DV-X Calculation of X-Ray Emission Spectra

Download as pdf or txt
Download as pdf or txt
You are on page 1of 25

X-Ray Emission Spectrum

Related terms:

Photon, Ligand, Chemical Element, Fluorescence, Transition Element, Tight Bind-


ing Model, X-Ray Emission

View all Topics

Learn more about X-Ray Emission Spectrum

DV-X Calculation of X-Ray Emission


Spectra
Jun Kawai, in Advances in Quantum Chemistry, 1998

1 Introduction
X-ray emission spectra of solids and molecules are methods of measuring electronic
structure of matter [1–5]. The x-ray emission spectra reflect the occupied electronic
structure as shown in Fig. 1 , while the x-ray absorption spectra reflect the unoccu-
pied molecular orbitals (MO). These x-ray spectra represent local (L) and partial (P)
electron density of states (DOS) because of the electric dipole selection rule, and thus
the x-ray spectroscopy is a powerful tool to study the electronic structure of matter.
The development of synchrotron radiation facilities makes it possible to measure
the x-ray emission spectra more easily with using a resonance condition [6,7], and
we can get information on the detailed electronic structure of matter using the
characteristics of synchrotron radiation such as resonance and polarization.
Fig. 1. Measured sulfur K (K-M2,3) x-ray emission spectrum of Na2SO4 and the
assignment of x-ray transitions for the SO42 − cluster. The x-ray intensity reflects the
S 3p atomic orbital component in the molecular orbitals.

(Taken and modified from Kawai et al. [1].)

The calculation of x-ray emission spectra of molecules or solids are one of the most
successful applications of the discrete variational (DV) Hartree-Fock-Slater (X ) MO
method using cluster approximation [8–10], which was originally coded by Ellis
and his coworkers [11–14] based on Slater’s X exchange potential [15]. The DV-X
method has several advantages for the calculation of x-ray transition process as
follows.

Firstly, it can calculate the inner-shell hole state. The atomic orbitals as well as
the molecular orbitals are relaxed by the inner-shell or valence shell hole in the
self-consistent field (SCF) calculation. This is because the DV-X method uses SCF
numerical atomic orbitals as the basis set in an atomic potential in a cluster and thus
the orbitals contract or expand according to the hole potential. They also self-consis-
tently contract or expand due to the formation of chemical bonds. Figure 2 shows
the Fe 3d atomic orbital used as one of the basis functions in the DV-X method for
the ground state or the 1 s− 1 hole state. It is found that the 3d wavefunction contracts
due to the core hole potential. The Gaussian type orbital (GTO) basis sets are usually
fixed and thus the relaxation effect due to the core hole is included either by the
change of MO coefficients or by the configuration interaction. Thus the expression
of core hole state by the fixed basis function requires a large basis set. On the other
hand, the numerical atomic basis functions are self-consistent themselves, and thus
are good eigenfunctions of the Hamiltonian. Therefore the DV-X method usually
needs only the minimal basis set for the MO calculations.

Fig. 2. Iron 3d atomic orbital used in the DV-X method as a basis function. Solid
line: ground state, and dotted line: 1 s− 1 hole state.

Figure 3 shows the electron density of states of [MnO6]10 − for the ground state and
the 1 s− 1 core hole state [16]. This cluster is a model cluster of MnO. In the ground
state, the Mn 3d and O 2p are separated in energy and the hybridization is week.
Thus MnO is an ionic solid. In the 1 s− 1 core hole state, however, the Mn 3d orbital
energy becomes as deep as that of the O 2p and they hybridize strongly each other.
Thus we can know, using the DV-X method, that the initial state of x-ray emission
(or final state of photoionization) of MnO is no longer ionic but covalent.

Fig. 3. Electron densities of states of [MnO6]10 − cluster for the ground state (left)
and the 1 s ↑− 1 hole state (right) calculated by the spin-included DV-X method. The
shadowed area indicates oxygen 2p component.

(Taken from Suzuki [16].)


The second advantage of the DV-X method is that the basis functions of the DV-X
method are atomic orbitals. Thus the number of nodes is exact as shown in Fig. 4
, where Si 2 s GTO in GAUSSIAN method is compared with the numerical basis
function used in the DV-X method [17]. The use of the atomic orbital wavefunction
makes it possible to perform the direct calculation of the electric dipole matrix
elements, e.g. < 1s| r| 2p>, using the DV-X MO, yielding better result than when
using a GTO basis MO.

Fig. 4. Comparison of atomic basis functions of Si 2 s used in the DV-X (solid line)
and GAUSSIAN 6-31G* (broken line).

(Taken and modified from Nakamatsu [17].)

The third advantage of the DV-X method is in the cluster approximation. The core
hole state as the initial state of x-ray emission or final state of x-ray absorption is
treated as an impurity atom in solid when using the band theory. However, since
the core hole potential usually affects the next nearest neighbor atoms at most, it is
enough for the calculation of x-ray emission spectra to use the clusters which include
the second nearest neighbor atoms to the x-ray emitting atom as is described below.

The fourth advantage is that the precision of the DV calculation is comparable to


the experimental precision. The energy resolution of x-ray optics is at most 0.01 eV,
which is the comparable value to the DV-X precision. Thus the infrared spectra
cannot be reproduced by the DV-X method.

We have calculated for these several years the x-ray emission spectra of solids and
molecules and we have found that the calculation of x-ray emission spectra is one of
the most successful applications of the DV-X method. In the present paper, results
of our ongoing research, as well as published results are described.

> Read full chapter


X-ray Emission Spectroscopy, Methods
George N. Dolenko, ... Jolanta N. Latośińska, in Encyclopedia of Spectroscopy and
Spectrometry, 1999

Characteristic X-ray emission spectra


Characteristic X-ray emission spectra consist of spectral series (K, L, M, N…), whose
lines have a common initial state with the vacancy in the inner level. Labels of basic
X-ray transitions are shown in Figure 2. All electron levels with the principal quantum
number n equal to 1, 2, 3, 4, etc. are named as K, L, M, N etc. levels and denoted
with corresponding Greek letters and digit indexes. The electron transitions which
satisfy the dipole selection rules

Figure 2. Scheme of the most important X-ray emission transitions; n, I and j are
correspondingly the principal, orbital and total quantum numbers of K, L1, L2, L3
levels, etc.

[1]

are most intense. The dependence of X-ray emission line energy on atomic number
Z is defined by Moseley's law:

[2]

where Z is the atomic number and is the shielding constant, which varies from
series to series. Therefore, any X-ray emission spectral line is the finger-print of an
element.

With X-ray emission excitation by electron bombardment (primary emission) all


emission lines of the ith series appear when the X-ray tube voltage U exceeds the
ionization potential of ith level (Vi). At higher U the intensity of all lines of the ith
series, Ii, increases because the electrons penetrate deeper into the target substance
and, therefore, the number of excited atoms in the target increases. In the Vi < U <
3Vi region, the intensity obeys the rule I i (U − Vi) 2. With a further increase in U
X-ray emission begins to be absorbed by the target atoms, therefore the increase in
Ii is reduced. At U ≥ 11 Vi, Ii decreases because now most of the excited atoms are so
deep in the target that their emitted radiation is absorbed by the target substance.

X-ray emission spectra are usually excited by X-ray photons because most chemical
compounds are decomposed by electron bombardment. With X-ray emission spec-
tra excitation by photons [secondary emission or fluorescence (XFS)] the fluorescent
line intensity depends on the exciting photon energy h I i = 0 if h < åV i. All lines
of the ith series appear ifh = åVi; however, Ii decreases little with further increase
in h . Therefore, to excite X-ray fluoresence one must use a target that contains
a substance with intense characteristic X-ray lines whose energy just exceeds eV i.
Using the continous radiation of an X-ray tube with a target consisting mostly of
heavy elements it is possible to excite X-ray fluorescence.

The intensity of a characteristic X-ray spectrum (both primary and fluorescent)


depends on the probability pr of a radiation transition in the atom having the vacancy
in the ith level. The value of pr is determined by the total probability of photon
emission when this vacancy is filled by outer electrons. However, with a probability
pA the vacancy may be filled by outer electrons without radiation as the result of the
Auger-effect (see Figure 1). For the K series of medium and heavy elements pr > pA,
for the light elements pr < pA. For all others series of any elements pr << pA. The ratio
f = pr /(pr + pA) is called the yield of characteristic radiation.

However, X-ray characteristic lines appear because of single atom ionization; in X-ray
emission spectra weaker lines are found to occur as a result of binary (or multiple)
atom ionization when two (or more) vacancies are formed simultaneously in different
electron shells. If, for example, one vacancy is formed in the K shell of atoms and
filled by electrons belonging to the L2,3 shell, atoms emit an Ê 1,2 doublet. If another
vacancy is formed simultaneously which too is filled by electrons from the L2,3 shell,
then the final state will have a binary ionization L2,3L2,3, and would correspond to the
emission of radiation with energy exceeding that of the Ê 1,2 doublet. As a result,
in an X-ray emission spectrum a short wavelength Ê 3,4 doublet, called a satellite
of the main Ê 1,2 doublet, would appear. Because of such processes of multiple
ionization X-ray emission spectra may have a large number of satellites of the main
lines. Usually, the satellite intensity is some orders of magnitude less than that of
the main lines. However, if target atoms are bombarded by heavy ions with great
energy, the probability of multiple atom ionization becomes higher than that of
single ionization. Therefore, in this case the intensity of the main emission lines is
essentially less than that of the satellites.

> Read full chapter


X-ray Absorption Spectroscopy in Biol-
ogy (BioXAS)
Martin C. Feiters, Wolfram Meyer-Klaucke, in Practical Approaches to Biological
Inorganic Chemistry, 2013

In the X-ray emission spectrum of a first row transition metal (such as the Mn2+
represented in Figure 6C, left), the K 1 and K 2 lines are well resolved and more
intense than the K 1 and K 3 lines, which are not resolved, by an order of magnitude;
these are in turn more intense than the K satellite lines K 2,5 and K . Not indicated
in Figure 6C, but usually present for transition metals which have a total electron
spin S ≠ 0 (such as Mn2+), is the K line at slightly lower energy than the K 1,3 line.
This results from emission from the metal 3p level combined with a spin flip of a
3d electron and is therefore sensitive to the spin state of the metal ion. Of the K
satellite lines, the cross-over emission line K is extremely sensitive to the nature
of the coordinating ligands, because it involves emission from the ligand’s 2s level
to the metal’s 1s core hole, and allows one to distinguish O from N and C ligands.
This ligand identification is of interest as it gives information complimentary to that
obtained from EXAFS, which can typically not discriminate between coordination by
ligands from the same row of the periodic table, see text. Examples are the variation
in the number of O ligands to Mn in the so-called Kok cycle in Photosystem II (Yano
and Yachandra, 2008), and the identification of central atom bound to Fe in the Fe,
Mo cofactor of nitrogenase as C (Lancaster et al., 2011). A typical theoretical approach
calculation of the multiplet ‘ligand field multiplets’ to interpret K and K main lines,
and molecular orbital theory for the K satellites (Glatzel and Bergmann, 2005); both
are outside the scope of this chapter.

FIGURE 6C. Overview of ligand edge, L edge, and X-ray emission transitions; the
y-axis only gives relative energies and is not to scale. Left panel, effect of ‘low-Z’
ligands (C, N, O, F) on the X-ray emission of a Mn2+ complex. K edge excitation (blue)
leads to a 1s core hole intermediate state (central green oval), which can emit X-ray
fluorescence at various wavelengths. The final state obtained with K 1 fluorescence
(red box) is identical to that obtained by direct L3 edge excitation (red). Right panel,
illustration of probing mixed 3p orbitals of Cl or S ligands with a transition metal’s
3d orbital, in this case the singly occupied 3dx2-y2 orbital of Cu2+ (other Cu 3d orbitals
grouped together for clarity), by either K edge or L3 edge XAS.

> Read full chapter

Measurements of Radioactivity
Vlado Valković, in Radioactivity in the Environment, 2000

5.4.1 Methods based on X-ray spectrometry


The electronic transitions which give rise to X-ray emission spectra involve core
electrons and are therefore relatively insensitive to the chemical and physical form
of the determinant (Bertin, 1978). As a result, analyses can be performed with a
minimum of sample preparation directly on materials in the condensed phase.
This insensitivity of sample matrix applies to the wavelength of the emitted X-rays,
not to their intensities and as quantitation is based on intensity measurement,
closely matched standards are required. X-ray emission spectra can be excited by
primary X-rays in a fluorescence experiment or by changed particles via collisional
excitation. The cross sections for excitation of X-ray emission are rather low and
this is combined with the low efficiency of collection, collimation, diffraction and
detection of the emitted X-rays. This low overall efficiency leads to a relatively
low sensitivity in some cases and is compounded by high backgrounds either from
scattered primary radiation in a fluorescence experiment or due to bremsstrahlung
in the charged-particle-excitation methods. Methods based on X-ray spectrometry
do not provide isotopic information about the sample. Nonetheless, there are a
number of radio analytical problems which can be solved by methods based on X-ray
spectrometry.

The following instrumental methods of elemental analysis are based on X-ray spec-
trometry.

X-ray Fluorescence Analysis XRF


Total Reflectance X-ray Fluorescence Analysis TXRF
Electron Microprobe Analysis EMA
Particle Induced X-ray Emission PIXE
Synchrotron Radiation Induced X-ray Emission SRIXE

XRF is the simplest of these methods. It allows bulk analysis of solid or liquid samples
with detection limits of approximately 0.1 μg. The method can thus only compete
with radiometric methods for the longest lived of radionuclides. It has approximately
the same sensitivity for
232
Th as alpha spectrometry but has the advantage that little sample preparation is
required and that analysis is rapid and easily automated. XRF would be the method
of choice for measurement of airborne thorium collected onto filter papers, for
example.

The more sophisticated methods address the problem of the low overall efficiency
of generation and acquisition of the X-ray spectrum. The low fluorescence cross
section is addressed by using a highly intense X-ray source, a synchrotron in the
SRIXE method. The high intensity of synchrotron X-rays allows the beam to be
focused and collimated whilst retaining significant intensity. The method can there-
fore be used in a microprobe mode and by moving the sample in a raster pattern
across the incident X-ray beam, elemental images can be generated with micron
spatial resolution. The scattered primary radiation background can be reduced by
using the total-reflectance technique in TXRF (Knoth and Schwenke, 1978). The
instrumental geometry limits scattering of primary X-rays in the direction of the
detector, however this is at the expense of increased sample preparation. The gains
in sensitivity achieved by each of these methods may be compounded in a method
which uses a total reflectance sample geometry in combination with a synchrotron
X-ray source.

The charged-particle-beam methods EMA and PIXE also allow elemental imaging
within the sample. These methods generally require that the sample be enclosed in a
vacuum. The approximately 15 keV electrons used in an EMA instrument, penetrate
only 1–2 μm into the sample. This rapid declaration of the charged particles gener-
ates bremsstrahlung X-rays which generate a strong background signal in the spec-
tral region of interest. EMA thus has relatively poor detection limits. The method can
be used for analysis of electrodeposits such as sources prepared for alpha-particle
spectrometry where the element of interest is present at high concentration in a very
thin surface layer. The approximately 2.5-MeV-proton beam used in PIXE analysis
penetrated much deeper into the sample than the EMA electron beam. The resulting
proton bremsstrahlung is less intense and backgrounds are therefore reduced. PIXE
can thus achieve much lower detection limits.

PIXE (Johansson and Campbell, 1988) and SRIXE (Jones and Gordon, 1989) have
similar imaging capabilities and detection limits but both suffer from the drawback
that they rely on major pieces of hardware, an accelerator in the PIXE experiment
and a synchrotron X-ray source for SRIXE.

> Read full chapter


Measurements of radioactivity
Vlado Valković, in Radioactivity in the Environment (Second Edition), 2019

4.4.1 Methods based on X-ray spectrometry


The electronic transitions which give rise to X-ray emission spectra involve core
electrons and are therefore relatively insensitive to the chemical and physical form
of the determinant (Bertin, 1978). As a result, analyses can be performed with a
minimum of sample preparation directly on materials in the condensed phase.
This insensitivity of sample matrix applies to the wavelength of the emitted X-rays,
not to their intensities and as quantitation is based on intensity measurement,
closely matched standards are required. X-ray emission spectra can be excited by
primary X-rays in a fluorescence experiment or by changed particles via collisional
excitation. The cross sections for excitation of X-ray emission are rather low and
this is combined with the low efficiency of collection, collimation, diffraction and
detection of the emitted X-rays. This low overall efficiency leads to a relatively
low sensitivity in some cases and is compounded by high backgrounds either from
scattered primary radiation in a fluorescence experiment or due to bremsstrahlung
in the charged-particle-excitation methods. Methods based on X-ray spectrometry
do not provide isotopic information about the sample. Nonetheless, there are a
number of radio analytical problems which can be solved by methods based on X-ray
spectrometry.

The following instrumental methods of elemental analysis are based on X-ray spec-
trometry:

X-ray fluorescence analysis (XRF)


Total reflectance X-ray fluorescence analysis (TXRF)
Electron microprobe analysis (EMA)
Particle-induced X-ray emission (PIXE)
Synchrotron radiation-induced X-ray emission (SRIXE)

XRF is the simplest of these methods. It allows bulk analysis of solid or liquid samples
with detection limits of approximately 0.1 μg. The method can thus only compete
with radiometric methods for the longest lived of radionuclides. It has approximately
the same sensitivity for 232Th as -spectrometry but has the advantage that little
sample preparation is required and that analysis is rapid and easily automated. XRF
would be the method of choice for measurement of airborne thorium collected onto
filter papers, for example.

The more sophisticated methods address the problem of the low overall efficiency
of generation and acquisition of the X-ray spectrum. The low-fluorescence cross
section is addressed by using a highly intense X-ray source, a synchrotron in the
SRIXE method. The high intensity of synchrotron X-rays allows the beam to be
focused and collimated while retaining significant intensity. The method can there-
fore be used in a microprobe mode and by moving the sample in a raster pattern
across the incident X-ray beam, elemental images can be generated with micron
spatial resolution. The scattered primary radiation background can be reduced by
using the total reflectance technique in TXRF (Knoth and Schwenke, 1978). The
instrumental geometry limits scattering of primary X-rays in the direction of the
detector, however this is at the expense of increased sample preparation. The gains
in sensitivity achieved by each of these methods may be compounded in a method
which uses a total reflectance sample geometry in combination with a synchrotron
X-ray source.

The charged-particle-beam methods EMA and PIXE also allow elemental imaging
within the sample. These methods generally require that the sample be enclosed in
a vacuum. The approximately 15 keV electrons used in an EMA instrument, pene-
trate only 1–2 μm into the sample. This rapid declaration of the charged particles
generates bremsstrahlung X-rays which generate a strong background signal in the
spectral region of interest. EMA thus has relatively poor detection limits. The method
can be used for analysis of electrodeposits such as sources prepared for -particle
spectrometry where the element of interest is present at high concentration in a very
thin surface layer. The approximately 2.5-MeV-proton beam used in PIXE analysis
penetrated much deeper into the sample than the EMA electron beam. The resulting
proton bremsstrahlung is less intense and backgrounds are therefore reduced. PIXE
can thus achieve much lower detection limits.

PIXE (Johansson and Campbell, 1988) and SRIXE (Jones and Gordon, 1989) have
similar imaging capabilities and detection limits but both suffer from the drawback
that they rely on major pieces of hardware, an accelerator in the PIXE experiment
and a synchrotron X-ray source for SRIXE.

> Read full chapter

Chemical Imaging Analysis


Freddy Adams, Carlo Barbante, in Comprehensive Analytical Chemistry, 2015

8.3.1 Particle-Induced X-ray Emission


The primary analytical technique in IBA is PIXE, which usually measures the X-ray
emission spectrum with an energy-dispersive semiconductor detector. As such, it is
quite similar – as an analytical technique and as an imaging tool – to other X-ray
analytical techniques that use X-ray or electron excitation sources. PIXE differs from
Scanning Electron Microscopy with Energy-Dispersive X-ray (SEM-EDX) (Chapter 7)
in the particle interaction mechanism. The impinging radiation extends deeper in
the sample, while X-Ray Fluorescence (XRF) (Chapter 6) penetrates even deeper, and
decays exponentially; X-ray excitation practically becomes a bulk analytical technique
in numerous situations. The three techniques share numerous common points
in methodology (e.g. the detection tools used for X-ray spectrometry) but differ
significantly in the projectile–sample interaction process.

As the bombarding particles are slowed down in subsequent layers of the sample,
they excite characteristic X-rays with a particular energy-dependent cross section
until the threshold energy for production is reached. The cross sections for the
excitation of K and L shell electrons decrease strongly with energy and are also
dependent on the atomic number. The Ion Beam Analysis Nuclear Data Library
(IBANDL) contains a number of experimental nuclear cross sections [4].

Quantitative Analysis
Calibration in PIXE analysis needs to take into account the energy loss of the
projectile in the sample and the absorption and fluorescence effects of the produced
X-rays. For a thin, uniform and homogeneous sample with negligible energy loss of
the bombarding particle and negligible absorption of the X-rays in the sample, the
yield of each X-ray peak can be calculated from the physical constants such as the
ionisation cross section at the particular particle’s energy and the factors that govern
the X-ray production. The latter are equivalent to those that apply to SEM-EDX
and micro-XRF in Chapters 6 and 7. The total number of incident particles can be
expressed and measured as particle current.

For thick homogeneous targets in which the bombarding particles are completely
stopped and the emitted X-rays are subjected to considerable absorption in the sam-
ple, the yield of each X-ray peak can be calculated according to iterative procedures
similar to those applied to electron and X-ray excitation [5].

The simplest techniques used for calibrating a PIXE system are based on the deter-
mination of sensitivity factors, which assign absolute concentration data to numbers
of counts in X-ray peaks. These are similar to those performed in SEM-EDX or
micro-XRF. Calibration procedures take into account the slowing down of bombard-
ing particles, as well as absorption and fluorescence effects, according to methods
similar to the ZAF correction procedures discussed in Section 7.4.6. The SigmaCalc
Web site of the IAEA Nuclear Data Section contains evaluations of scattering and
reaction cross-section data. It is part of the IBANDL, a Web site of the IAEA Nuclear
Data Section containing extensive scattering and reaction cross-section data relevant
to IBA [4].
Overall, PIXE quantification is rather troublesome, with quite a long traceability
chain. However, in favourable circumstances and for homogeneous samples, a sys-
tematic comparison between the theoretical and experimental calibration methods
proves that the determination of elemental concentrations can be performed with
an overall accuracy of 3–5% [5,6].

Ion Beam Focussing


The focussing of ion beams of megaelectron volt energy is carried out by means of
magnetic and electrostatic quadrupole lenses in different configurations. Magnetic
quadrupoles are used to focus megaelectron volt energy particle beams and typ-
ically produce beam spots of 1–10 μm with currents of 100–1000 pA. Submicron
beamlines usually have a two-stage focussing system consisting of a first stage
focussing on a small spot, which is then used as object in a second stage and
demagnified to submicrometre dimensions. Micro and submicron beamlines are
equipped with post-lens magnetic scanning systems that raster the beam over a
large sample area to obtain 2-D information on the elements in the sample. PIXE can
be applied with an external beam, enabling the analysis of materials in the air or other
environments. In PIXE-tomography, the characteristic X-ray emission is detected as
the sample is rotated over a number of angles. PIXE-tomography is most commonly
performed for minor and major elements, but the quantification can be performed
only with an accuracy close to 10% if the sample contains low-Z elements, which
cannot be detected with PIXE or are available through other measurements. PIXE
has no depth-profiling capability and cannot detect low atomic number elements
for the same reason as micro-XRF (absorption, Auger yields, decreased detection
efficiency).

PIXE and other particle-induced analytical techniques are well covered in the hand-
book published by Johansson et al. [7]. Several more recent surveys are available
[5,8–10].

There are numerous applications of PIXE in various fields. Figure 8.3 shows a typical
example of PIXE mapping in the geological field for the measurement of PIXE
element maps of fluid bubbles inside silicate glass inclusions in quartz phenocrysts.
Figure 8.4 illustrates the results of a combination of PIXE with 3-D depth profiling
with RBS, see next section, that was used to establish the distribution of trace
elements in individual corneal and retinal areas in frozen sections from rats. The
distribution of endogenous trace elements in the cornea and retina was found to
be nonhomogeneous. The most abundant metal in the cornea is calcium followed
by zinc. Iron and copper are present in small amounts localised particularly to the
epithelium. Iron is also identified in corneal fibroblasts (keratocytes).
Figure 8.3. Optical microscope (a and b) and PIXE images of fluid bubbles, rich in
Cu and Cl, inside silicate glass inclusions in quartz phenocrysts from the Okataina
volcanic region of New Zealand. The magnified bubble is shown as an inset in the
left image. (see colour plate).

From P. Davisson, V.S. Kamenetsky, Chem. Geol. 237 (2007) 372 [11].

Figure 8.4. PIXE false colour images of the distribution of endogenous trace ele-
ments in the rat cornea. (a) Diagram showing different corneal structures. (b) Stained
corneal section showing cell nuclei. Pictures of phosphorus, sulphur, potassium,
chlorine, calcium, zinc, iron, copper. White arrow: Keratocyte, (corneal fibroblast).
Scale bar, 50 mm. (see colour plate).

From M. Ugarte, et al., Metallomics 6 (2014) 274 [12].

> Read full chapter

X-ray absorption and emission spec-


troscopy in biology
Martin C. Feiters, Wolfram Meyer-Klaucke, in Practical Approaches to Biological
Inorganic Chemistry (Second Edition), 2020

As an example of a photon-in–photon experiment with incident energy in and emit-


ted energy out, the analysis of the 2p to 1s emission stimulated by 1s-3d absorption
of a 3d transition metal is given in Fig. 7.7, left. The accurate measurement of
out in X-ray emission spectra requires element-specific analyzer crystals that have

a high-energy resolution, close to the natural linewidth. In the Johann approach,


this is achieved by point-to-point scanning using a crystal bent with radius 2R,
which is placed with the sample and detector on a Rowland circle with radius R
to ensure that radiation from the sample that is selected by the crystal hits the
detector (Fig. 7.7, right). In a wavelength dispersive arrangement, X-rays emitted by
the sample are diffracted and the photons of different energies detected separately
with a position-sensitive detector.

Figure 7.7. (Left) Principle of a hard X-ray photon-in/photon-out experiment. (Right)


Experimental setup at a storage ring beamline, with the sample, analyzer crystals,
and detector in Rowland circle geometry. The arrows indicate the motion of the
components when a spectrum of the emitted X-rays is taken.

Source: Adapted from Glatzel, P., Sikora, M., Fernández-García, M., 2009. Eur. Phys.
J. Special Top. 169, 207–214.

> Read full chapter

Scanning Electron Microscopy (SEM)


and Energy-Dispersive X-Ray (EDX)
Spectroscopy
M. Abd Mutalib, ... J. Jaafar, in Membrane Characterization, 2017
2.3 Principles of Energy Dispersive X-Ray Spectroscopy
EDX spectroscopy is involved in the detection of elemental composition of substance
by using scanning electron microscope. EDX is able to detect elements that possess
the atomic number of higher than boron and these elements can be detected at
concentration of least 0.1%. The application of EDX includes material evaluation
and identification, contamination identification, spot detection analysis of regions
up to 10 cm in diameter, quality control screening, and others.

Upon collision with the electron beam in typical SEM, the samples interact with
the beam and produce characteristic X-rays. Due to the principle that none of the
elements have the same X-ray emission spectrum, they can be differentiated and
measured for its concentration in the sample [6]. The X-ray is the result of the primary
beam of electron interaction with the nucleus of the sample atom. Primary electron
beam will excite the electron in the nucleus of an atom, ejecting it from the nucleus
and creating an electron hole. An electron from the outer shell (higher energy) of the
atom will replace the missing ejected electron and releases the superfluous X-ray.
The emitted X-ray consists of X-ray continuum (generated by the deceleration of
electron) and characteristic X-ray (generated as a result of higher shell electron filling
the electron hole in the nucleus shell) as shown in Fig. 9.6.

Figure 9.6. Schematic representation of the types of X-ray spectrum emitted upon
bombardment of fast electron.

Reprinted with permission from Jerosch J, Reichelt R. Scanning electron


microscopy studies of morphologic changes in chemically stabilized ul-
trahigh molecular weight polyethylene. Biomed Tech Berl 1997;42:358–62.
http://dx.doi.org/10.1007/978-0-387-49762-4_3. Copyright 2016, Springer.

X-ray continuum is not paramount for the identification of elements in the sample
and need to be identified to differentiate them. The intensity of the X-ray continuum
is contributed by the factors such as probe current, accelerating voltage supplied,
and the atomic number of the sample. On the other hand, the characteristic X-ray
will be recorded by the energy dispersive spectrometer for the measurement of the
elemental composition in the specimen.

> Read full chapter

X-ray Emission Spectroscopy, Applica-


tions
George N. Dolenko, ... Jolanta N. Latośińska, in Encyclopedia of Spectroscopy and
Spectrometry, 1999

MO structure investigation
Any variations in the composition, structure, stereochemistry or coordination char-
acter of a molecule change its chemical properties and MO (molecular orbital)
structure. MO changes are clearly observed by the fine structure of XFS (X-ray
fluorescence spectroscopy). This makes it possible to relate some features of the
chemical behaviour of compounds to their electronic structure and opens a way
to various ‘chemical property–electronic structure parameter’ correlations which
are frequently of help for explaining and predicting the chemical properties of
compounds.

The transitions from valence atomic levels to vacancies in inner shells form X-ray
valence emission lines, reflecting the structure of the valence levels or zones. Elec-
tron transitions between different inner levels form inner X-ray emission lines. The
study of the fine structure of different valence emission lines of all the atoms in a
compound allows detailed investigation of the structure of valence levels or zones.
Research into the shifts of inner X-ray emission lines allows one to investigate
effective charges on the corresponding atoms.

For example, consider the X-ray emission spectra of sulfur. The initial state of a sulfur
atom for X-ray emission is that with a vacancy in the K or L2,3 level. This vacancy
is rapidly filled (within 10−16 − 10−14 s) as a result of transitions obeying the dipole
selection rules, i.e. 2p → 1s (K lines), 3p → 1s (K lines) or 3s → 2p, 3d → 2p (L2,3 lines)
transitions. The energy released in this case is emitted from the atom as either an
Auger electron or an X-ray quantum (Figure 1).
Figure 1. Scheme showing the change of one-electron energies of 1s, 2p levels and
the energy of the A K line in the ions A+ and A− with respect to the neutral atom A°.

Whereas SK are inner lines, SK and SL2,3 are valence lines. The energies of atomic
np → 1s transitions can be represented by the equations

[1]

[2]

where h is the energy of the emission quantum, Efin is the final energy of the system,
Einit is the initial energy of the system, E−nl is the energy of the system with an nl
electron removed and np is the one-electron energy of the np level.

Thus, in a one-electron approximation the distance between individual maxima in


a spectral series is equal to the difference in one-electron energies of the corre-
sponding atomic levels. In molecules the 3s, 3p and 3d electrons of the sulfur atom
are involved in chemical bonding to form an MO system. In this case, the SK
spectrum (S3p → S1s interatomic transitions), for example, corresponds to MOi →
S 1s transitions, and the distances between spectral maxima correspond to energy
differences of the appropriate occupied molecular levels:

[3]

The intensity of X-ray emission lines is determined by the relation (for the SK series
as an example):

[4]

where NSnp is the np level population of the sulfur atom, E1 is the energy of Snp → S1s
transitions, Snp is the wavefunction of sulfur np orbitals and S1s is the wavefunction
of the sulfur 1s orbital.
For molecules this expression is transformed to the equation

[5]

where i is the wavefunction of ith MO.

In the generally accepted MO LCAO model the wavefunction of any ith MO can be
represented as an AO sum:

[6]

where j is the wavefunction of jth AO and cij are the coefficients.

Then, with account taken of the dipole selection rules, Equation [5] is transformed
into

[7]

where ci is the coefficient of i at the wavefunction of S3p AO ( S3p). Thus, the ratio
of partial intensities in the SK spectrum of a molecule is equal to that of the squares
of the coefficients of the S3p wavefunction in the expansion of the corresponding
MOs in terms of AOs:

[8]

The above is also true for other X-ray emission series. Important features of X-ray
emission spectra are the comparative ease of interpretation and the possibility
of investigating the electronic structure from the ‘viewpoint’ of any atom of the
molecule investigated.

Example: electronic structure of the sulfate ion


Information concerning the electronic structure of molecules provided by XFS can
be well illustrated with the sulfate ion as an example. The wavefunction of any ith
valent MO of the sulfate ion can be described by the equation

[9]

All possible X-ray fluorescence spectra of the sulfate ion are presented in Figure 2
whereas Figure 3 shows an MO diagram constructed from a full set of these spectra.
‘Adjustment’ to the scale of the ionization potentials [IP] of valence electrons is ef-
fected by subtracting the X-ray transition energies from the IPs of the corresponding
inner levels (S1s, S2p, O1s) determined by the use of data of X-ray photoelectron
spectroscopy:
Figure 2. Full set of X-ray fluorescence spectra of the sulfate ion on an energy scale
corresponding to the ionization potentials of valence electrons.
Figure 3. Scheme of the sulfate ion MOs obtained from the full set of X-ray fluores-
cence spectra.

[10]

[11]

[12]

From these, the following equations are derived

[13]

[14]

[15]

All MOs with c5 ≠ 0 are displayed in the O K spectrum (O2p → O1s transitions),
those with c2 ≠ 0 in the SK spectrum (S3p → S1s transitions) and those with c1 ≠ 0
and c3 ≠ 0 in the SL2,3 spectrum (S3s → S2p and S3d → S2p transitions). From the
spectra shown in Figure 3 it follows that the highest occupied MO 1t1 (maximum
A in the OK spectrum) consists of only O2p electrons; next, the 3t2 and 1e levels
(maxima B, C and W) significantly correspond to the π bond S3d – O2p; the 2t2 level
then follows (maxima D and F), which corresponds to the strong S3p – O2p bond.
The 2a1 level (maxima E and V), consisting of the S3s AO and, possibly, the O2s with
a small admixture of O2p AO, lies even deeper. Deep 1t2 and 1a1 MOs consisting
mainly of the O2s AO are seen as low intensity long-wavelength maxima G and M
respectively.

Consequently, much information on the MO structure of a chemical species under


investigation can be derived from X-ray emission spectra.

> Read full chapter

Oxide-based Systems at the Crossroads


of Chemistry
John Meurig Thomas, in Studies in Surface Science and Catalysis, 2001

2 BACKGROUND
The only major ex situ technique that I propose to dwell upon is high-resolution
(transmission) electron microscopy (HRTEM) which, nowadays (see later) can also be
used for in situ studies. Its value as a tool in post mortem (as well as pre natal) studies
of many catalysts has been highlighted on numerous previous occasions — see, for
example, refs. [2–5]. I also wish to say a few words about electron crystallography[4].

So far as microporous oxide catalysts are concerned — and here I focus predomi-
nantly on zeolitic ones or their aluminophosphate analogues — the great merit of
HRTEM[4–7] is that it can be used to “read off ”, in real-space images, the symmetry
elements of the structure, and often yield a reasonably accurate value of the number
of tetrahedral sites in a pore aperture. (In general, 5-, 6-, 8-, 10- or 12-membered
(T sites) can be straightforwardly discerned from the HRTEM images recorded
down the appropriate zone axis). This real-space approach also proves invaluable
in identifying intergrowth structures. I have given elsewhere[8] an account of how
several of the ostensibly different members of the ZSM family of synthetic zeolites
turned out to be no more than intergrowth variants of well defined end-members.
HRTEM, twenty years ago, was shown[9] to be capable of “seeing” in real space the
intergrowth variants of ZSM-5 and ZSM-11.

With CCD detectors and other instrumental advances[4,6] the age of electron crys-
tallography is well and truly with us. Small, ultrathin specimens of zeolites (or
mesoporous silicas) can be processed by a combination of electron diffraction and
HRTEM imaging, so as to yield a full, hitherto unknown, structure, thanks to the
pioneering work of Terasaki and his colleagues in Tohuku University.

Surface structures of complex oxides are also amenable to direct imaging and deter-
mination by HRTEM (which has the advantage over scanning tunnelling microscopy
in being able, through the observation of parallel X-ray emission spectra, to yield
elemental composition)[10]. There are two examples which I wish to cite.

First, La2CuO4, a material that is of great relevance in the field of warm supercon-
ductors, has been studied in detail by my former colleague Wuzong Zhou, now of
the University of St. Andrews. A strictly stoichiometric powdered sample of La2CuO4
was viewed down the [110] direction by HRTEM. The outline of the [001] surface is
seen in Figure 1.
Figure 1. A typical region of the surface of La2CuO4 viewed down the [110] direction.
The outer layers of the solid have the structure and composition of La2O3 (after Zhou
et al[11])

Significantly, it is found that the surface composition does not correspond to


La2CuO4. The images leave no doubt (and microanalysis confirms it) that the exterior
surfaces (the last three layers or so) have the so-called C-La2O3 structure and com-
position[11]. Zhou and I have recently reported[10] that, in the warm superconductor
HgBa2CuO4 +  , the exterior and immediate sub-surface region is stoichiometrically
quite different from the bulk. There are well-known reasons, discussed by Freund[12]
recently, which lead us to expect polar surfaces (of oxides) to undergo reconstruction
on exposure to vacuum (or an ambient gas). What is surprising here, in this pow-
dered specimen of (nominally) La2CuO4, is that there is extensive reconstruction and
reconstitution at the surface regions.

Low-energy electron diffraction (LEED) studies conducted by Freund et al[13], show


that, in simple oxides belonging to the corundum family (e.g. Al2O3, Cr2O3 and
Fe2O3), there is quite extensive reconstruction (in a direction perpendicular to the
[0001] surface), amounting to between 50 and 60 percent change, compared with the
inter-planar distances inside the bulk of the oxide. With the complex oxides that have
zeolitic structure, however, the degree of reconstruction is rather minute. Thus, in
zeolites L (idealised formula Na3K6[Al9Si27O72]) Terasaki et al has shown[14] (Figure 2)
by HRTEM that the surface structure essentially retains that of the bulk, with the
individual cancrinite cages being intact right up to the last surface layer.
Figure 2. Unlike many simple metal oxides, such as those possessing the corundum
structure, there is very little reconstruction at the outer layers of zeolite L

In addition to being able to cope with the structure determination of beam-sensitive


oxides by the methods alluded to above, many other developments in HRTEM
now underway are likely to contribute to further physico-chemical elucidation of
oxide surfaces, simple and complex. Thus, at the ex situ level, following Bovin’s
recent classic work[15], one can already see the advantage in recording energy-filtered
transmission electron micrographs (EFTEM) for the elemental mapping of ‘frozen’
particles. This approach of Bovin’s promises to be able to follow sub-monolayer
structural changes at solid surfaces undergoing reactions with liquids. There is also
the (expensive) prospect of doing photo-emission electron microscopy (PEEM) using
synchrotron radiation as the primary beam[16].

Gai[5,17,18] has used HRTEM successfully as an in situ technique in her pioneering


work on the dynamic reduction of vanadyl pyrophosphate (also known as VPO) under
reaction conditions. She was able to chart the formation of extended defects by a
glide shear mechanism. Ordering of the glide shear (so-called GS defects) leads
to a new structure by transformation of the orthorhombic vanadyl pyrophosphate
[(VO)2P2O7] into an anion-deficient tetragonal structure. The GS defects essentially
preserve anion vacancies and do not lead to a collapse of the lattice (Figure 3),
and are distinct from the well-known crystallographic shear (CS) planes, which
eliminate oxygen ion vacancies. These GS defects, as Gai has shown elsewhere[5],
have important consequences in the VPO-catalysed conversion of butane to produce
maleic anhydride[19].
Figure 3. In situ studies using HRTEM (See text) reveal that so-called glide shear
defects, identified by Gai[19] play an important role in the catalytic action of (VO2)P-
2O7 during the oxidation of butane to maleic anhydride. P1 and P2 are partial screw
dislocations.

> Read full chapter

ScienceDirect is Elsevier’s leading information solution for researchers.


Copyright © 2018 Elsevier B.V. or its licensors or contributors. ScienceDirect ® is a registered trademark of Elsevier B.V. Terms and conditions apply.

You might also like