distinct binding energies; consequently, if one plots the X-ray absorbance of a specific element as a
function of energy1, the resulting spectrum will appear similar to Figure 1.

Figure 1: Transitions that contribute to XAS edges. (CC BY-SA 3.0; Atenderholt via Wikipedia)

Theory of XANES
Core hole
As stated above, a core hole is the space a core electron occupied before it absorbs an X-ray photon and ejected from its core
shell. core holes are extremely energetic (electronegative) which leads to their unstable nature. the average life span of a core
hole is close to 1 femtosecond. A core holes is created through processes in which either a core electron absorbs an X-ray
photon (X-ray absorption) or absorbs part of the an X-ray photon's kinetic energy (X-ray Raman scattering). The successor
process is the decay of a core hole which can take place either through Auger electron ejection of X-ray Fluorescence.

Absorption edge
As the energy of X-ray radiation is scanning through the binding energy regime of a core shell, a sudden increase of absorption
appears, and such phenomenon corresponds to absorption of the X-ray photon by a specific type of core electrons (eg. 1s
electrons of Cu). This gives rise to a so-called absorption edge in the XAS spectrum due to its vertical appearance. The name
of the absorption edges are given according to the principle quantum number, n, of the excited electrons (Table 1).
Table 1: Absorption edges.
K edge 1s

6/25/2021 1 https://chem.libretexts.org/@go/page/1869
L edge 2s 2p

M edge 3s 3p 3d

N edge 4s 4p 4d 4f

Energies of absorption edges in X-ray absorption spectra reveal the identity of the corresponding absorbing elements.
However, more useful information can be obtained by a closer examination of a giving absorption edge (Figure 1 ).

Figure 2: Figure courtesy of Wikipedia.


Illustrated by figure 2, the absorption edge is often much more complex than simply an abrupt increase in absorption
illustrated in figure 1. There are weak transitions below the absorption edge, namely pre-edge structures, as well as significant
absorption features in the immediate neighborhood of the absorption edge and well above the edge. The structure found in the
immediate neighborhood of the absorption edge, conventionally within 50 eV of the absorption edge, is referred to as X-ray
Absorption Near Edge Structure (XANES). Beyond XANES, the oscillatory structure cause by the interference between the
outgoing and the back-scattered photoelectron waves is referred to as Extended X-ray absorption Fine Structure (EXAFS),
which can extend to 1000 eV or more above the absorption edges3.

Dipole selection rule


XANES directly probes the angular momentum of the unoccupied electronic states: these states can be bound states (exitonic)
or unbound states (continuum), discrete or broad, atomic or molecular. The dipole selection rule for transition determination is:
Δl= ±1, Δj= ±1, Δs= 0. Commonly observed ALLOWED transitions are tabulated in Table 2.
Table 2: Spin and Orbitally allowed transitions.
Initial state Final state

s p

p s, d
d p, f
f d,g

White Line
In certain XANES spectra, the rising absorption edge might lead to a sharp intense peak referred as “white line”. The reason is
that in the past, X-ray absorption spectra were recorded using photographic plates, and the strong absorption of certain
wavelength leads to an unexposed band on the photographic plates which after develops in negative, appear as a white vertical
stripe, hence the term “white line”.

X-ray absorption measurement


X-ray fluorescence
As a core electron absorbs an X-ray photon and is ejected from its core shell, the absence of the electron left in the core shell
leads to a core hole state. Core hole states are highly excited and can relax in mainly two ways: Auger electron emission or X-

6/25/2021 2 https://chem.libretexts.org/@go/page/1869
ray fluorescence. For higher-energy excitation (e.g., for the K edges of elements with atomic numbers greater than 40), X-ray
fluorescence is the primary relaxation process. The scheme of an X-ray fluorescence is illustrated in Figure 3.

Figure 3: A schematic illustration of X-ray Fluorescence measurement.


The intensity of X-ray fluorescence is described by Equation 1.
IF
A =( ) (1)
I1

The intensity of X-ray fluorescence is directly proportional to the X-ray absorption cross-section of the sample. However, in
practice as a beam of X-ray is shined on a sample, a variety of X-rays are emitted, they can be fluorescent X-ray from the
sample and background X-ray from the sample scattering. In order to improve the sensitivity, energy-resolving solid-state
fluorescence detectors2 are used to selectively distinguish background radiation from signal of interests.

Figure 4: Typical wavelength dispersive XRF spectrum


Transmittance of X-ray flux
As X-ray was transmitted through a sample, it becomes attenuated. The intensity ratio of the incoming X-ray and the outgoing
X-ray is proportional to the exponential of the absorption coefficient times the thickness. The spectrum will show a sudden
decrease in transmittance as the beam of scanning X-ray meets the absorption edge. The obtained transmittance spectra are
usually converted to absorption spectra afterward. X-ray Absorption is described by Equation 2.
I0
A = ln( ) (2)
I1

Disadvantage: This method is limited to moderately concentrated samples (eg. greater than 500 ppm). In cases of certain
samples or solvents, the incoming X-ray photons could be nearly completely absorbed, and leaves the detector little signal to
detect. For example, dichloromethane (CH2Cl2) is nearly opaque to low energy X-rays, and this will cause complication in
data interpretation if dichloromethane is largely present in the sample or used as the solvent.

Interpretation of XANES
Oxidation state sensitivity
As the oxidation-state of the absorption site increases, the absorption edge energy increases correspondingly. This observation
can be explained using an electrostatic model: atoms with a higher oxidation state require more energetic X-ray to excite its
core electron because the nucleus is less-shielded and carries a higher effective charge.
However, an alternative interpretation of edge energies is more suitable. This interpretation treats the edge features as
continuum resonances. A continuum resonance refers to a short lived the excitation process in which a core electron is excited
into a higher energy state that is usually above the continuum. An example is the potential well created by the absorbing and

6/25/2021 3 https://chem.libretexts.org/@go/page/1869
scattering (between nearest neighboring) atoms. As the absorber–scatterer distance gets shorter, the energy of the continuum
state increases as 1/R2. Since higher oxidation-states implies shorter bond lengths in molecules, the edge energies increases as
the oxidation-states increases.
As stated above, XANES is oxidation sensitive. Moreover, multiple scattering is particularly important in the XANES region.
In principle, one can argue that it is possible to determine the three-dimensional structure of the absorbing atom to its
environment from analysis of the XANES features. Experimentally, this has been proven true. The XANES region is quite
sensitive to small structural variations. For instance, two sites with identical EXAFS spectra can nevertheless have distinct
XANES spectra. Such sensitivity is intuitively, at least in part, due to the fact that geometrical differences between sites alter
the multiple scattering pathways, and thus the detailed structure in the immediate vicinity of the absorption edge.

Bound state transition


Weak pre-edge structures usually result from bound state transitions. The pre-edge structures prior to K edges of first row
transition metals arise from 1s to 3d transition. These pre-edge structures are observed for every first row transition metal as
long as its 3d orbital is not fully occupied. Although the 1s to 3d transition is forbidden by dipole selection rules, it is
nevertheless observed due to 3d to 4p orbital mixing and as well as direct quadrupolar coupling.
As the 3d to 4f mixing improves, the 1s to 3d transition increases, which means that such a trend can be utilized as tool to
probe the molecular geometric properties of the absorption sites. As the 1s to 3d transition increases, the geometry of the
absorption site distorts away from a centrosymmetric geometry.

References
1. PENNER-HAHN, J. E. X-ray Absorption Spectroscopy. University of Michigan, Ann Arbor, MI, USA
2. Cramer, S. P.; Tench, O.; Yocum, M.; George, G. N. Nucl. Instrum. Methods Phys. Res. Sect. A-Accel. Spectrom. Dect.
Assoc. Equip. 1988, 266, 586–591.
3. Que, L. Jr. (ED.). (2000). Physical Methods in Bioinorganic Chemistry : spectroscopy and magnetism. Sausalito, Calif. :
University Science Books, c2000.

Problems
Examining the X-ray absorption spectrum of a first row transition metal closely, one would notice that the "L edge" contains
three edges namely L1, L2 and L3 edge in a energy decreasing order. L3 edge intensity is twice as much as of the L2 and L3
intensity. Rationalize this observation.

Answer
The so called L1 edge corresponds to the excitation of a 2s electron which requires more energy than a 2p electron. The 2p
electron excitation is split into two edges namely L2 and L3, as a 2p electron gets excited, an open shell 2p5 electronic
configuration forms, consequently, spin–orbit coupling of such a system occurs. A 2p5 excited states corresponds to two
terms, 2P1/2 which has higher energy and give rise to the L2edge, and 2P3/2 which has lower energy and give rise to the L3
edge. Due to degeneracy, the L3edge has twice the edge jump of the L2and L1edges.

Contributors and Attributions


Mo Zhang

6/25/2021 4 https://chem.libretexts.org/@go/page/1869
XANES - Theory II
XANES, x-ray absorption near edge structure, is sometimes also called NEXAFS, near edge x-ray absorption fine structure. It
is defined as partial analysis of x-ray absorption, and the range of XANES is between the threshold and where EXAFS begins.
It can be used to determine the oxidation state and coordination number of the metal center in a complex; as well as the
covalency and site symmetry of the molecule.

Introduction
Absorption

M4 and M 5edges
3d
M2 and M 3edges
3p
M1 edge
3s

L3 edge
2p (J=3/2)
L2 edge
Energy

2p (J=1/2)

L1 edge
2s

K edge
1s

Figure 1. Illustration of a x-ray absorption spectrum. K edge, three L edges and five M edges are shown-figure form Wikipedia
XAS, x-ray abosorption spectroscopy, is a widely used technique for atomic local structure determination. A crystalline
monochromator is usd to select the certain radiation with correspond energy that can excite a core electron. And there are
mainly two region in the XAS spectrum, XANES and EXAFS.
The difference between XANES and EXAFS is loosely defined as space difference in terms of which regime they represent in
a XAS spectrum. Apparently the conceptual principles of these two systems are technically the same, which is one of the
limitations of this definition, because there is no clear principle definition to distinguish these two systems. In terms of energy
level, people usually think the XANES is in the range where the potential is 50 eV above the edge; the EXAFS is in where the
potential between 50 - 1000 eV above the edge.

NEXAFS EXAFS
(= XANES)
Absorption

0 100 200
E – EK (eV)
Figure 2. Schmatic illustation of XANES and EXAFS regime-figure from Wikipedia

6/29/2021 1 https://chem.libretexts.org/@go/page/1870
Based on this, Bianconi also suggested that when a core electron gets excited at XANES, the wavelength of this photoelectron
should be the distance between the absorbing atom and the closest atom/ligand.

Theory
During scanning, since the purpose is to excite a core electron, there will be no absorption until the energy of the radiation
matches the ionization energy of the core electron. Then a hole will be left in the inner shell. A electron from higher energy
state will fall into this vacancy, which cases releasing of energy. This released energy can either undergo a fluorescence
emission to release a photon, or it can further excite and eject an electron from a higher energy state. This ejected electron is
called auger electron.

(a) (b)
Vac

2p Valence Level
2s
1s
M, ...

L 2,3
L1
Electron collision

E Auger
2p
2s
1s

Auger electron emission

Figure 3. Two types of illustration of auger electron emission. (a) sequential step of auger electron emission. (b) illutration of
the same process by appling spectroscopic notation-figure form Wikipedia
Unlike traditional photoemission spectroscopy, in which the photoelectron is directly measured; however, in XANES, the
intensities of the light come through samples are measured. Besides, the effect of scattering from auger electrons,
photoelectrons, and even emitted photon are all included. In order to undstand how XANES works in the XAS and be able to
extract useful information from it, learning how the scattering happens and works in the x-ray absorption is very important.

Absorption edge
XANES in spectrum includes three portions: pre-edge, which is stroingly affected by bond lengths of molecules because of
the exponentially decay of wavefunctions, edge, which is the big jump at 0 eV that indicate the ionization energy of core
electrons, and XANES, which is the region from 0 to 50 eV above the threshold. As Figure 2 shows, the binding energy
regime of a core shell is scanned by the radiation at certain frequency correspond to certain energy. There will be no
absorption until the energy of the radiation matches the ionization energy of one core electron. There will be a big jump in the
spectrum. What need to be notice here is that there is no selection rule for transition, because when the photon hit the core
electron, it not only excites the electron, but also give kinetic energy to that electron to form a photoelectron. This
photoelectron will be ejected and leave the atom. The most common edges are shown in Figure 1, K-edge (1s), L-edge (2s and
2p), M-edge (3s, 3p and 3d). These edges are corresponding to the priciple quantum number, n=1 (K-edge), n=2 (L-edge) and
n=3 (M-edge), which indicate the order of electron shells. K, L and M are all x-ray notation.

Scattering
From the other perspective, the x-ray absorption spectroscopy can be used to determine the coordination number of a metal
center and structure of a molecule based on how the scattering of photoelectrons or auger electrons affect the intensity of the
radiation that come through the sample. When energy of the photon is high, in EXAFS, at continuum states which are high in
energy, the atom can only have very weak effect on the system because the photoelectron that is ejected from the core orbitals
is only weakly scattered by the surrounding atoms which are further from the absorbed atom. So all the electronic properties,
or the interest in electron structure of the absorbed atom are centered on the low-lying extended states.
The theoretical analysis of the electron structure theory is to solve Schrodinger equation. Based on scattering theory, one can
say that at high energy, EXAFS region, for the scattering of the photoelectron or auger electron are very weak, the main
contribution of the scattering to the wave function of the final state is from the path that the excited electron is scattered only

6/29/2021 2 https://chem.libretexts.org/@go/page/1870
once. This process can be referred as single scattering. Since when the excited electron is scattering at one atom, there must be
another scattering happening on the other atom on the opposite position, so the scattering in EXAFS can be treated as
geometrical and the information is much easier to extract.
In the XANES region, when the energy is low, since neighboring atoms are not only close to the center atom, but also
relatively close to each other, the excited electrons tend to bounce between the neighboring atoms before it hits back to the
absorbed atom. So people call it multiple scattering in XANES regime compare to the single scattering in the EXAFS regime.
The information details between the absorbing atoms and its neighbors, in terms of spatial arrangement, such as, radial
distance, bond angles, and orientations relative to one another, are highly based on the multiple scattering. The multiple
scattering expression is shown below,
ij i ij i ij j ij i ik k kj j
τ ′
=t δ δLL′ + t G ′
t ′ (1 − δ )+∑t G ′′
t ′′
G ′′ ′
t ′
+⋯ (1)
LL l l LL l l LL l L L l
′′
k,L

L stands for the pair of angular momentum quantum number. t


i
l
is the l-wave t-matrix of atom i given in terms of the phase
shift δ by
l
i

i
i 2iδl
t (ϵ) = i/2k[ e (ϵ) − 1] (2)
l

ij
and G LL

is a real space structure constant.
Each term in the expression represent an indvidual scattering path. This XANES cross-section is expressed as the sum of the
sum of individual scattering path.

Figure 4. Illustration of single scattering in EXAFS (left) and multiple scattering in XANES (right)-figure from Wikipedia
The scattering of excited electrons in XANES regime is very sensitive. Changing of charge distribution leads to changing of
the chemical environment of the absorbed atom. And the absorption edge will shift in the XANES regime because the core-
level energies are changed. Every molecule have their distinct charge distributions, and this is how one is able to figure out
their structure and other spatial information based on XANES. However, there are a lot of physical effects evolved in the near-
edge region, which may cause some problems when trying to extract specific information out of it.

6/29/2021 3 https://chem.libretexts.org/@go/page/1870
Figure 5. Illustration of different oxdation states provides different x-ray absorption spectra in XANES regime, which due to
different chemical environment.-figure from Wikipedia

One-electron approximation
When there are more than one electron in a system,for example, in a hydrogen molecule, the interelectron term in the
Hamiltonian is shown as below
2 2 2 2 2 2 2 2
ℏ ℏ e e e e e e
2 2
H =− ∇ − ∇ − − − − + + (3)
1 2
2m 2m 4πϵ0 ra1 4πϵ0 ra1 4πϵ0 rb1 4πϵ0 rb2 4π ϵ0 R 4πϵ0 r12

the last six terms, which are expression of potential energy of the four particles (two protons and two electrons) in the system,
make impossibility to solve the Schrodinger equation. Then an approximation is made that instead of consider all these
potential energy, the Haniltonian is expressed as the one electron which contains potential energy as the average of all the
interactions.
fi ϕi = ϵi ϕi (4)

where f is a sum of one-electron function, and ϕ is a one-electron wavefunction.


i i

This one-electron approximation, which is defined as the product of wave functions of many particles can be represented by a
wave function of a single particle, is employed in the XANES theory. Even though density functional theory with local density
approximation (LDA), which includes the exchange-correlation effect, can be used to figure out the electronic structure, it can
be only applied in the ground state. In excited state, people need a quasiparticle description, and apparently one-electron state
can give a pretty good approximation to the quasiparticle states. One thing about the one-electron eigenvalues in density
functional theory is that instead of being the quasiparticle energies, they actually help the central object to obtain the correct
total energy. From this, people know that excitation energy of local density approximationone-electron state cannot be
determined, which, in XANES, means that LDA core eigenvalue can not accurately determine the core electron’s binding
energy. However, the main spectra feature of interest in XANES regime is in the relative energy region above the threshold,
instead of the threshold energy. So calculating the x-ray absorption spectra by using LDA one-electron eiganstate can give a
pretty good approximation.

Advantages and Disadvantages


The most significant advantage of XANES, which also refers to as XAS, is that this technique is specific to elements. The
signals in a XANES spectrum, llike fingerprints, represent only specific elements. Metal ions in molecules may be "silent" to
some of spectroscopic techniques, such as EPR; however, XAS spectra can be taken of samples with a metal of interest, it
provides detailed information about the oxidation state of the metal atoms. The sensitivity to multiple scattering makes it
possible to extract three dimensional structure information from XANES spectra. Even though XANES can be used to
determine local structure and many characteristics of atoms in a molecule, no technique is perfect. XANES also has its

6/29/2021 4 https://chem.libretexts.org/@go/page/1870
drawbacks. X-radiaiton can be destructive to samples, causing damage to the sample during the measurement process. In XAS
spectra, espeically in XANES regime, oxidation state and the identity of scattering atom are difficult to determine because of
the effect of mutiple scatterings by photoelectrons and auger electrons.

References
1. Winefordner, L. D.(ED.) (1988). X-Ray Absorption. Chem. anal. 1988 92 53-83.
2. Bianconi, A. Appl. Surf. Sci. 1980, 6, 392

Contributors and Attributions


Shuai Wang

6/29/2021 5 https://chem.libretexts.org/@go/page/1870
XAS - Theory
XAS, or X-ray Absorption Spectroscopy, is a broadly used method to investigate atomic local structure as well as electronic
states. Very generally, an X-ray strikes an atom and excites a core electron that can either be promoted to an unoccupied level, or
ejected from the atom. Both of these processes will create a core hole. If the electron dissociates, this produces an excited ion as
well as photoelectron and is studied by X-ray Photoelectron Spectroscopy (XPS).

Introduction
The electrons that are excited are typically from the 1s or 2p shell, so the energies are on the order of thousands of electron volts.
XAS therefore requires high-energy X-ray excitation, which occurs at synchrotron facilities. X-ray energy is about 104 eV
(where "soft x-rays" are between 100 eV- 3 keV and "hard x-rays" are above 3 keV) corresponding to wavelengths around 1
Angstrom. This wavelength is on the same order of magnitude as atom-atom separation in molecular structures, so XAS is a
useful tool to deduce local structure of atoms. XAS is also utilized in analyzing materials based on their characteristic X-ray
absorption "fingerprints." It is possible to deduce local atomic environments of each separate type of atom in a compound. XAS
is particularly convenient because it is a non-destructive method to examine samples directly. Structures can be determined from
samples that are both heterogeneous and amorphous.

When an X-ray strikes an atom, one of the core electrons is either excited to a higher energy unoccupied state (a transition
studied by XAS) or into an unbound state, called the continuum. When electrons are ejected from an atom of a solid material,
this is essentially the photoelectric effect and is studied by X-ray Photoelectron Spectroscopy. Below is a diagram illustrating an
atom absorbing an x-ray with the resultant ejection of a core electron into the continuum. Note that the level K refers to the n=1
level, L refers to the n=2 level, and M refers to the n=3 level.

The wave vector of a photoelectron


Determining the wave vector of the electron dictates which subset of XAS will be used (EXAFS, NEXAFS, etc.). The kinetic
energy E and threshold energy E (amount of energy the photoelectron needed to promote the electron into the continuum) of
k 0

the photoelectron can be addressed, after considering the de Broglie equation:


h
λ = (1)
p

where
p is the momentum of the electron,
λ is the wavelength, and
h is Planck’s constant.

The wave vector of the electron is denoted as k, and can be defined as


2π 2πp
k = = (2)
λ h

and thus the electron kinetic energy is written


2 2 2
1 1 p h k
2
Ek = mv = ( ) =( ) (3)
2 2 m 2π 2m

5/28/2021 1 CC-BY https://chem.libretexts.org/@go/page/1871


In terms of the incident x-ray energy, E , and the threshold energy E ,
0

Ek = E − E0 = hν − E0 (4)

The expression for the wave vector k of the photoelectron can thus be written as
−−−−−−−−−−−−−−−−−
2

k = √( ) 2m (hv − Eo ) (5)
h

Depending on values of k, different subsets of XAS will be used. For example, if k = 0 , or if 0 < k < 2/R (where R is the
distance between the x-ray absorbing atom and it’s nearest neighbor), low-energy Near Edge X-ray Absorption Fine Structure
(NEXAFS) and NEXAFS will be used, respectively. If k > k , where k is equal to 2/R, Extended X-ray Absorption Fine
c c

Structure (EXAFS) is used. It is apparent from the diagram below that there are three different types of excited photoelectrons
upon absorption of an x-ray quantum. For the sake of completeness, these are briefly discussed below.
The first type of photoelectron transition to an unoccupied valence state, as it does not have enough energy to completely leave
the atom. The second type of photoelectron actually has just enough kinetic energy to be able to escape into the continuum.
Multiple scattering processes occur here, between multiple surrounding atoms that neighbor the absorbing atom. Lastly, the third
type of photoelectron has a very high kinetic energy, and weak back-scattering occurs in a single scattering process between only
one neighbor atom.

Returning to the formula for the kinetic energy of an electron

Ek = E − E0 = hv − E0 (6)

can be rewritten as

Ek = hv − EB − ϕ (7)

Here, E is the binding energy of the core electron to the atom, while the expression E + ϕ is the ionization energy. A more
B B

in-depth analysis of the ionization energy can be found in the section regarding Koopman's theorem.

Beer’s Law and the relation to XAS


Beer's Law can be applied to XAS for an x-ray beam that is both narrow and monochromatic, striking the absorber at a 90
degree angle. The absorber has a uniform composition and thickness. The important result of this derivation will be an
expression describing the absorption coefficient, μ , of X-ray absorption. Beginning with the basic model of
M

5/28/2021 2 CC-BY https://chem.libretexts.org/@go/page/1871


Beer’s law can be written
I1
log( ) = kΔm = kd Δd (8)
I2

where m and d refer to the mass and thickness of the sample, respectively. The terms k and k are proportionality constants
d

whose units are dependent on the units of m or d, the wavelength, and also the elements comprising the sample. The previous
equation can be written in terms of natural logarithms, in which case
I1
log( ) = μM Δm = μM ρΔd (9)
I2

where the symbol μ has been introduced, and is termed the mass absorption coefficient. Units of μ are cm2/g, and amounts
M M

are not dependent on the physical or chemical state of the element as X-ray absorption is primarily an atomic property. The term
2
Δm is the mass difference of two samples in units of g/cm of irradiated sample area, and Δd is the thickness of the sample in

units of centimeters. Density of the sample is ρ in units of g/cm3. This equation is limited to the sample area receiving uniform
radiation by the X-ray beam. In thickness measurements, the linear absorption coefficient ?l is introduced, and the previous
equation can be written
l1
ln( ) = μb Δd (10)
l2

The linear absorption coefficient has a characteristic value depending on both the element and the quantity of atoms that are in
the beam path. The value of Δd is the thickness of the sample that is irradiated in units of centimeters. It is obvious that

μ1 = μM ρ (11)

Furthermore, one can define the atomic mass absorption coefficient that measures the “cross-section” of the absorbing atom as
M
μ = μM (12)
N

where M is the absorbing atom’s atomic mass and N is Avogadro’s number. Values of μ M for a sample can be calculated using
μM = WA μA + WB μB + WC μC + … (13)

where the sample contains elements A, B, and C of weight fractions W , W , and W , and mass absorption coefficients of μ ,
A B C A

μ , and μ . Values of μ , μ , and μ


B C A B C at different wavelengths are available in many sources. As stated previously, the value of
the mass absorption coefficient is primarily an atomic property. It is not dependent on the state of the atom; for instance, a
bromine atom in the form of vapor, potassium bromide or sodium bromate, liquid or solid bromide, has the same chance of
absorbing an x-ray quantum in all of these forms.

More on the absorption coefficient, μ M

At higher photon energies, μ decreases steadily. However, when the necessary energy for a core electron transition is achieved,
M

μM increases sharply and causes an absorption edge. This occurs when the photon energy just matches the energy needed to
either promote a core electron into an empty valence level, or to completely eject the electron from the atom (electron is
promoted into the continuum). Each edge occurs at its own critical absorption wavelength. The associated energies are the
electron binding energy in the K, L, and M… shells (based on the Bohr atomic model, n = 1 for K edges, n = 2 for L edges,
n = 3 for M etc.) of the particular atom absorbing the X-rays.

5/28/2021 3 CC-BY https://chem.libretexts.org/@go/page/1871


The resultant spectrum of a sample will show these edges at the X-ray photon energies that equal the ionization potentials of the
bound electrons in the component atoms. Seen in the figure below are the K-edge and three L-edges of a typical X-ray
absorption spectrum. This nomenclature indicates which core orbital the electron originated from. The K-edge is a consequence
of the 1s – 3p transition, where the L-edges are from 2s – 5p (L ), 2p – 5d (L ), and 2p – 5d
I 1/2 3/2 II 3/2
(L ) transitions.
5/2 III

Elements have characteristic edges, a few of which can be seen in Table 1.

There are a few factors that directly influence the exact energy of the edge. The charge density of the absorbing atom is very
important, and is itself influenced by the valence or oxidation state. The energy of the X-ray absorbing edge is proportional to
the oxidation state of the absorbing atom; the edge will occur at a higher energy the more positive the oxidation state of the
atom. This can be simply explained by considering that it is increasingly difficult to remove an electron from an atom that bears
a higher positive charge.
Another factor influencing the edge energy is the atomic number of the absorbing atom. As atomic number increases, so does the
corresponding edge energy. A plot of the energy of the edge as a function of atomic number can be seen below. The X-ray
photon energy region of approximately 2-30 keV limits the accessible K edges. Because current synchrotron sources do not
produce high intensity X-ray energies greater than 30 keV, elements P to Sn can be analyzed, but anything past this cannot. The
energy required to separate a 1s electron from elements of atomic numbers greater than Sn is too high for current synchrotron
sources. However, because L edges occur at lower energies than K edges, it is possible to examine the rest of the elements using
XAS. Lr, the heaviest element, has an L edge occurring at 22 keV, an energy that synchrotrons can indeed produce.
III

5/28/2021 4 CC-BY https://chem.libretexts.org/@go/page/1871


Koopman's Theorem and the relation to x-ray absorption
As stated previously, x-ray absorption is capable of exciting and dissociating a core electron from a neutral atom. This aspect of
x-ray absorption is studied by XPS, but we will discuss this here for completeness. Calculating the energy for this ionization
potential is of interest. Koopman’s Theorem (KT) is a method that relates experimental ionization potentials to the molecular
orbital energy levels. Considering an orbital ?i of energy E , Koopman’s Theorem says that the value of – E is equal to the
i i

ionization potential needed to remove an electron from ?i. Stated differently, the ionization potential is the negative value of the
HOMO energy. Koopman’s Theoroem is an application of the Hartree-Fock approximation, as the value of E is calculated via i

the HF approximation. This is the simplest method to evaluate ionization potentials.


Koopman’s theorem assumes that a multi-electron atom has an electronic wavefunction which can be described using a Slater
determinant. The Slater determinant is a set of one-electron wavefunctions, where each wavefunction in the determinant is the
eigenfunction of the related Fock operator. Another assumption made is that when an electron is added or removed from the
system, the remaining electrons’ corresponding Fock operators do not change. Thus the Koopman’s theorem predicted value of a
final wavefunction has a higher energy than the actual final wavefunction. This is because in reality an addition or removal of an
electron to or from the initial wavefunction will change the system’s Fock operator. This would result in a re-organizing of the
one-electron wavefunctions, which Koopman’s theorem disregards.
Koopman’s theorem uses the Hartree-Fock theory in describing the value of E , where the Hartee-Fock theory describes
i

ionization or electron-gain of a system. A more accurate method would be electron correlation, which is typically calculated by
post Hartree-Fock methods.

A review of the Hartree-Fock Approximation


Assume that a single Slater determinant can describe the closed-shell system having N electrons. The Slater determinant, |ψ ⟩, i

can be written as a set of N orthonormal spin orbitals, χ , χ ,…χ .


1 2 N

| ψ0 ⟩ = | χ1 , χ2 , … , χN ⟩ (14)

Recall that the Hartree-Fock method seeks to find the spin orbital Y that minimizes E , using the electrostatic Hamiltonian, H,
0 0

operating on |ψ ⟩.
0

N N N
1
E0 = ⟨ψ0 |H | ψ0 ⟩ = ∑⟨i|h|i⟩ + ∑ ∑{[ii|jj] − [ijβj]} (15)
2
i=1 i=1 j=1

where

⟨i|h|i⟩ = ⟨χi |h| χj ⟩ = ∫ dx1 χ1 ∗ (x1 )h(x1 )χj (x1 ) (16)

1
∗ ∗
[ij|kl] = [ χi χj | χk χl ] = ∫ dx1 dx2 χ (x1 )χj (x1 ) χ (x2 )χl (x2 ) (17)
i k
r12

5/28/2021 5 CC-BY https://chem.libretexts.org/@go/page/1871


and h(x ) is the usual one-electron core-Hamiltonian. Based on the minimization condition on E , the following expression can
1 0

be recognized, where the occupied spin orbitals ? span a subspace of the Fock operator, f .
N

f | χn ⟩ = ∑ εω | χ3 ⟩ (18)

j=1

This holds true for any a from 1,2,3….N. Also, the Fock operator matrix elements between occupied spin orbitals and virtual
spin orbitals must be zero. If there is a set of N spin orbitals that satisfies the previous equation, and which has a unitary
transformation performed on it, then a new set of spin orbitals will be produced that satisfy the same equation. The unitary
transformation is chosen such that the matrix will have diagonal elements of
ϵij = ⟨χi |f | χj ⟩ (19)

The corresponding spin orbitals are referred to as "canonical spin orbitals." Also, the Fock operator is defined as
N

⟨χi |f | χj ⟩ = ⟨χi |h| χj ⟩ + ∑ [ij|kk] − [ik|kj] (20)

k=1

Evaluating the Ionization Potential


The energy needed to remove an electron from an atom can now be evaluated. To get a close approximation of the ionization
potential, the neutral and ionic states require a good description. The neutral state is said to be at the HF level. A rough
assumption can be made with respect to the electron density of the ionic system. The ionic system has an electron density of the
neutral system minus the density of the orbital from which the electron has been removed. A “frozen orbital” approximation has
been used here, in that it is assumed the other N-1 electrons do not change their spatial distributions upon ionization. Thus the
ionic molecule can be described using the neutral molecule’s Hartree-Fock spin orbitals.
To summarize thus far,
1. The value of E is the HF energy of an N-electron system.
0

2. The HF canonical spin orbitals of the N-1 electron system are used to build up the mean value of H over an N-1 electron
single determinant.
Koopman's theorem thus says that the difference between E and the mean value of H is a good approximation of the ionization
0

potential of the corresponding N-electron system. If the ionization system is described by the N-1 -electron single determinant,
and the canonical HF spin orbitals of the neutral system are ?χ , then it is possible to write an expression for the ionization
i

potential as
(N −1) N (N −1) (N −1) N N
−ϵcc = Ec − E0 = ⟨ Ψc |H | Ψc ⟩ − ⟨ Ψ0 |H | Ψ0 ⟩ (21)

where
N −1
| Ψc ⟩ = | χ1 χ2 , … , χc−1 , χc+1 , … , χN ⟩ (22)

The matrix element ϵ is an eigenvalue of the Fock matrix if one uses the canonical spin orbitals in the calculation. Koopman
cc

showed that the use of the canonical spin orbitals as HF spin orbitals produce the lowest energy of the ion. Thus it is possible to
approximate the ionization potentials by the eigenvalues of the Fock matrix.

Fermi's Golden Rule applied to XAS


Analysis of both the initial and final states of the absorbing atom is necessary to determine the probability of a core electron to
absorb an x-ray photon. XAS measures the transition of an electron initially in a deep core state and finally in a previously
unoccupied state. Fermi’s golden rule gives the transition probability, to which the x-ray absorption coefficient is proportional.
This relationship can be expressed as
2
μ(E) ∼ ∑ |⟨ψi |A(r) ⋅ p| ψf ⟩| δ(E − Ef ) (23)

where the photoelectron energy is E=?? – E . ψ and ψ are the initial and final eigenstates and have energies E and E . The
i i f i f

initial state is the ground state of the atom. These wavefunctions are typically calculated using a Self-Consistent Field (SCF)

5/28/2021 6 CC-BY https://chem.libretexts.org/@go/page/1871


approximation. The coupling to the x-ray field is represented by A-p, where p is the momentum operator. A(r) is the vector
potential of the applied electromagnetic field, and is considered to be a classical wave of polarization. This can be seen as ϵ^ ⊥ k
or A(r, t) ≅ϵ^A e . The entire expression is summed over unoccupied final states of energies E .
0
ik⋅r
f

Usually the Golden Rule is reduced to just a one-electron approximation, and calculations are based on the dipole
approximation. The dipole approximation assumes that the spatial dependence of the electromagnetic field can be ignored, or
= 1 . However, the one-electron final state that should be used is still debated. The final state rule is often implemented in
ik⋅r
e

current research. In this approximation, the final state is calculated while considering the screened core-hole. The Hamiltonian
used is the final state, one-particle Hamiltonian.
However, for transition metal L2,3 edges (excitation of a 2p electron to a 3d shell), the one-electron approximation is rendered
invalid. This is a result of the initial state wave function containing a partially filled d-shell. Upon excitation of a 2p electron to a
3d shell, there are two partially filled shells that have large overlap. Mathematically, this overlap can be written as
2 2 ∗ 5 n+1 ∗ n 2 5 n+1 n 2
|⟨ψf |A(r) ⋅ p| ψi ⟩| = |⟨ψi 2p3d|A(r) ⋅ p| ψi ⟩| = |⟨ψ 2 p 3 d |A(r) ⋅ p| ψ 3 d ⟩| = |⟨2 p 3 d |A(r) ⋅ p|3 d ⟩| (24)
i i

Note that the wave function for the final state is written as ψ i 2p3d .
A little more on the "Interaction Hamiltonian"
The overall interaction Hamiltonian can be written as a sum of the radiation field Hamiltonian, the atomic electron Hamiltonian,
and the interaction Hamiltonian.
H = HRAD + HAT OM + HIN T (25)

where the radiation field Hamiltonian, HRAD, can be written

HRAD = ∑ ℏωk (nk,λ + 1/2) (26)

k,λ

where the whole expression is summed over the wave vector, k, and degrees of freedom, λ . The term in parenthesis is simply the
zero point energy. The kinetic term p /2m and potential energy term V (r ) together make up the expression for the Hamiltonian
2
i

of the atomic electron. The potential energy term considers both the Coulombic interaction with the nucleus, as well as the
electron-electron repulsion and spin-orbit interaction.
2
p
i
HAT OM = ∑ [ + V (ri )] (27)
2m
i

Lastly, the interaction Hamiltonian is described as a slight perturbation comprised of two terms. The first of these terms shows
the vector field A interacting on the momentum operator p. Another way of stating this is the electron moments being acted on
by the electric field E. The second term in the interaction Hamiltonian shows how the magnetic field B, where B=? × A acts on
the electron spin ?.
e e
HIN T (l) = ∑ p)i ⋅ A(ri ) + ∑ σi ⋅ ∇xA(ri ) (28)
mc mc
i i

So Fermi's Golden Rule gives the transition probability, W , between a transition between a system's initial state ( ?i) and final
fi

state (?f) upon absorption of a photon of energy ℏΩ. This is expressed in the equation
2π 2
Wf i = |⟨ϕf |T | ϕi ⟩| ∂(Ef − Ei − ℏΩ) (29)

The role of the delta function is to take care of the conservation of energy. A transition occurs if the energy of the final state is
equal to the energy of the initial state plus the energy from the x-ray. The transition rate is given by the squared matrix. The
Lippman-Schwinger equation gives an expression for the transition operator, T, as
1
T = HIN T + HIN T T (30)
Ei − H + iΓ/2

where H is the Hamiltonian of the unperturbed system, and gamma stands for the excited state's lifetime broadening. In order to
describe the one-phone process of x-ray absorption, the equation for T is solved iteratively and in first order where
T =H
1 (1) . For an electric dipole transition, the transition operator is expressed as
IN T

5/28/2021 7 CC-BY https://chem.libretexts.org/@go/page/1871


−−−−−
2πℏΩ
T1 = ∑ e√ eq ⋅ r (31)
V5
q

and for an electric quadrupole transition, the transition operator is described as


^
T1 (EQ) ∝ ek,λ ⋅ Q ⋅ k (32)

where Q is the quadrupole operator


1
2
Q = rr − r ∂ij (33)
3

Dipole Selection Rules


As stated previously, the dipole transition is able to describe x-ray absorption. When Fermi's golden rule is written in terms of
this operator, the probability of a transition, W , per unit time is
fi

2 3
e 4Ω 2
Wℏ = n|⟨ϕf |r| ϕi ⟩| ∂ (Ef − Ei − hΩ) (34)
2
ℏc 3c

where the fine structure constant is the first term, . The symbol omega is the excitation frequency and n is the number of
e

ℏc

photons of the radiation field.The wave functions of an atom can be given quantum numbers J and M, where J is the overall
momentum quantum number and M is the magnetic quantum number. If the integral term is written in terms of J and M, the
following equation arises, where the radial and angular part of the matrix element can be separated as stated by the Wigner-
Eckart theorem.

′ J−M
J 1 J ′
⟨ϕf (JM )| ⋅ r| ϕi (J M )⟩ = (−1 ) [ ] ⟨ϕf (J)| ⋅ r| ϕi (J )⟩ (35)
qλ q
−M q M

The total momentum quantum number, J, cannot change by more than one unit. Therefore, ΔJ=+1, 0 or -1. The magnetic
quantum number, M, can change depending on the x-ray polarization. This can be stated as ΔM=q. The incident x-ray has an
angular momentum quantum number l value of lhv=1, and conservation of this angular momentum yields Δl=+1 or -1. The
excited electron has an l value different from the original core state by 1. The spin quantum number, s, has the selection rule of
Δsj=0, as the x-ray has no associated spin and this must be conserved. When an electron in the 1s core state is excited, the only

state that is accessible is the p state. However, from the p state, either the s or d final states are accessible.
The line strength of the transition can be written in terms of the radial part of the previous equation, or as
2 ′ 2
S =e ⋅ |⟨ϕf (J)| eq ⋅ r| ϕi (J )⟩| (36)

.
2 2
S =e ⋅∣
∣⟨ϕf (J) | eq ⋅ r| ϕi (Jη ) ⟩ (37)

Then the transition probability, W , can be rewritten to include the S term.


ij

3 ′ 2
1 4Ω J 1 J
WA = n[ ] S∂ (Ef − Ei − hΩ) (38)
2
ℏc 3c −M q M

If the squared-J symbol is assumed to be unity, this can be written as


3
1 4Ω
Wn = nS∂ (Ef − Ei − hΩ) . (39)
2
ℏc 3c

Photon flux, the x-ray absorption cross-section, and the oscillator strength
As seen in the previous equation, the transition probability is actually proportional to n, then number of photons, which is
directly related to photon flux. The formula for photon flux F can be written as p

2
Ω
Fp = n (40)
2
πℏc

5/28/2021 8 CC-BY https://chem.libretexts.org/@go/page/1871


The x-ray absorption cross section, σ, is given in m2 in the following equation, and is directly proportional to the value of f
which is the oscillator strength.
2
Wf i 4π Ω
σ = = S∂(Ef = Ei − ℏΩ)
Fp 3c

2 2
2π e
σ = f
mc

The penetration depth, λ , of the x-rays is directly influenced by the cross section via
p

1
λp = (41)
ρσ

with the term ρ being the density of the system. Here, the units of cross-section are typically given in values of angstroms-
squared. Also, it is common to define the term "inverse density", or the space occupied by one atom of interest, as
Vat = 1/ρ (42)

and the equation for penetration depth can be re-written as


Vnt
λp = (43)
σ

Practice Problem
1. For La metal, which has a V value of cubic angstroms, and a cross section of 0.15 square angstroms, what is the penetration
at

depth?
2. What is the difference between XAS and XPS?
3. What information does the absorption edge yield?

References
1. Liebhafsky, H., et al. X-rays, Electrons, and Analytical Chemistry. John Wiley & Sons. 1972.
2. Que, L. Physical Methods in Bioinorganic Chemistry. Sasuolito, California: University Science Books, 2000.
3. Skoog, D.; Holler, F.; Nieman, T. Principles of Instrumental Analysis. Fifth ed.; Brooks/Cole: 1998.
4. Ankudinov, A.L., Rehr, J.J., "New Developments in the Theory of X-ray Absorption and Core Photoemission." Journal of
Electron Spectroscopy and Related Phenonemon. 2001, 114, 1115-1121
5. Rehr, J.J., "Theory and Calculations of X-ray Spectra: XAS, XES, XRS, and NRIXS." Radiation Physics and Chemistry.
2006, 75, 1547-1558
6. Albers, R.C., Rehr, J.J., "Theoretical Approaches to X-ray Absorption Fine Structure." Reviews of Modern Physics. 2000,
72, 621-654
7. De Groot, F., Kotani, A. Core Level Spectroscopy of Solids. CRC Press: 2008.

Contributors and Attributions


Sarah Gee (UCD)

5/28/2021 9 CC-BY https://chem.libretexts.org/@go/page/1871


Photoacoustic Spectroscopy
Photoacoustic spectroscopy (PAS) uses acoustic waves produced from materials which are exposed to light to measure its
concentration. PAS is unique in that it combines heat measurements with optical microscopy. Gases have been the ideal
samples used but more research has been increasing gradually to use PAS efficiently for solid and liquid samples. When
measuring a sample, it takes measurements directly through looking at the internal heat instead of the effects of the light on the
surroundings. This makes PAS highly accurate and useful for sensitive detectors. Interest sparked after Alexandar Graham Bell
wrote about his findings when he discovered the acoustic effect in 1880.
Bell accidently stumbled onto this effect as he was experimenting with this invention of the photophone. He noticed that a
clear acoustic sound formed whenever the sunlight hitting the sample was interrupted. Bell realized that the absorption of light
by the material caused the sound wave which is now known as the photoacoustic effect. The ultraviolet and infrared spectra
also worked and experimented by Bell. However, the apparatus was not sophisticated enough to show any promise in accurate
results and the development of PAS was put on a halt. It was not until the introduction of more sophisticated equipment did the
development of PAS start again. Today, most set ups use other sources than sunlight and microphones instead of the ear to
measure the waves emitted from a material accurately.

Theory
Generally, when a material absorbs light, there are many paths the energy can go on. Light is always conserved as shown by
the equation,
1 +α +T +R (1)

where is \( \alpha\) the absorbance, T is the transmittance, and R is the reflectance. Light that hits the sample must either be
absorbed, transmit through the material, or reflect off of the material. PAS focuses on the light path that is absorbed as that is
where heat is released. As light strikes the sample, the photons are absorbed and electrons are excited from the energy created.
This energy is then released as heat and as the heat expands, acoustic waves are formed. This process is explained below and
shown in figure 1.

Figure 1. Process diagram for light absorption to acoustic waves.


As light is absorbed electrons are excited either electronically or vibrationally. When looking at electronic excitation, electrons
jump to a higher energy level. As they drop back to its ground state, the extra energy is given off as heat. Collision
deactivation, another form of heat formation, involves the colliding of atoms. The collision of atoms give off energy in the
form of heat. However in the case of electronic excitation, the energy can also be dissipated through chemical reactions or
radiative emissions as seen in figure 1. Chemical reactions involve any reactions with its surroundings as energy is used to
initiate those reactions. Radiative emissions involve the energy given off as photons, rendering it useless for PAS which
requires heat. This reduces the amount of heat formed as energy is spent somewhere else. It is possible to have chemical
reactions form heat, but only a portion of the energy absorbed goes towards heat. On the other hand, with vibrational energy,
chemical reactions and radiative emissions have little effect. The lifetimes of the vibrations are long enough prevent chemical
reactions and radiative emissions from interfering. Therefore, the atoms have as much time as needed to complete the process
of collision deactivation which will effectively use the full amount of energy to transfer to heat.

5/9/2021 1 https://chem.libretexts.org/@go/page/40643
Figure 2. Depiction of collision deactivation. Kinetic energy of the atoms is transferred into heat(red arrows).

With the formation of heat, thermal expansion also occurs. The expansion of heat creates localized pressure waves and in turn,
can be measured as an acoustic wave. However, as with the case of the formation of energy, heat can also be lost through the
surroundings. Heat diffusion lowers the temperature around the emitted energy source which in turn lessens the pressure fields.
With acoustic waves sent after every pulse of light, a sensor can then measure the waves. By adjusting the wavelength of each
pulse of light, the corresponding acoustic wave can be measured and plotted to form a spectrum of the material.

Figure 3. Inside components of photoacoustic spectrometer.


With the advancement in technology, amplifiers, light sources, and sensors have significantly improved. Figure 3 shows a
common set up of the inside of a photoacoustic spectrometer. Light sources typically use infrared lasers or wire filaments such
as tungsten that produce high intensities of light. To send pulses of light to hit the sample, the light source is either turned off
and on to produce the pulsing effect or a rotating disc with openings to control the pulses of light going through. A mirror
directs the pulses of light to hit a set of filters, which can change to alter the wavelength of the light hitting the sample.
Once the light passes through the filter, it hits the contact window, which is where the sample is held. Two microphones are
placed inside to pick up the acoustic waves and is sent to measure its electrical signal. Different wavelengths are tested and a
spectrum for the sample is created.

Importance of Photoacoustic Spectrometry


Differences in PAS
Though PAS may seem similar to other infrared techniques such as Fourier transform infrared spectroscopy (FTIR), it has
many unique aspects. PAS does not measure the effect relative to the background but directly from the sample, making it
extremely accurate. Samples with multiples gases can be singled out and measured. Samples can also be in tiny amounts as the
sample case is small. However, it can currently only measure samples as small as a few milliliters, which is still small
compared to other techniques.
Applications
Because of the high sensitivity and accuracy of PAS, it is ideal to use it in gas detectors. Gas levels in the atmosphere can be
measured and provide details on any dangers of rising toxic gases. It is also useful in determining the materials of an unknown
samples. Each material has its own unique spectrum and by observing the acoustic waves produces, one can match the waves

5/9/2021 2 https://chem.libretexts.org/@go/page/40643
to specific profiles of materials. PAS is also used for high resolution imaging by analyzing the topography of the sample.
Using the topography and the profiles of electric signals, one can create an image with shaded colors to indicate the different
materials. The cost of creating these devices have decreased and is gradually being used more widely in gas detectors and
sample analysis in labs.

Questions
1. Light hits a sample of aluminum. The data collected shows that the sample had a transmittance 0.12 and reflectance value
of 0.8196. Assuming light is conserved, what is the absorbance of the material?
2. Does vibrational or electronic excitation of electrons produce more heat? Why is that?
3. Why does PAS require pulses of light instead of a continuous steady source of light to hit the sample?

Answers
1. Using equation 1 stated above, light must be conserved and so 1-0.12-0.8196= α
α = 0.0604
2. Vibrational excitation of electrons produce better heat as all the energy from the absorbed light transfers into heat. This is
because the vibrational lifetimes of the atoms is long enough so that the transfer of energy to heat is not interrupted by other
processes.
3. To obtain a signal, it must come in the form of a wave. Having a constant source of light will prevent waves from forming
and the spectrum will only show up as a straight line. The point of the pulsing light is to have the material absorb the energy,
convert it to heat, send out the acoustic waves, and let it rest before allowing it to absorb another amount of energy.

References
1. Hummel, Rolf E. Electronic Properties of Materials: An Introduction for Engineers. Berlin: Springer-Verlag, 1985. Print.
2. Miklos, Andras, Stefan Schafer, and Peter Hess. "Photoacoustic Spectroscopy, Theory."AccessScience (n.d.): n. pag.
Academic Press. Web. 24 Nov. 2015.
3. Svelto, Orazio. "Nonradiative Decay and Energy Transfer." Principles of Lasers. 4th ed. New York: Plenum, 1976. 50-56.
Print
4. Zharov, V. P., and V. S. Letokhov. "Optoacoustic Spectroscopy of Condensed Media." Laser Optoacoustic Spectroscopy.
Berlin: Springer-Verlag, 1986. 45-58. Print.

Contributors and Attributions


Kevin Quan, Materials Science and Engineering Department, UC Davis undergraduate

5/9/2021 3 https://chem.libretexts.org/@go/page/40643
Mössbauer Spectroscopy
Mössbauer spectroscopy is a versatile technique used to study nuclear structure with the absorption and re-emission of gamma
rays, part of the electromagnetic spectrum. The technique uses a combination of the Mössbauer effect and Doppler shifts to
probe the hyperfine transitions between the excited and ground states of the nucleus. Mössbauer spectroscopy requires the use
of solids or crystals which have a probability to absorb the photon in a recoilless manner, many isotopes exhibit Mössbauer
characteristics but the most commonly studied isotope is 57Fe.

Introduction
Rudolf L. Mössbauer became a physics student at Technical University in Munich at the age of 20. After passing his
intermediate exams Mössbauer began working on his thesis and doctorate work in 1955, while working as an assistant lecturer
at Institute for Mathematics.In 1958 at the age of 28 Mössbauer graduated, and also showed experimental evidence for
recoilless resonant absorption in the nucleus, later to be called the Mössbauer Effect.In 1961 Mössbauer was awarded the
Nobel Prize in physics and, under the urging of Richard Feynman, accepted the position of Professor of Physics at the
California Institute of Technology.

Mössbauer Effect
The recoil energy associated with absorption or emission of a photon can be described by the conservation of momentum.In it
we find that the recoil energy depends inversely on the mass of the system. For a gas the mass of the single nucleus is small
compared to a solid.The solid or crystal absorbs the energy as phonons, quantized vibration states of the solid, but there is a
probability that no phonons are created and the whole lattice acts as the mass, resulting in a recoilles emission of the gamma
ray. The new radiation is at the proper energy to excite the next ground state nucleus. The probability of recoilles events
increases with decreasing transition energy.
PR = Pγ (1)

2 2
Pγ = Pγ (2)

2

2M ER = (3)
2
c

2

ER = (4)
2
2M c

Doppler Effect
The Doppler shift describes the change in frequency due to a moving source and a moving observer.f is the frequency
measured at the observer, v is the velocity of the wave so for our case this is the speed of light c , v is the velocity of the
r

observer, v is the velocity of the source which is positive when heading away from the observer, and f is the initial
s 0

frequency.
v + vr
f =( ) f0 (5)
v + vs

c
f =( ) f0 (6)
c + vs

In the case where the source is moving toward a stationary observer the perceived frequency is higher.For the opposite
situation where the source travels away from the observer frequencies recorded at the observer will be of lower compared to

7/7/2021 1 https://chem.libretexts.org/@go/page/1787
the initial wave. The energy of a photon is related to the product of Planck's constant and the frequency of the electromagnetic
radiation. Thus for increasing frequencies the corresponding energy also increase, and the same is true in the reverse case
where frequencies decrease and therefore energy decreases.
hc
E = = hv (7)
λ

The energy differences between hyperfine states are minimal (fractions of an eV) and the energy variation is achieved by the
moving the source toward and away from the sample in an oscillating manner, commonly at a velocity of a few mm/s.The
transmittance is then plotted against the velocity of the source and a peak is seen at the energy corresponding to the resonance
energy.

In the above spectrum the emission and absorption are both estimated by the Lorentzian distribution.

Mössbauer Isotopes
By far the most common isotopes studied using Mössbauer spectroscopy is 57Fe, but many other isotopes have also displayed
a Mössbauer spectrum. Two criteria for functionality are
1. The excited state is of very low energy, resulting in a small change
in energy between ground and excited state. This is because
gamma rays at higher energy are not absorbed in a recoil free
manner, meaning resonance only occurs for gamma rays of low
energy.
2. The resolution of Mössbauer spectroscopy depends upon the
lifetime of the excited state. The longer the excited state lasts the
better the image.
Both conditions are met by 57Fe and it is thus used extensively in
Mössbauer spectroscopy. In the figure to the right the red colored
boxes of the periodic table of elements indicate all elements that have isotopes visible using the Mössbauer technique.

Hyperfine Interactions
Mössbauer spectroscopy allows the researcher to probe structural elements of the nucleus in several ways, termed isomer shift,
quadrupole interactions, and magnetic splitting. These are each explained by the following sections as individual graphs, but in
practice Mössbauer spectrum are likely to contain a combination of all effects.

Isomer Shift
An isomeric shift occurs when non identical atoms play the role of source and absorber, thus the radius of the source, Rs, is
different that of the absorber, Ra, and the same holds that the electron density of each species is different. The Coulombic
interactions affects the ground and excited state differently leading to a energy difference that is not the same for the two
species.This is best illustrated with the equation:

RA ≠ RS (8)

ρS ≠ ρS (9)

7/7/2021 2 https://chem.libretexts.org/@go/page/1787
EA ≠ ES (10)

2 2 2 2
δ = EA − ES = nZ e (ρA − ρS )(Res − Rgs ) (11)
3

Where delta represents the change in energy necessary to excite the absorber, which is seen as a shift from the Doppler speed 0
to V1. The isomer shift depends directly on the s-electrons and can be influenced by the shielding p,d,f electrons.From the
measured delta shift there is information about the valance state of the absorbing atom

The energy level diagram for δ shift shows the change in source velocity due to different sources used. The shift may be either
positive or negative.

Quadrupole Interaction
The Hamiltonian for quadrupole interaction using 57
Fe nuclear excited state is given by
eQVZZ
2 2 2
HQ = [3 I − I (I + 1) + η(I − Iy )] (12)
Z X
12

where the nuclear excited states are split into two degenerate doublets in the absence of magnetic interactions. For the
asymmetry parameter η = 0 doublets are labeled with magnetic quantum numbers m = 3/2 and m = 1/2 , where the
es
+
− es
+

3/2 doublet has the higher energy. The energy difference between the doublets is thus
+
m es= −

−−−−−−
2
eQVzz η
ΔEQ = √1 + (13)
2 3

The energy diagram and corresponding spectrum can be seen as

Magnetic Splitting
Magnetic splitting of seen in Mössbauer spectroscopy can be seen because the nuclear spin moment undergoes dipolar
interactions with the magnetic field

E(mI ) = −gn βn Bef f mI (14)

7/7/2021 3 https://chem.libretexts.org/@go/page/1787
where g is the nuclear g-factor and β is the nuclear magneton. In the absence of quadrupole interactions the Hamiltonian
n n

splits into equally spaced energy levels of

The allowed gamma stimulated transitions of nuclear excitation follows the magnetic dipole transition selection rule:
+
ΔI = 1andΔmI = 0, 1 (15)

mI (1)

is the magnetic quantum number and the direction of β defines the nuclear quantization axis. If we assume g and A are
isotropic (direction independent) where g = g = g and B is actually a combination of the applied and internal magnetic
x y z

fields:
H = gβS ⋅ B + AS ⋅ I − gn βn B ⋅ I (16)

The electronic Zeeman term is far larger then the nuclear Zeeman term, meaning the electronic term dominates the equation so
S is approximated by ⟨S⟩ and

1
+
⟨Sz ⟩ = ms = −
(17a)
2

and
⟨Sx ⟩ = ⟨Sy ⟩ ≈ 0 (17b)

Hn = A⟨S⟩ ⋅ I − gn βn B ⋅ I (18)

Pulling out a −gn βn followed by I leaves


A⟨S⟩
Hn = −gn βn (− + B) I (19)
gn βn

Substituting the internal magnetic field with


A⟨S⟩
Bint = − (20)
gn βn

results in a combined magnetic field term involving both the applied magnetic field and the internal magnetic field
Hn = −gn βn (Bint + B) ⋅ I (21)

which is simplified by using the effective magnetic field B ef f

Hn = −gn βn Bef f ⋅ I (22)

References
1. Gütlich, P., Link, R., & Trautwein, A. (1978). Mössbauer spectroscopy and transition metal chemistry. Inorganic chemistry
concepts, v. 3. Berlin: Springer-Verlag.
2. G J Long and F Grandjean (1993) Mössbauer Spectroscopy Applied to Magnetism and Materials. New York., Science Vol.
1,, eds.
3. Lawrence Que, J., Ed. (2000). Physical Methods in Bioinorganic Chemistry. Sausalito, CA, University Science Books.
4. Filatov, M. (2007). "On the calculation of Mössbauer isomer shift." The Journal of Chemical Physics 127(8): 8.

7/7/2021 4 https://chem.libretexts.org/@go/page/1787
5. Introduction to mossbauer spectroscopy. 2013]. Available from
http://www.rsc.org/Membership/Networking/InterestGroups/MossbauerSpect/part2.asp.
6. Bubek, Moritz, and Dennis Rettinger. 2004. Mossbauer-effekt.
7. Haskins, J. R. 1965. Advanced mossbauer-effect experiments. American Journal of Physics: 646.
8. Kistner, O. C., and A. W. Sunyar. 1960. Evidence for quadrupole interaction of Fe57m, and influence of chemical binding
on nuclear gamma-ray energy. Physical Review Letters 4 (8): 412.
9. Preston, R. S., S. S. Hanna, and J. Heberle. 1962. Mössbauer effect in metallic iron. Physical Review 128 (5): 2207.

Problems
1. The magnetic splitting of m = 0 intensity of transition is related to si n (θ) of the angle between the incoming gamma
I
2

ray and the effective magnetic field. When is the intensity of transition at max?
2. Why is it important for the sample to be in solid or crystalline state?
3. What case will result in a delta shift of 0.00 mm/s?
4. Why is the Doppler effect important to Mössbauer spectroscopy?
5. Why are both the emission and absorption distributions the same? (both estimated with Lorentzian functions)

Contributors and Attributions


Kevin MacDow (UCD)
Thumbnail: Magnetic splitting of the nuclear energy levels. (Public Domain; Hokur via Wikipedia)

7/7/2021 5 https://chem.libretexts.org/@go/page/1787

You might also like