Vdoc - Pub Understanding Nanomaterials
Vdoc - Pub Understanding Nanomaterials
Vdoc - Pub Understanding Nanomaterials
nanomaterials
Malkiat S. Johal
understanding
nanomaterials
understanding
nanomaterials
Malkiat S. Johal
CRC Press
Taylor & Francis Group
6000 Broken Sound Parkway NW, Suite 300
Boca Raton, FL 33487-2742
© 2011 by Taylor & Francis Group, LLC
CRC Press is an imprint of Taylor & Francis Group, an Informa business
This book contains information obtained from authentic and highly regarded sources. Reasonable
efforts have been made to publish reliable data and information, but the author and publisher cannot
assume responsibility for the validity of all materials or the consequences of their use. The authors and
publishers have attempted to trace the copyright holders of all material reproduced in this publication
and apologize to copyright holders if permission to publish in this form has not been obtained. If any
copyright material has not been acknowledged please write and let us know so we may rectify in any
future reprint.
Except as permitted under U.S. Copyright Law, no part of this book may be reprinted, reproduced,
transmitted, or utilized in any form by any electronic, mechanical, or other means, now known or
hereafter invented, including photocopying, microfilming, and recording, or in any information stor-
age or retrieval system, without written permission from the publishers.
For permission to photocopy or use material electronically from this work, please access www.copy-
right.com (http://www.copyright.com/) or contact the Copyright Clearance Center, Inc. (CCC), 222
Rosewood Drive, Danvers, MA 01923, 978-750-8400. CCC is a not-for-profit organization that pro-
vides licenses and registration for a variety of users. For organizations that have been granted a pho-
tocopy license by the CCC, a separate system of payment has been arranged.
Trademark Notice: Product or corporate names may be trademarks or registered trademarks, and are
used only for identification and explanation without intent to infringe.
Visit the Taylor & Francis Web site at
http://www.taylorandfrancis.com
and the CRC Press Web site at
http://www.crcpress.com
This book is dedicated to my
beautiful daughter Simran
Contents
Preface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xv
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xvii
About the Author . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xix
vii
viii Contents
Appendix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 281
Glossary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 283
Preface
TO THE STUDENT
Nanoscience is a rapidly changing field where new innova-
tions and discoveries are being made every day. To write a
book that captures even a fraction of what the scientific liter-
ature has produced over the last few years would be a monu-
mental task. With this in mind, the topics presented in this
book are carefully selected to provide a basic understanding
of the field. Many important topics, such as computational
chemistry and solid state physics, have been given limited
coverage, largely because I want this book to be accessible
to any student who has taken introductory college-level sci-
ence courses.
This book is written for a full- or half-semester course in
nanoscience with an emphasis on understanding nanoma-
terials. The stress on “understanding” is the key behind the
objective of the text: to provide fundamental insight into the
molecular driving force underlying self-assembly processes,
as well as to explain how to characterize the resulting nano-
materials. Knowledge of self-assembly and characterization is
essential for an understanding of these interesting systems.
It should be noted that this book does not draw heavily from
scientific papers; rather, it should be used in conjunction with
the primary literature.
TO THE INSTRUCTOR
I have drawn relevant material from many scientific disci-
plines, assuming only a basic level of competency in physics,
chemistry, and biology. Mathematical rigor has been limited
to presenting key results and simple proofs. Instructors should
use their discretion in placing emphasis on or “filling holes”
in areas that may seem somewhat inadequate or limited in
scope. The half-course model is suggested for teaching mate-
rial directly from the book and solving the end-of-chapter
problems. For a full-semester course, the book should be used
in a course that requires students to refer to the primary lit-
erature. The latter may be more suitable for intermediate to
xv
xvi Preface
Writing this book would not have been possible without the
support of friends and family. My research students and mem-
bers of my Soft Nanomaterials class (2008 and 2009) have been
crucial in helping me to develop a textbook that meets the
needs of students interested in this field. I would like to express
sincere gratitude to Robert Rawle, Theodore Zwang, Michael
Haber, Michael Gormally, and Thomas Lane for making valu-
able contributions, including editing and extensive proofread-
ing of the manuscript. In addition, for their detailed review of
the manuscript and constructive comments, I’m very grateful
to Professors Lisa Klein at Rutgers University, Joseph Tracy at
North Carolina State University, and Marcus Lay at University
of Georgia. Last but not least I would like to thank my editor,
Luna Han, for her invaluable suggestions and her unwavering
commitment. This book would not have been written without
her dedication.
Malkiat S. Johal
Claremont, California
[email protected]
xvii
About the Author
xix
One
A Brief Introduction
to Nanoscience
1
2 Understanding Nanomaterials
~10 nm
~~10 nm
10 nm
1.3 SELF-ASSEMBLY
In general there are two strategies for designing molecular sys-
tems. A “top-down” approach (or step-wise design) describes
the breaking down, or decomposing, of a system to construct
the material of interest or to gain insight into its compositional
subunits. Nanoscale materials are carved into shape by physi-
cal methods such as lithography. In contrast, a “bottom-up”
approach describes the piecing together, or synthesis, of
f ragments to form the larger molecular system. In the most
fundamental sense, nanotechnology refers to the ability to con-
struct molecular assemblies from the bottom up, using current
methods and tools to make useful products. The bottom-up
approach may take advantage of specific chemical reactions or
may involve intermolecular interactions between molecular
fragments. Self-assembly, which is one of the important terms
used in nanoscience, describes a process in which a collec-
tion of disordered building blocks (molecules or nano-objects)
come together to form an organized structure. One common
example is the crystallization of small ions into a definite lat-
tice structure. Techniques such as x-ray diffraction show that
such highly organized structures are comprised of a repetition
of smaller “unit cells” of nanoscale dimension. Self-assembled
structures may also result from just a few molecules, but the
cooperative interplay between these molecules within the
structure may impart a specific function or property.
It is also worthwhile to note that self-assembly may lead to
discrete or extended entities. A discrete entity is well defined
in terms of the number of molecules it contains. A simple
example is the dimer that is formed when two acetic acid mole-
cules interact through hydrogen bonding (Figure 1.2). Another
good example, which is discussed in detail later, is a surfac-
tant “micelle”—a stable, often spherically shaped aggregate
comprised of a few hundred molecules. An extended entity
is undefined in at least one dimension. An example is a thin
film with a well-defined thickness, say, one molecule thick,
but with undefined length and width. A polymer molecule,
in which the exact number of monomeric building blocks is
O H O
C CH3
CH3 C
O H O
Figure 1.2 The acetic acid dimer. The dashed lines represent hydrogen bond interac-
tions (discussed in Chapter 2) between the carbonyl oxygen and the hydroxyl hydrogen. The
dimer is an example of a discrete entity.
6 Understanding Nanomaterials
Intermolecular Interactions
and Self-Assembly
CHAPTER OVERVIEW
Nanostructures assemble, often spontaneously, from simple
molecular building blocks. It is therefore important to begin
this chapter with a discussion of the forces between such mol-
ecules. The types of intermolecular interactions (for example
ion-ion, ion-dipole, dipole-dipole, dipole-induced dipole,
London forces, hydrogen bonds, and electrostatic forces) will
ultimately determine the degree and type of intermolecular
aggregation, as well as the structure of the resulting aggregate.
Such interactions are examined in both bulk media and on sur-
faces. This chapter concludes with some coverage on electronic
structure, and how simple quantum mechanical models can
be applied to predict some optical properties of nanomaterials.
In particular, conjugation in simple organic molecules is used
to make important connections between electronic structure,
intermolecular interactions, and molecular self-assembly.
2.1 INTERMOLECULAR FORCES AND
SELF-ASSEMBLY
This section introduces selected fundamental physical ideas
relating to the assembly and properties of nanomaterials in
order to provide a sufficient background for understanding
subsequent chapters. Intermolecular interactions play a central
role in surface chemistry and the process of self-assembly, both
of which affect the structure and properties of nanomaterials.
Such interactions also determine the properties of surfactants,
influence adsorption phenomena, and even affect the interaction
between molecules and electromagnetic radiation.
Self-assembly is the process during which molecular frag-
ments spontaneously and often reversibly organize themselves
into nanomaterials. The organization of these molecular
9
10 Understanding Nanomaterials
dV ( r )
F (r ) = − (2.1)
dr
q1q2
V (r ) = (2.2)
4π ε 0 r12
dV ( r ) q1q2
F (r ) = − = (2.3)
dr 4π ε 0 r122
12 Understanding Nanomaterials
= − 8.36 × 10−19 J
where 1 J = 1 kg m2 s –2 and 1 C = 1 A s.
Na+ Cl–
Figure 2.1 A schematic depiction of Na+ and Cl– “in contact” with each other.
Intermolecular Interactions and Self-Assembly 13
δ−
δ+ δ+
H H
Figure 2.2 The electronegative oxygen pulls electron density away from the hydrogen
atoms, leaving them with a partial positive charge δ+. Each of the hydrogens has a dipole
moment that points toward the partial negative charge of the oxygen, resulting in a net
dipole moment that passes through the O and bisects the Hs.
14 Understanding Nanomaterials
μ = qL (2.4)
qion μ cos θ
V ( r , θ) = (2.5)
4π ε 0 r122
+= qL
q+ q–
Figure 2.3 The dipole moment µ between two charges q+ and q– separated by a dis-
tance L is calculated as µ = qL.
q+
Ion θ µ
L
r12
q1 = ze
q–
μ 1μ 2
V ( r , θ1, θ2, φ) = −
4π ε 0 r123
(2 cos θ1 cos θ2 − sin θ1 sin θ2 cos φ (2.6))
q+1
q+2
θ1 φ θ2
µ1 L1 L2 µ2
r12
q–2
q–1
– +
δ– – + δ+
+ – +
– +
–
Figure 2.6 As a cation approaches a polarizable atom or molecule, its electric field
produces a distortion of the electron cloud surrounding the polarizable atom or molecule.
The proximity of the cation to the molecule results in an effective charge separation and
produces an induced dipole in the molecule.
18 Understanding Nanomaterials
μ induced = α E (2.7)
F q
E (r ) = = (2.8)
q 4π ε 0 r 2
−q1α
V (r ) = (2.9)
2(4π ε 0 )2 r124
The center of a sodium ion (Na+) is located 0.35 nm from the cen-
ter of a gold atom. If the atomic radius of gold is 144 pm, by what
percentage of its atomic radius would the electron cloud of the gold
atom be shifted due to the presence of the sodium ion?
Solution The electric field induced by Na+ at a distance 0.35 nm
from its center is calculated using Equation 2.8.
J
= 1.18 × 1010
Cm
Then, using Table 1.1, the dipole moment induced in the gold
atom is
⎛ J ⎞
μ = α E = 4π ε 0 (5.8 × 10−30 m3 ) ⎜1.18 × 1010 = 7.61 × 10−30 C m
⎝ C m ⎟⎠
μ 7.61´10−30Cm
L= = = 4.75 × 10−11m = 47.5 pm
q 1.602 × 10−19 C
μ (3 cos2 θ + 1)1/2
E ( r , θ) = (2.10)
4π ε 0 r 3
− μ 12 α (3 cos2 θ + 1)
V ( r , θ) = (2.11)
2(4π ε 0 )2 r126
20 Understanding Nanomaterials
−3 α 2I −Cdispersion
V (r ) = = (2.12)
4(4π ε 0 ) r126
2
r126
−3 α1α 2 I 1I 2
V (r ) = (2.13)
2(4π ε 0 )2 r126 (I 1 + I 2 )
Figure 2.7 Atoms in crystal lattices can often be modeled as tiny, hard spheres in order
to calculate their atomic radius. X-ray or neutron diffraction methods can then be used to
experimentally determine the atomic radius.
Intermolecular Interactions and Self-Assembly 23
∞
⎛ σ⎞
V (r ) = ⎜ ⎟ (2.14)
⎝r⎠
n
⎛ σ⎞
V (r ) = ⎜ ⎟ (2.15)
⎝r⎠
V(r) 0 r V(r) 0 r
σ σ
(a) (b)
Figure 2.8 (a) The hard sphere model of overlap repulsion. r is the intermolecular dis-
tance and σ is the molecular diameter. (b) The soft sphere model of overlap repulsion. Note
that r can assume some values slightly smaller than σ without V(r) becoming infinitely large,
as is the case with the hard sphere model.
24 Understanding Nanomaterials
⎡⎛ σ ⎞ 12 ⎛ σ ⎞ 6 ⎤
V ( r ) = 4ε ⎢ ⎜ ⎟ − ⎜ ⎟ ⎥ (2.16)
⎢⎣⎝ r ⎠ ⎝r⎠ ⎥
⎦
σ
V(r) 0 r
–ε
Figure 2.9 The Lennard-Jones total intermolecular potential curve. –ε is the minimum
energy. r is the intermolecular distance.
Intermolecular Interactions and Self-Assembly 25
rhydrogen bond
H H
Figure 2.10 The hydrogen bond between two water molecules. The electronegative oxy-
gen atom pulls much of the electron density surrounding the hydrogen atom to itself, giving
the oxygen a large partial-negative charge and leaving the hydrogen atom with a partial-
positive charge and very little electron density. The oxygen atom of a neighboring water
molecule can therefore approach much closer to the hydrogen atom than would normally be
possible for a dipole-dipole interaction.
Intermolecular Interactions and Self-Assembly 27
–
–- x
–
Diffuse electrical
double layer –
– –
– –
– – –
– – – –
Stern/Helmholtz – –
–
layer – – – – – – – – – –
+ + + + + + + + + + + + + + + + + + + + + 0
Positively charged surface
Figure 2.11 The Stern/Helmholtz layer and the diffuse electrical double layer. Ions
within the Stern/Helmholtz layer are bound to the surface, although generally not rigidly.
d 2Ψ ⎛ e ⎞
dx 2
= −⎜
⎝ ε ε 0 ⎟⎠ ∑z ρ
i
i io (
exp − zieΨ( x )/kT ) (2.18)
where zi is the valency of the i-th electrolyte (i.e., +1 for Na1+),
e is the standard unit of charge, k is Boltzmann’s constant,
and T is temperature. Ψ(x) is the electrostatic potential at a
distance x away from the surface. The zero of the potential can
Intermolecular Interactions and Self-Assembly 31
d 2Ψ ⎛ e ⎞
=
dx 2 ⎜⎝ ε ε 0 ⎟⎠ ∑ z ρ ( z eΨ(x )/kT )
i
i io i (2.19)
d 2Ψ
= κ 2 Ψ( x ) (2.20)
dx 2
where
1/2
⎛
⎜ ∑ρ
i
ze ⎞
2 2
i0 i
⎟ (2.21)
κ=⎜ ⎟
⎜ ε ε 0 kT ⎟
⎝ ⎠
and has units of m–1. In this case, ρi0 is defined as the number
density of the ith electrolyte in the bulk solution.
The second-order differential equation in 2.20 is called the
Debye-Hückel equation and has a well-known solution of
Ψ( x ) = Ψ0e − κx (2.22)
P( x ) = kT ∑ ρ (x )
i
i (2.23)
Neutral surface
–
–
–
– –
– –
– –
–
– – –
–
– –
–
– – – – – – – – – –
+ + + + + + + + + + + + + + + + + + + + +
Positively charged surface
Figure 2.12 The interaction between a neutral and charged surface is repulsive.
–
– –
–
– – –
– – – –
– –
– – –
– – – – – – – –
+ + + + + + + + + + + + + + + + + + + + +
Positively charged surface
Figure 2.13 The interaction between two surfaces of like charge is repulsive.
2.3 INTERMOLECULAR FORCES AND
AGGREGATION
Supramolecular chemistry is dominated by the host of non-
covalent interactions present in molecular subunits. A simple
illustration can help us understand the interplay between the
various interactions discussed so far, and how this interplay
leads to molecular self-assembly into nanomaterials with a spe-
cific structure. Consider the set of generic molecules shown in
Figure 2.14. For simplicity, only characteristics emphasizing
interactions are shown such as ionic moieties, hydrophobic
regions, dipoles, and hydrogen bonding groups. The mole-
cules represent molecular building blocks and the aggregation
of these individual blocks will be affected by intermolecular
interactions.
The organization of the building blocks into more com-
plex structures will largely be driven by a minimization of
energy and thermodynamic constraints. The latter factor, for
example, can be an entropic gain due to the hydrophobic effect.
Minimization of energy will be achieved by minimizing unfa-
vorable interactions such as bringing two like-charged moi-
eties to the same vicinity. Possible ways the molecular building
blocks may assemble are shown in Figure 2.14(b). Self-assembly
will lead to a three-dimensional aggregate, but if the assembly
Intermolecular Interactions and Self-Assembly 35
Hydrocarbon chain
Dipole
Hydroxyl group –
(a)
+ + + +
– –
(b)
(c)
Figure 2.14 (a) Representation of four molecular building blocks containing various
interacting functionalities. (b) Possible aggregation patterns driven by (i) like-charge repul-
sion, (ii) dipole-dipole and H-bonding interactions, (iii) opposite-charge attraction, and (iv)
strong hydrophobic interactions. Hydrophobic interactions probably play a role in all of these
aggregates. (c) Strong substrate-molecule interactions cause the molecules to tilt in order to
minimize like-charge interactions.
hc
ΔE = hν = (2.24)
λ
h2n2
E= (2.25)
8ma2
e –-
0 a
Figure 2.15 An electron confined on a line between 0 and a. At points 0 and a, the
infinite potential barriers prevent the electron from crossing over.
38 Understanding Nanomaterials
h2
ΔE = E 2 − E1 = (n22 − n12 )
8ma 2
The subscripts 1 and 2 denote lower and upper energy levels, respec-
tively. If a increases from 200 nm to 300 nm, ∆E will decrease by
22/32 or about 45%.
Probability Density
Energy Level
n=5 E = 25h2/8ma2
n=4 E = 16h2/8ma2
n=3 E = 9h2/8ma2
n=2 E = 4h2/8ma2
n=1 E = h2/8ma2
0 a
(a) (b)
Figure 2.16 (a) The allowed energy levels of an electron confined on a line between 0
and a. The energy values are given as a function of the quantum number n. As n increases,
the spacing between the energy levels increases. (b) The corresponding probability density
as a function of the quantum number n. The regions of zero probability represent nodes. The
number of nodes is equal to n – 1 (ignoring the zero probability at 0 and a).
C C C
Unhybridized
p-orbital
C C C
containing
one electron
C-C σ-framework
CH2=CH-CH=CH-CH=CH2
(a)
n=2 n=2
Excitation
n=1 n=1
(b)
Figure 2.17 (a) Overlap of unhybridized p-orbitals on each carbon atom produces a
π molecular orbital resulting in the delocalization of electron density along the hexatriene
chain. (b) The particle in a box model as applied to the hexatriene molecule results in an
energy-level diagram showing electron pairs in three levels. Excitation of an electron to the
n = 4 occurs by the absorption of energy.
2
h2 ⎛ nx2 ny ⎞
E 2D = ⎜ + ⎟ (2.26)
8m ⎝ a 2 b 2 ⎠
2
h2 ⎛ nx2 ny nz2 ⎞
E 3D = ⎜ + + ⎟ (2.27)
8m ⎝ a 2 b 2 c 2 ⎠
h2
E= m2 (2.28)
8π 2I
I = me r 2 (2.29)
42 Understanding Nanomaterials
h2 2 2
ΔE = (2 − 1 ) (2.30)
8 π 2I
C6H6
Unhybridized C C
p-orbital C C
containing
C C
one electron
C-C σ-framework
(a)
E = 4h2 8π2I m = ±2
E = h2 m = ±1
8π2I
E=0 m=0
(b)
Figure 2.18 (a) Overlap of unhybridized p-orbitals on each carbon atom produces a π
molecular orbital resulting in the delocalization of electron density along the benzene ring.
(b) The particle in a box model as applied to the benzene molecule results in an energy level
diagram showing electron pairs in three levels. The two energy levels corresponding to m =
±1 are doubly degenerate.
Intermolecular Interactions and Self-Assembly 43
C C
C π π* C
C C
Figure 2.19 A partial MO energy-level diagram for ethylene emphasizing the π bonding
(Ψ1) and antibonding (Ψ2) MOs. MOs are constructed by the linear combination of the AOs
of ethylene, in this illustration the unhybridized p-orbitals (ϕ1 and ϕ2) on each carbon atom.
The two electrons from each p-orbital are placed in the bonding MO (Ψ1). The MOs due to
σ-bonding are not shown.
Intermolecular Interactions and Self-Assembly 45
C C C C
C C C C
C C C C π π* C C C C
C C C C
C C C C
Figure 2.20 A partial MO energy-level diagram for butadiene emphasizing the π bonding
(Ψ1 and Ψ2) and antibonding (Ψ3* and Ψ4*) MOs. MOs are constructed by the linear combina-
tion of the AOs of butadiene, in this illustration the unhybridized p-orbitals on each carbon
atom grouped together and described as ϕ1 and ϕ2. The four electrons from each p-orbital are
placed in the bonding MOs (Ψ1 and Ψ2). The MOs due to σ-bonding are not shown.
46 Understanding Nanomaterials
Ψ8*
Ψ6*
Ψ4*
Ψ7*
Ψ2 * Ψ5*
Ψ6*
Ψ3*
Ψ4*
Ψ5*
Energy
Ψ3 Ψ4
Ψ2
Ψ1 Ψ3
Ψ2
Ethylene Ψ1 Ψ2
Ψ1
Butadiene
Ψ1
Hexatriene
Octatetraene
2 4 6 8
Conjugation Length (number of C atoms)
Figure 2.21 The MO energy-level diagram for a series of linear conjugated hydrocar-
bons. The figure shows how ∆E for the π → π* transition changes as the conjugation length
increases.
Intermolecular Interactions and Self-Assembly 47
J-aggregate H-aggregate
Unaggregated
monomers
Blue shift
Red shift
λmax
(a)
(b)
(c)
H H
Figure 2.23 (a) π-π stacking between planar aromatic rings. (b) Structures of anthra-
cene and triphenylene. (c) The edge-face interaction between two planar aromatic ring
systems.
CH3
CH3 CH3 CH3
H3C
CH3
CH3 CH3 CH3
CH3
Na → Na + (solvated ) + e − (solvated )
(c) BraH3N-CH=CH-CH=CH-CH2-CO2Na
(d) O(CH2)10SO3Na
N
N
where X = OH or NO2
(e) CONHC18H37
N
N
N
SO3H
O
H 3C C N N O (CH2)10 SO3–
NC N N O (CH2)10 SO3–
Three
Rudiments of Surface
Nanoscience
CHAPTER OVERVIEW
Surfaces and interfaces occur everywhere in nature, from bio-
logical cells to the vast expanse of oceans. They play a key
role in nanoscience since they are often used as platforms for
the growth of nanomaterials. Understanding nanomaterials
would therefore not be complete without covering important
elements of surface science. Interfaces represent the two-
dimensional plane between two different bulk phases of mat-
ter, such as oil and water. We will describe a surface as an
interface where one of the bulk phases is a gas, usually air.
Physical and chemical processes occurring at such regions
tend to be very different from corresponding processes in the
bulk phase as such processes are confined to a region in which
one dimension is on the nanoscale order. For instance, a spe-
cial interaction known as the hydrophobic effect influences
chemistry at surfaces, often responsible for the formation of
films of nanoscale thickness. In this chapter, a consideration
of hydrophobicity and the surface energy of solids and liquids
will lead to a discussion of contact angles and wetting phe-
nomena. This naturally leads to a discussion of self-assembled
monolayers and adsorption phenomena. An understanding
of how intermolecular interactions influence the adsorption
and aggregation of molecules into nanostructures (such as
micelles) is provided by considering the amphiphilic nature
of surfactant molecules.
55
56 Understanding Nanomaterials
%
B
$
A
Figure 3.1 Bulk and surface interactions between molecules in a pure phase (e.g., H2O).
Bulk molecule A is surrounded symmetrically by its neighbors. Surface-bound molecule B is
surrounded asymmetrically by its neighbors.
(3.1)
dG surf = γdA
58 Understanding Nanomaterials
Hydrophilic
moiety
SO4–Na+
Hydrophobic
(CH2)11 moiety
CH3
Air
Aqueous phase
Figure 3.2 A monolayer of the anionic surfactant sodium dodecylsulfate (SDS) at the
aqueous-air interface. The monolayer is in equilibrium with SDS molecules in the bulk aque-
ous phase. The hydrophobic moieties of SDS are pointing away from the aqueous phase and
the polar headgroups are buried in the aqueous phase. This orientation of the SDS molecules
on the surface stabilizes the air–water interface.
(a)
dh
(b)
free energy given by Equation 3.1, show that surface tension is F/2L.
Prove that the units of surface tension are N m –1.
Solution Work must be done to pull up the frame and create a film
[Figure 3.3(b)].
Work done (dw = F dh) is equal to the surface free energy (γ dA),
where A is the area of the surface. The soap film has two sides or
surfaces, so in this example area A is actually 2A. Also, A is the
length L times the distance moved dh.
Therefore, F dh = γ 2dA = γ 2L dh.
Rearranging gives
F
γ= .
2L
Since γLV and θ can easily be measured (see Chapter 4), the
value of γSV – γSL can be determined. If the liquid completely wet-
ted the solid (θ = 0°), then the value of γSL = 0. This is a hypothet-
ical situation in which there is no “tension” between these two
phases. In fact a plot of cosθ versus γLV is linear with a negative
slope. The line can be extrapolated to the value cosθ = 1 and
the corresponding surface tension measured (on the x-axis).
Cosθ = 1 corresponds to a contact angle of 0°, or complete wet-
ting. The corresponding surface tension is called the critical
62 Understanding Nanomaterials
Liquid
γLV
Solid
(a)
θr
θa
(b)
(c)
Figure 3.4 A liquid drop on a solid surface. (a) The drop is non-wetting with a contact
angle θ. (b) The drop is running down the solid with advancing (θa) and receding (θr) contact
angles. (c) Complete spreading. The solid dark lines in (a) highlight the various interfaces
with different surface tensions: the liquid-vapor tension (γLV), the solid-vapor tension (γSV),
and the solid-liquid tension (γSL).
Figure 3.6 The various wetting states on rough surfaces: (a) Wenzel state, (b) Cassie-
Baxter state, and (c) intermediate wetting.
3.2 ADSORPTION PHENOMENA:
SELF-ASSEMBLED MONOLAYERS
A molecule approaching a surface may experience a net attrac-
tive force and consequently become trapped or confined at the
surface. Such a species is called the adsorbate, and adsorption
is the physical process by which adsorbate molecules accu-
mulate onto a solid surface. Desorption is the opposite pro-
cess in which molecules leave the surface and enter the bulk
phase. The solid surface in question is referred to either as the
substrate or an adsorbent. The former is usually a planar sur-
face and the latter can be a high-surface-area porous solid. The
sticking of a reactant molecule to a surface was first proposed
by Michael Faraday in 1834 as the initial step of a surface-cat-
alyzed reaction. In this section we are interested in the kind
of adsorption that results in a monolayer and has a thickness
on the order of nanometers. The surface itself does not need to
be flat but may be rough or even porous. The extent of adsorp-
tion depends not only upon the types of intermolecular forces
involved (van der Waals, electrostatic, hydrogen bonding), but
also on the surface area; the greater the surface area of the
substrate, the greater the extent of adsorption.
The best adsorbents are those with large total surface areas,
such as silica gel (SiO2, surface area >1000 m2/g) and activated
carbon. Silica gel is commonly used in chromatographic col-
umns to enhance the separation of solute mixtures by taking
advantage of the different degrees of adsorption of the various
components. An interesting example of an adsorption phe-
nomenon occurs in polar stratospheric clouds in the upper
atmosphere, which are highly porous and act as substrates for
the adsorption of gases such as HCl.
In this and the next section we focus on the adsorption of
molecules from solution or gaseous phase onto a solid sub-
strate. However, it is important to appreciate that adsorption
is a surface phenomena and can occur at any interface. For
example, molecules may adsorb from an aqueous solution to
the aqueous-air interface. Liquid-liquid interfaces, such as the
boundary between an oil and water phase, also represent regions
at which adsorption may occur. Surface adsorption plays a cen-
tral role in the formation of nanomaterials. Adsorbents pres-
ent themselves as platforms for the self-assembly of molecules
into nanostructures. Specific examples in which nanomateri-
als are synthesized this way are presented in Chapter 5. At
this point it is worth mentioning that solid substrates can be
chemically modified so that adsorption can be selective. This
Rudiments of Surface Nanoscience 69
CH3CH3CH3CH3CH3CH3CH3CH3CH3CH3
CH3
S S S S S S S S S S S
Gold coated substrate
SH
CH3(CH2)17OH
OH OH OH OH OH OH OH
CH3 CH3 CH3 CH3 CH3 CH3 CH3 CH3 CH3 CH3
S S S S S S S S S S S
C B A
V(r)
0 r
Physisorption
P
ΔE
Chemisorption
Figure 3.9 A one-dimensional Lennard-Jones potential energy curve for the chemisorp-
tion and physisorption of a molecule on a planar surface. A, B, and C represent various
distances between the surface and the molecule.
(a)
(c)
(b)
Bulk Concentration
Figure 3.10 Adsorption isotherms describing (a) Langmuir adsorption, (b) multilayer
adsorption, and (c) cooperative adsorption.
74 Understanding Nanomaterials
ka
( ) ()
A g + S s AS s
kd
() (3.5)
The constants ka and kd are the rate constants for the adsorp-
tion and desorption process, respectively. At equilibrium, the
forward and reverse rates are the same, and so from basic
kinetics,
or
ka [ AS ]
= =K (3.7)
kd [ A ][ S ]
νd = kdθβ (3.8)
νa = ka(1–θ)[A] (3.9)
Rearranging gives
1 kd 1
= 1+ = 1+ (3.11)
θ ka [ A ] K [ A]
number of molecules nN A nN A P N A P P
[ A] = = = = =
V V nRT RT kT
(3.12)
76 Understanding Nanomaterials
1 kT 1
= 1+ = 1+ (3.13)
θ KP aP
V
θ= (3.14)
Vm
In this equation, V is the volume of the gas adsorbed to the
surface, and Vm represents the volume of gas corresponding
to monolayer coverage. Complete monolayer coverage corre-
sponds to θ = 1, and V = Vm. Equation 3.14 can be substituted
into Equation 3.13, giving Equation 3.15.
1 1 1
= + (3.15)
V aPVm Vm
9×107
6×107
7×107
8×107
5×107
1/[V(m3)]
4×107
3×107
2×107
1×107
0
0 1×1012 2×1012 3×1012 4×1012 5×1012 6×1012
1/[P(torr)]
Figure 3.11 A plot of 1/V versus 1/P for the adsorption of N2 on mica at
273.15 K.
3.3 × 10−8 m3
= 1.47 × 10−6 mol
2.24 × 10−2m3mol-1
8.85 × 1017molecules
= 2.21 × 1021m−2
(0.02m)2
78 Understanding Nanomaterials
V cz
θ= = (3.16)
Vm ( ) ( )
1 − z ⎡⎣1 − 1 − c z ⎤⎦
⎛ ΔH 1 − ΔH vap ⎞
c = exp ⎜ ⎟ (3.17)
⎝ RT ⎠
V 1
= (3.18)
Vm 1 − z
C1 > C2 > C3
c1
c2
V/Vm
c3
z = P/P*
Figure 3.12 The effect of changing the parameter c on the BET isotherm.
O–Na+
O S O Anionic (SDS)
+
N(CH3)2Br– Cationic (DDAB)
O–
Zwitterionic
N OS (DDAPS)
+ O
+ +
Br–(CH3)2NCH2 CH2N(CH3)2Br–
CnH2n+1 CnH2n+1
O– O
+ O
N P O
O
O O Zwitterionic
(DMPC)
O
Figure 3.13 The molecular structure of some common surfactants, with the hydro-
phobic and hydrophilic moieties indicated. Note that the hydrocarbon chains are shown
in all-trans conformation. This conformation is rarely adopted in micellar structures or in
adsorbed surfactant films at interfaces. The surfactants shown are sodium dodecylsulfate
(SDS), didodecyl dimethylammonium bromide (DDAB), zwitterionic dodecyl-N,N-dimethyl-3-
ammonio-1-proponate-sulfonate (DDAPS), non-ionic C12E3, cationic Germini surfactant, and
the zwitterionic lipid molecule 1,2-dimyristoyl-sn-glycero-3-phosphocholine (DMPC).
Air-water
interface
(a)
(b)
(c)
(d)
Figure 3.14 Surfactant molecules in the bulk aqueous phase are in equilibrium with a
monolayer at the air-water interface. The population of the molecules increases with increas-
ing bulk concentration (a)–(c). At the CMC (d) a saturated monolayer is formed and micelles
are present in the bulk phase.
Oil phase
Oil-water
interface
Aqueous phase
Figure 3.15 When an aqueous surfactant phase above its CMC is brought in contact
with a pure oil phase, the micelles are able to solubilize small oil droplets. The aqueous phase
is a microemulsion. A monolayer of the surfactant likely populates the oil-water interface.
The diagram also shows a “reverse” micelle in the oil phase containing a water droplet.
above its CMC. As a result the micelles swell with oil and
increase in size (Figure 3.15). This swollen micellar phase
is thermodynamically stable and is known as a microemul-
sion. In essence, tiny oil droplets are solubilized in water.
Microemulsion phases are generally made by putting an
aqueous surfactant phase in contact with an oil phase. The
84 Understanding Nanomaterials
Surface Tension
CMC
Figure 3.16 The drop in surface tension of air-water interface with increasing sur-
factant concentration. The intersection of the horizontal (constant surface tension) and the
sloped lines corresponds to the CMC of the surfactant.
1 dγ
Γ=− (3.19)
RT d ln C
1 dγ
Γ=− (3.20)
2RT d ln C
65 65
60 60 (i)
55 55
(ii)
γ/mN m–1
γ/mN m–1
50 50
Saturated (iii) Saturated
45 monolayer 45 monolayer
40 40
35 35
CMC CMC
30 30
0 1 2 3 4 5 6 7 8 9 10 –2.5 –1.5 –0.5 0.5 1.5 2.5
[CTAB]/mM Log(C)
(a) (b)
Figure 3.17 The surface tension of aqueous CTAB solutions versus (a) concentration
and (b) logarithm of the concentration. The surface tension values were obtained at 20°C
using the Wilhelmy plate method. Beyond the concentration corresponding to the CMC, the
surface tension values become constant, indicating a saturated monolayer at the aqueous-
air interface.
1 1
Nm −1 = Jm−1m−1 = mol m−2
(JK−1mol−1 ) ⋅ K (JK−1mol−1 ) ⋅ K
1 dγ 1
Γ=− =− ( −3.23 × 10−3Nm−1 )
RT d lnC ( 8. 314 JK )(
−1mol−1 298K
)
= 1.30 × 10−6 mol m−2
88 Understanding Nanomaterials
1
σ (m2 /molecule ) =
NAΓ
σ (nm2 /molecule ) =
(109 nmm )2
6.023 × 1023 molecule
mol
s
⋅Γ
nm2 ⋅mol
1.6603 × 10−6 m2 ⋅molecules
=
Γ
nm2 ⋅mol
1.6603 × 10−6 m2 ⋅molecules
=
1.30 × 10 −6 mol
m2
= 1.28nm2 /molecule
∫
w = − ∏ dσ
KT α
Π= −
σ − β σ2
65
60
55
50
45
40
0 0.02 0.04 0.06 0.08 0.1
Concentration (mol/L)
92 Understanding Nanomaterials
Πσ
Z=
kT
This should be equal to 1 at all surface pressures
for an ideal gas.
7. Consider the following plot of surface tension versus
SDS concentration. The squares represent the data
from a purified sample of SDS (>99.99%, recrystallized
thrice from ethyl acetate), and the circles show the
data from a batch used as received from a popular ven-
dor (>99.0%). The latter sample contains trace amounts
of the corresponding alcohol (dodecanol).
80
70
Surface Tension (mN/m)
60
50
40
30
20
0 2 4 6 8 10 12 14
SDS0 (mM)
RT Γ
dγ = − dC
C
K ⎡⎣ A ⎤⎦
θ=
1 + K ⎡⎣ A ⎤⎦
K adC
σΓ =
1 + K adC
K adC
σΓ =
1 + K adC
94 Understanding Nanomaterials
RT Γ
dγ = − dC
C
RT
γ = γ0 −
σ
(
ln 1 + K adC )
⎛ o ⎞
ΔGad RT
γ = ⎜γ0 + ⎟ − ln C
⎝ σ ⎠ σ
Rudiments of Surface Nanoscience 95
dG = –8πr dr γ + ∆P 4πr2 dr
2γ
ΔP =
r
⎛ 1 1⎞
ΔP = γ ⎜ + ⎟
⎝ R1 R2 ⎠
Characterization at
the Nanoscale
CHAPTER OVERVIEW
This chapter provides a broad survey of some of the common
techniques used to characterize nanomaterials. Novel tech-
niques and methodologies to probe nanoscale dimensions are
being developed continually, and many of these are based on
variations of common tools and practices used in surface science
and materials research laboratories. The aim of this chapter is
to provide a modest level of background in a sufficient range of
techniques to allow one to make rational choices in determin-
ing which method or combination of methods are best suited to
characterize a particular feature of a nanomaterial. The meth-
ods introduced in this chapter range from traditional surface
science tools, spectroscopic methods, and gravimetric tech-
niques to more specialized characterization approaches such as
nonlinear optical methods and interferometic techniques. The
chapter focuses on the principles behind the techniques and
the interpretation of data. Specific examples of the application
of these methods to nanomaterials are found in Chapter 5.
97
98 Understanding Nanomaterials
2r
Figure 4.1 A fluid moving through a narrow capillary tube. The distance h the fluid
travels depends on the surface tension (γ) and the contact angle (ϕ). The distance h also
depends on the diameter 2r of the capillary.
rhΔρg
γ=
2cos φ (4.1)
Force balance
h
φ w
Figure 4.2 The Wilhelmy plate method. A platinum or paper plate of length I and width
w is immersed into a fluid to a depth h. The forces acting on the plate are measured using
a force balance. ϕ indicates the contact angle the fluid makes against the plate. If a paper
plate is used, this angle can be reduced to zero.
F = 2(w+t)γcosϕ (4.3)
F
γ= (4.4)
2(w + t )
C ⋅ Δf
Δm = − (4.5)
n
where m is mass, f is frequency, n is the overtone number of
the crystal (n = 1, 3, 5, 7, …), and C is a constant that depends
upon the specific quartz crystal.
A quartz crystal microbalance detects minute changes
in mass by applying an AC potential across a quartz crys-
tal to induce its resonance frequency and then monitoring
104
Crystal oscillating
with frequency
induced by AC
Pure crystal With material After AC potential
potential
oscillation adsorbed on the surface is removed
Time Time
Amplitude
Amplitude
(a) (b)
Figure 4.3 (a) The resonant frequency of a piezoelectric quartz crystal in a QCM system will oscillate more slowly when a material is adsorbed to its surface. Thus
a decrease in the resonant frequency is observed during an increase in mass adsorbed to the surface. (b) Dissipation is a measure of how quickly the QCM crystal stops
oscillating after the AC potential is removed. As such, it describes the relative thickness and rigidity of thin films adsorbed to the surface of the quartz crystal. The decaying
oscillations of the crystal after the circuit is broken obey an exponentially decaying sinusoidal curve. The numerical data of the decay can be fitted to Equation 4.7, which
allows for the calculation of a time constant τ, which in turn can be used to calculate the dissipation of the crystal using Equation 4.8.
Understanding Nanomaterials
Characterization at the Nanoscale 105
Edissipated
D= (4.6)
2π E stored
1
D=
π⋅ f ⋅τ (4.8)
Nano-assembly
4.3 ELLIPSOMETRY
There are many approaches to determining the thickness of a
nanofilm. Measuring the absorbance of light through a nano-
film is one way to determine its thickness. However, as will
be shown, determining thickness from absorbance measure-
ments requires knowledge of both molar absorptivity and the
concentration of the material comprising the film. It may be
challenging to obtain the concentration of the molecules in
such a film.
Dragging a stylus over the surface of the substrate is another
approach to determining film thickness. This method is the
basis of the technique known as profilometry. A profilometer
can be used to measure a film’s thickness and roughness. In
this method, a diamond stylus is moved vertically in contact
with the surface of the nanofilm, then moved laterally across
the surface for a specified distance and specified contact force.
A profilometer can measure small surface variations in vertical
stylus displacement as a function of position. A typical profi-
lometer can measure small vertical features ranging in height
from 10 nm to 50 µm. In fact, most of the world’s surface finish
standards are written for contact profilometers. Unfortunately,
the technique is “invasive” in that the stylus makes contact
with the surface and may well damage the film. There are non-
contact profilometers that use light as a way of measuring the
height of surface features. However, non-contact profilometry
Characterization at the Nanoscale 109
x
x y
y
z
Amplitude
z x
Position
(a) (b)
Figure 4.5 Classical representation of light as an electric field. The solid arrows’ vec-
tors represent the field’s orientation and intensity. The magnitude of the vectors oscillates
in time, and their orientation defines the polarization axis. (a) is a wave representation
of the light and (b) is a linear representation showing to which axis the electric field is
confined.
x + x = x
(a)
x y x y y
z + z = z
(b)
Figure 4.6 (a) Linear and (b) wave representations of linearly polarized light of the
same amplitude but various orientations. The polarizations are along (i) the x-axis and
(ii) the y-axis. Vector addition of (i) and (ii) creates a linear polarization 45° from the
x-axis (iii).
Surface normal
Ep Plane of incidence
Es
φi φr
Figure 4.7 Light reflection off a planar surface. The plane of incidence contains the
incident beam, the surface normal, and the reflected beam. If the polarization of the light is
along the plane of incidence, then it is called p-polarized light. If the polarization vector is
perpendicular to the plane of incidence, then the light is called s-polarized light.
x x
z z
(a)
y y y
90° out of phase
x + x x
(b)
y y y
Phase difference
is not 90° n
x + x x
(c)
Figure 4.8 (a) Two mutually perpendicular electric field components that are 90° out
of phase will combine to generate circularly polarized light. (b) Vector addition of the two
components leading to circularly polarized light. (c) Vector addition of the two components
leading to elliptically polarized light.
Characterization at the Nanoscale 113
φ1
n1
n2 d φ2
φ3
n33
Figure 4.9 The reflection, refraction, and transmission of light through a model mul-
tiple interface system. In our case, the layer of thickness d can be a nanofilm assembled on
a solid support. The ni represent the refractive index of each phase.
114 Understanding Nanomaterials
n2 cos φ1 − n2 cos φ2
r12p = (4.9)
n2 cos φ1 + n2 cos φ2
n2 cos φ1 − n2 cos φ2
r12s = (4.10)
n2 cos φ1 + n2 cos φ2
r12p + r23p e − i 2α
Rp = (4.11)
1 + r12p r23p e − i 2α
r12s + r23s e − i 2α
Rs = (4.12)
1 + r12s r23s e − i 2α
where
⎛d⎞
α = 2π ⎜ ⎟ n2 cos φ2, i = −1
⎝ λ⎠
Δ = δ1 − δ 2 (4.13)
Rp
tan ψ =
Rs (4.14)
Rp
ρ= (4.15)
Rs
Rp
ρ= = tan ψe iΔ = tan ψ (cos Δ + i sin Δ) (4.16)
Rs
Light source
Polarizer
Detector
φi φr
Analyzer
Quarter
wave plate
Sample
Figure 4.10 Layout of a typical ellipsometer. The angle of incidence (ϕi) is typically 70°.
The light source is a helium-neon laser, and the detector is a photomultiplier tube.
118 Understanding Nanomaterials
4
Ellipsometric Thickness (nm)
0
0 1 2 3 4 5 6 7 8 9 10
Bilayer Number
Figure 4.11 Thickness data as determined by ellipsometry for the layer-by-layer con-
struction of a typical polyelectrolyte nanoassembly. The bilayer represents a layer of poly
cation complexed with a layer of polyanion.
⎛n ⎞
θcritical = arcsin ⎜ 2 ⎟ (4.17)
⎝ n1 ⎠
Glass prism
Incident Reflected
light light
Metallic layer
(Au or Ag)
Surface plasmon
wave
Aqueous solution
Figure 4.12 The generation of a surface plasmon wave. When polarized light is shone
at the correct angle through a glass prism onto a metal-water interface, then a surface
plasmon wave is generated that propagates along the interface. This results in a reduction
of intensity of the reflected light.
Characterization at the Nanoscale 121
Sample To waste
Flow cell
Glass prism
Polarized
Detector
light source
Figure 4.13 Typical setup of an SPR instrument. A polarized light source is used to
excite the SPR effect at the gold-water interface. The detector monitors the changes in the
intensity of the reflected light as a function of angle of incidence in order to detect changes
in the SPR angle that may result from the adsorption of a thin nanofilm to the gold sensor
surface. A flow cell allows for the easy exposure of the gold sensor surface to the desired
sample solution.
Characterization at the Nanoscale 123
Evanescent
Cladding region field (~100 nm)
Screen
Light source
Interference fringes
Slit in the far field
Double slit
Figure 4.15 The classic Young’s double-slit experiments. Interference between the two
wavefronts emerging from the double slit produce an interference pattern on a screen placed
some distance away.
Sensor waveguide
Laser Reference waveguide
Silicon oxide
Interference fringes
in the far field
Figure 4.16 Typical architecture of a DPI flow cell. Light enters the stacked waveguide
and upon exiting generates an interference pattern in the far field due to a phase shift. The
phase shift occurs due to the adsorption of material onto the sensor waveguide.
Characterization at the Nanoscale 127
Polarization 2
Response
Time
Maxwell’s
equations
Polarization 2
Polarization 1
Absolute RI (density)
Figure 4.17 (a) Typical representation of the fringe pattern observed in a DPI experi-
ment. (b) The sensor phase response for the two polarization states. (c) Effective refractive
index plots showing a unique solution for TE and TM modes. The point at which the curves
cross gives the absolute refractive index and thickness of the film.
n film − nbuffer
δ= (4.19)
dn film /dc
130 Understanding Nanomaterials
Increasing energy
Wavelength (m)
10–13 10–11 10–9 10–7 10–5 10–3 10–1 101
Microwave
X-ray Ultraviolet
Infrared
Gamma ray
Radio
Visible
X-ray absorption
0.1–100 Å Inner electron energy states
and diffraction
hc
E = hv = (4.20)
λ
hc
E1 − E 0 = hv = (4.21)
λ
Non-radiative
relaxation
Non-radiative
relaxation
Intersystem
crossing
Excited Electronic State
and associated
vibrational states
Triplet state
ΔE2
Figure 4.19 Energy diagrams depicting the transitions that occur during fluores-
cence (a) and phosphoresence (b). Note that the energies of the photons (either those being
absorbed or those being emitted) are identical to the corresponding difference in energy
states of the molecule.
134 Understanding Nanomaterials
⎛I ⎞
A = log ⎜ o ⎟ = − log T
⎝ I ⎠
(4.22)
A = cεl
(4.23)
Thin film
Transparent
substrate
d
Figure 4.20 Transmission mode UV-vis absorption spectroscopy can be used to determine the thickness of a thin nanofilm. Beer’s law is used just as with bulk phase
measurements, but the thickness of the film replaces the path length of the sample cell in the equation, as shown.
137
138 Understanding Nanomaterials
Reference cell
Beam
chopper
Light source
Detector
Sample cell
0.40
n
0.35 HN
SO2
0.30 PAZO
0.25
Absorbance
0.20 N
N
0.15 λmax
0.10 CO2–
OH
0.05
0.00
250 350 450 550 650
Wavelength (nm)
(a)
0.5
0.4
Absorbance at λmax
0.3
0.2
0.1
0
0 1 2 3 4 5 6 7 8 9 10
Bilayer Number
(b)
Figure 4.22 (a) The absorption spectrum of PAZO (structure shown in the inset). λmax
represents the wavelength at which the absorbance is a maximum. (b) A plot of absorbance
at λmax as a function of bilayer number. The bilayer represents a layer of polycation com-
plexed with a layer of polyanion.
Estimate the slope of the line of the graph shown in Figure 4.22 and
use it to predict an absorbance value of a film composed of 15 bilay-
ers. How would you determine the concentration of PAZO in the film
from the PAZO’s molar absorptivity? What other technique would be
useful in this determination?
Solution A linear fit to the data yields a slope of ~0.02 (Figure 4.23).
The intercept is close to zero. The slight negative number in the
intercept is likely due to a baseline shift. Thus the equation of the
line is y = 0.02x, where y = A. When x = 15, the absorbance (A)
is 0.75. From Beer’s law, A = cεl, where l represents path length,
or in this case the thickness of the film. We can estimate l using
ellipsometry, and if we know ε (molar absorptivity), we can obtain
the concentration, c.
0.40
0.35
0.30
0.25
Absorbance
0.20
0.15
0.10
0.00
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
Bilayer Number
Figure 4.23 The absorbance (at λmax) data of a polyelectrolyte multilayer film.
The bilayer represents a layer of polycation complexed with a layer of polyanion.
Characterization at the Nanoscale 141
120 0.20
0.18
(a) (b)
100
0.16
0.14
80
Fluorescence Intensity
0.12
Absorbance
60 0.10
0.08
40
0.06
0.04
20
0.02
0 0.00
350 450 550 650 750
Wavelength (nm)
Figure 4.24 Typical absorption (a) and fluorescence emission (b) spectra of PPV, an
optically active polymer. The fluorescence intensity is in arbitrary units. The molecules are
excited at ~475 nm (the wavelength corresponding to the peak absorbance) and emits light
at ~575 nm (the wavelength corresponding to the peak fluorescence).
Characterization at the Nanoscale 143
g a na2
Φ fa = Φ std 2
(4.24)
g std nstd
Sample cell
Light source
Transmitted light
Emitted light
Detector
Em
Emitted light itte
d ligh
t
Detector
Figure 4.26 Methods for monitoring the fluorescence of samples in a cuvette in solu-
tion phase (a) or in a thin film (b). The detector is strategically placed so that it will inter-
cept the maximum amount of fluorescent light without picking up any of the transmitted or
reflected light.
1 k
ν= (4.25)
2πc μ
m1m2
μ= (4.26)
m1 + m2
Asymmetric Symmetric
Bend
Stretch Stretch
O O O
H H H H H H
O C O O C O O C O
Figure 4.27 The vibrational modes of H2O and CO2 and the IR absorption associated
with each vibrational mode. The symmetric stretch in CO2 does not produce a net change in
the dipole moment of the molecule and is therefore IR inactive.
Characterization at the Nanoscale 147
ν 1 (4.27)
wavenumber = =
c λ
Evanescent field
IRE Crystal
Figure 4.29 A schematic diagram of an ATR setup for an FTIR spectrometer. The eva-
nescent wave that is generated at the surface of the IRE can penetrate the overlying region
up to several microns.
E0 = hν0 E0 = hν0
Characterization at the Nanoscale
Etc.
2nd Vibrational state ΔE
1st Vibrational state
Energy state Ground state ΔE
Figure 4.30 Energy diagrams modeling the Rayleigh and Raman scattering processes. A photon interacts with a molecule, promoting it to a non-quantized virtual
state. The molecule quickly returns to a lower energy state, emitting a scattered photon. If the molecule returns to the same state in which it began, Rayleigh scattering has
occurred. If the molecule returns to a higher or lower vibrational energy state, then Raman scattering has occurred.
151
152 Understanding Nanomaterials
μ = αE (4.28)
μ = αE + βE 2 + γE 3 + … (4.29)
The constants χ1, χ2, and χ3 are known as first-, second-, and
third-order susceptibilities. They are related to the sum com-
ponents of the corresponding hyperpolarizabilities averaged
over orientations of the molecules in the bulk material. The
complete mathematical descriptions of the various χ terms
are ignored here. We can separate each polarization term in
Equation 4.30, such that P1 = χ1E represents the linear polar-
ization, and P2 = χ2E2 represents the second-order nonlinear
polarization, etc. In fact, many of the nonlinear optical effects
used to study nanomaterials are based on the second-order
nonlinear polarization. We focus on this term.
Characterization at the Nanoscale 157
P2 =
χ2
2
{( E 2
1 ) ( ) (
+ E22 + E12 cos 2ω1t + E22 cos 2ω 2t )
(4.32)
( ) (
+ 2E1E2 cos ω1 + ω 2 t + 2E1E2 cos ω1 − ω 2 t )}
This is an interesting result because of the terms 2ω1, 2ω2,
ω1 + ω2, and ω1 – ω2, which tell us that when two intense light
frequencies (ω1 and ω2) interact with a material, we can pro-
duce resulting light of doubled frequencies (2ω1, 2ω2), and even
resulting light of combined (ω1 + ω2) and subtracted (ω1 – ω2)
frequencies. The doubled frequency production is known as
second-harmonic generation (SHG), and the latter two prod-
ucts are usually referred to as sum-frequency and difference-
frequency generation (SFG and DFG). Two photons can pass
through a material and combine in energy to produce a sin-
gle photon. If we had only one frequency of light, then two
photons would combine to produce a single photon of twice
the energy (and hence frequency) of the incident photon. In
this case, there is no distinction between SHG and SFG.
We are combining two photons to produce one, so the conversion
efficiency of this process has a theoretical upper limit of 50%.
Not all materials generate a second-order polarization, and
those that produce this effect do so at varying efficiency. The
key determining factor is χ2, the second-order susceptibility.
The mathematical properties of χ2 depend on factors such as
molecular orientation, and these properties ultimately deter-
mine the nonlinear optical conversion efficiency. The proper-
ties of χ2 are discussed in the next section.
λ 500 × 10 m
− 9
λ 900 × 10 m
− 9
SHG:
(a) 2(5.996 × 1014) = 1.199 × 1015 Hz, or ~250 nm (deep UV light)
(b) 2(3.331 × 1014) = 6.662 × 1015 Hz, or ~450 nm (blue light)
SFG: 5.996 × 1014 + 3.331 × 1014 = 9.327 × 1014 Hz, or ~320 nm
(UV light)
H H H H H H
C
C C C
(a) Inversion Inversion
C (b)
C C C
C
C
H H
H H H H
H
H
C Inversion C
(c)
C
C
H
H
Figure 4.31 An inversion operation on simple molecules. (a) Ethylene does not have
inversion symmetry because the positions of all the hydrogen atoms are changed after
the operation. (b) All atoms in ethylyne are unchanged after an inversion. This molecule
has inversion symmetry. If ethylyne is on a surface, as in (c), the center of symmetry
is broken. Surfaces represent non-centrosymmetric regions where there is no inversion
symmetry.
π-conjugated “bridge” N
Donor group O
(CH2)10CH3
2ω 2deff
2 2
L sin2( ΔkL /2)
I 2ω = (I ω )2 (4.33)
c 3ε o nω2 n2ω ( ΔkL /2)2
⎛ω ⎞ 2 2⎛ I I ⎞
I SFG = 128π 3 ⎜ SFG K SFG K vis K IR χ2 ⎜ vis IR ⎟ (4.34)
⎝ hc 3 ⎟⎠ ⎝ AT ⎠
Infrared laser
(variable frequency)
Visible laser
(fixed frequency) Sum-frequency
emission
θVIS
Figure 4.33 Typical geometry of an SFG experiment in which a prism is used to couple
the various laser beams onto a surface. The various angles of the IR, visible, and SFG beams
are indicated in the circle.
beams (A), the various geometric Fresnel factors (K), and the
second-order susceptibility.
In practice, in order to generate SFG light, precise incidence
angles of the fixed frequency visible and tunable IR light have
to be used. Furthermore, the emitted SFG light is observed
at a precise angle. This is known as phase matching, and the
appropriate angles can be calculated using Equation 4.35.
Figure 4.33 shows the typical geometry of an SFG experiment,
indicating the various angles shown in Equation 4.35.
H H
H H
C C
H H
Inversion
C C
H H
C H C H
H H
C C
H H
H H
Figure 4.34 Portion of an all-trans alkyl chain. The CH2 are locally centrosymmet-
ric because after an inversion operation, the positions of all atoms remain unchanged.
Therefore, these groups in this conformation will not produce SHG or SFG light. A “kink” in
the chain or a gauche defect will create a non-centrosymmetric environment of CH2 groups,
rendering these groups SFG active.
Figure 4.35 (a) A self-assembled dodecane thiol monolayer at the gold-water interface. An all-trans conformation due to strong inter-alkyl chain hydrophobic
interactions creates a close-packed structure with the terminal CH3 groups pointing toward the water phase. The SFG spectrum in (b) shows only features due to the non-
centrosymmetric CH3 groups. The CH2 groups are centrosymmetric and so do not appear in the SFG spectrum.
165
166
2.6
Water 2.4
Carboxyl groups
2.2
randomly oriented
2
1.8
alkyl chain
containing 1.6
“defects”
Figure 4.36 (a) A self-assembled mercaptododecanoic acid monolayer at the gold-water interface. The large anionic headgroup prevents the monolayer from packing
tightly and creates kinks and gauche defects in the chain. This randomized the headgroups. The SFG spectrum in (b) shows only features due to the non-centrosymmetric
CH2 groups.
Understanding Nanomaterials
Characterization at the Nanoscale 167
4.8.1 Absorption
When x-rays pass through a thin layer of matter, the inten-
sity of the x-rays is diminished as a result of absorption and
scattering. The effect of scattering can be ignored in wave-
length regions at which significant absorption occurs. Each
element has its own absorption spectrum with well-defined
x-ray absorption peaks that can be used to identify it. In the
x-ray region of light, there is enough energy to cause observ-
able changes in the electronic state of a molecule. When x-ray
energy corresponding to the binding energy of core elec-
trons is absorbed, the core electron is ejected from the atom.
This results in an excited ion. There is a higher probability
of this happening when an x-ray beam with energy equal to
the binding energy of the core electron is used. As the x-ray
beam increases in energy, away from the core electron binding
energy, the probability of the corresponding wavelength being
absorbed diminishes, and so the amount of the x-ray beam
absorbed will decrease. If the x-ray beam is too low in energy,
the corresponding wavelength will not be appreciably present
and will not be able to eject the core electron. This will cause
an abrupt decrease in the amount of the x-ray beam absorbed.
The wavelength of x-rays ranges from about 10 to 10 –6 nm, but
conventional x-ray spectroscopy generally uses only wave-
lengths between 2.5 and 0.001 nm because this range contains
the x-rays with energies corresponding to core electron-bind-
ing energies, which differ between elements.
4.8.2 Fluorescence
The excited ion, which is a result of the x-rays’ ejecting of
a core electron, will fluoresce through transitions of elec-
trons in higher energy levels to the vacancy left by the
ejected core electron. These transitions allow the excited ion
to return to its more stable ground state. This fluorescence
is measurable and is often used in nanoscience techniques
168 Understanding Nanomaterials
4.8.3 Diffraction
When x-rays pass through matter, the radiation interacts
with electrons in the matter in such a way that the path of
the x-rays can be altered. This scattering effect of matter on
x-rays is known as diffraction. In a crystal, or any ordered
sample, the x-rays scatter in ways that produce higher-intensity
areas and lower-intensity areas, also termed constructive and
destructive interference, respectively. By comparing these
high- and low-intensity areas, it is possible to determine the
architecture and the ordering in matter at the nanoscale. It
is, of course, slightly more complicated in practice, because
when an x-ray beam strikes an ordered crystal, which can be
thought of as multiple layers of atoms, each subsequent layer
of atoms scatters some of the beam and lets the remainder of
the beam through. So that this diffraction can take place in a
way that allows us to relate it to the materials structure, the
space between layers of atoms in the material must be about
the same distance as the wavelength of the radiation used to
probe it, and the atoms comprising the system must be highly
ordered with few defects.
In order to use the high- and low-intensity areas that result
from shining x-rays through an ordered crystal, a way to relate
them to the structure must be known. Fortunately, in 1912, W.
L. Bragg determined this relationship. The Bragg equation for
constructive interference is as follows:
2 d sin θ = n λ (4.36)
6 10
5
5
4
(nm)
0
3
400
2
300 200
200
1 100
(um) 100 (um)
0 0
nm
Piezoelectric X
transducers
Y
STM tip
Substrate
Figure 4.38 Schematic diagram of the piezoelectric transducers that control the posi-
tion of an STM tip near the sample surface. For many piezoelectric transducers used, a
distance of as little as 1 nm can be affected with a single applied volt. When the STM tip is
positioned sufficiently close to the sample surface and a potential is applied across them
both, a tunneling current will be induced between the tip and the sample. This tunneling
current is the basis of the STM measurement.
I ≈ e −2κh (4.37)
A
B
Cross section:
0.8
0.6
z/Å
0.4
0.2
A B
0
0 5.5 11 16.5 22
x/Å
Figure 4.39 STM image of a graphite surface. Notice the angstrom-level resolution
that is characteristic of STM. (Image from Atamny et al, Phys. Chem. Chem. Phys., 1999, 1,
4113–4118. Reproduced by permission of the PCCP Owner Societies.)
176 Understanding Nanomaterials
Laser
Detector
Cantilever
AFT tip
Substrate
Piezoelectric scanning stage
Figure 4.40 Schematic diagram of an atomic force microscope. The AFM tip is held in
constant force against the sample surface by the cantilever. A laser beam is reflected off the
back of the cantilever to monitor the tip height. The sample is scanned underneath the AFM
tip by a piezoelectric stage.
Characterization at the Nanoscale 177
AFM tip
Gold Thiol
RS SR
Deposition (H-SR) SR
RS
SR
RS
RS SR
SR
Figure 4.41 A method for the creation of a CFM tip. An AFM tip is coated with Au and
then is functionalized using thiol-gold chemistry. The functional group R can then be used
to probe the surface.
178 Understanding Nanomaterials
AFM tip
Delivery
moiety
Molecular
cargo
Cell
membrane
B C
1 µm
20 nm 20 nm
Figure 4.43 Imaging of the nanoneedle before and after the molecular cargo is attached. (a) shows an SEM image of the carbon nanotube affixed to the AFM tip.
(b) is a TEM image of the carbon nanotube shown in (a). (c) shows a TEM image of the nanoneedle after the molecular cargo has been attached (a quantum dot in this case).
(Image taken from Chen et al, 2007, PNAS, 104, 8218–8222. Copyright 2007 National Academy of Sciences, USA. With permission.)
Understanding Nanomaterials
Characterization at the Nanoscale 181
⎛ λ ⎞
δ ≈ 0.61 ⎜ ⎟⎠ (4.38)
⎝ n sin(β )
h
λ=
p (4.39)
1
V =K= mev 2 (4.40)
2
where me is the rest mass of the electron (9.11 × 10 –31 kg) and v
is the electron’s velocity. From classical physics we also know
that p = mv, so using Equation 4.40 we can write
⎛1 ⎞
p = mev = 2me ⎜ mev 2 ⎟ = 2meV (4.41)
⎝2 ⎠
h h
λ= = (4.42)
p 2meV
h
λ= 1
⎡ 2
⎛ V ⎞⎤
⎢2meV ⎜⎝ 1 + 2m c 2 ⎟⎠ ⎥ (4.43)
⎢⎣ e ⎥⎦
Characterization at the Nanoscale 183
⎛ λ⎞
δ ≈ 0.61 ⎜ ⎟ (4.44)
⎝ β⎠
Electron gun
Condenser
Electromagnetic lens
Sample
Electromagnetic lens
⎛ λ ⎞
δ ≈ 0.61 ⎜
⎝ NA ⎟⎠ (4.45)
200 nm 200 nm
(a) (b)
Figure 4.45 SEM images of two NSOM probes that have been constructed by tapering
a fiber-optic cable into a very fine point and then coating that tip with aluminum, leaving
only a very small aperture at the point. The ends have also been flattened using a focused
ion beam. In these NSOM probes the apertures are approximately (a) 120 nm and (b) 35 nm.
(Image reprinted with permission from Veerman et al., 1998, Appl. Phys. Lett., 72, 3115–3117
as shown in Dunn, Chem Rev, 1999, 99, 2891–2927. Copyright 1998 American Insitute of
Physics.)
188 Understanding Nanomaterials
Light source
d<λ
Sample
(b)
Detector
(a)
Figure 4.46 The different operating modes of NSOM. (a) In Transmission mode,
the light travels from the probe through the sample to a detector on the other side. (b) In
Reflection mode the light from the probe is reflected off the sample surface and captured by
the detector. (c) In Transmission-Collection mode, the sample is illuminated from underneath
and captured by the probe, through which the light travels to a detector. (d) In Reflection-
Collection mode, the sample is illuminated externally to the probe and the reflected light is
captured by the probe and channeled to the detector. (e) In Illumination-Collection mode, the
probe is responsible for both illuminating the sample and collecting the light that reflects
from its surface.
Optical
components
Laser
NSOM probe
Fiber optic cable/
feedback mechanism Sample on
piezoelectric stage
Detector
2 µm 2 µm
2 µm 1 µm
2 µm
Figure 4.48 NSOM images of spherical and toroidal liquid crystals suspended in a
polymer. (Image from Mei et al, Langmuir, 1998, 14, 1945–1950 as shown in Dunn, Chem
Rev, 1999, 99, 2891–2927. With permission.)
πD
α= (4.46)
λ
Unpolarized light
Figure 4.49 The scattering of unpolarized light through a sample. The intensity of the
scattered light is measured as a function of the angle (ϕ) between the incident beam and
the scattered beam.
2
n2 ⎛ dn ⎞
K o = 2π 2
λ 4N ⎜⎝ dc ⎟⎠ (4.47)
d2 I
Rφ = 2 (4.48)
1 + cos φ I o
194 Understanding Nanomaterials
2
2π 2n2 ⎛ dn ⎞ dΨ
ΔRφ = RTc (4.49)
λ 4 ⎜⎝ dc ⎟⎠ dc
K o (cm − CMC )
ΔRφ = 3 (4.50)
10 /M + 2B(cm − CMC )
K o (cm − CMC ) −3 1
10 = + 2 × 10−3B (cm − CMC )
ΔRφ M
Intensity 632
Wavelength (nm)
(a)
Intensity
632
Wavelength (nm)
(b)
Figure 4.50 (a) In a simplistic model all scattered light is detected at one wavelength
(the wavelength of the incident light). (b) The actual observed scattered light; the wave-
lengths are distributed around the expected wavelength due to Doppler shifting of the light
from Brownian motion of the particles.
A nanopartical
(e.g., SiO2, 95 nm diameter)
Correlation
A protein of 5 nm diameter
Time
Figure 4.51 A plot of correlation versus time delay between subsequent measurements
for two different particles. As expected, the larger particles stay correlated for a longer
period of time, indicating that they diffuse more slowly.
k BT
dH = (4.51)
3π ηD
As one might expect, the larger the particle, the slower it
diffuses, as shown in the Figure 4.51; this is why the diffusion
rate and diameter have an inverse relationship. The number
of particles observed at each diameter can be plotted versus
198 Understanding Nanomaterials
15 nm 400 nm
Intensity
Characterization at the Nanoscale
Correlation
0
Figure 4.52 A plot of signal intensity versus particle size along with the correlation plot from which such a graph is derived.
199
200 Understanding Nanomaterials
Time 0 5 10 15 20 25 30 35 40 45 50 55 60 65
(min)
∆F 0 0 –2 –5 –7 –15 –25 –42 –55 –60 –63 –64 –65 –65
(Hz)
⎛ ηρ ⎞
Δf = − fo3/2 ⎜ l l ⎟
⎝ πρq μ q ⎠
A B C D
Polymerase
0 0
(b)
∆m/ng cm–2
–100
∆F/Hz
dATP + dGTP
(a)
100
–200 Step 1 Step 3
Step 2
–300
200
0 10 20 30 31 32 33
Time/min.
206 Understanding Nanomaterials
CHAPTER OVERVIEW
This final chapter is written more like a scientific review of
nanomaterials, with the goal of helping the student to transition
from the fundamental textbook-style material to the primary
scientific literature. The previous chapters should have given
the student the necessary scientific background and terminol-
ogy to understand the material presented in this chapter. This
chapter can be thought of as two separate sections: tuning opti-
cal and energetic properties of materials using nanomaterials
(Sections 5.1 to 5.4) and functionalizing surfaces using nanoma-
terials (Sections 5.5 to 5.9), brought together as they both define
useful ways to employ nanotechnology. There are many other
potential applications of nanomaterials, but herein an overview
is given of how to functionally use various technologies, each
of which is constantly evolving the state of nanoscience.
207
208 Understanding Nanomaterials
Figure 5.1 Excitation of donor molecule (D) by blue light (450 nm) and its subsequent
fluorescence of green light (550 nm). The emitted green light is absorbed by acceptor mol-
ecule (A) which finally emits red light (620 nm).
Types and Uses of Some Nanomaterials 209
Distance d
Glass substrate
(a)
Id/I∞
(b)
Id/I∞
do d (nm)
Figure 5.3 Quantum yield profiles illustrating energy transfer from D to A. (a) Green
fluorescence yield from D as a function of distance d. (b) Red fluorescence yield from A as
a function of distance d. As d increases, energy transfer to A is reduced resulting in more
fluorescence from D and less from A.
5.1.2 Photorelaxation
The functionality of the aforementioned dye assemblies
hinges on the differences in absorbance and emission spectra,
or Stokes shift, of the two entities (not to be confused with
Stokes shifts in Raman spectra). As a molecule is subjected to
energy, in this case energy in the form of photons, it enters an
excited, quantized electronic state. This excited state cannot
be sustained—in fact, exciton lifetime is approximately 10
nanoseconds—and the molecule quickly reverts to its ground
state. Upon returning to ground electronic state, the mole-
cule is left with excess energy from the initial excitation, and
must thus emit that energy in several forms. In dyes, energy is
dissipated via the emission of a photon. Visible color is emit-
ted from photorelaxing dyes in very specific cases when the
energy band gap, or energy difference between ground and
excited states, lies between the energies available to photons
with a visible wavelength. Since the band gap and emission
of a photon from an excited molecule can be quantified by
hc
E BandGap = E Photon =
λ (5.1)
Δλ
Fluorescence
Absorption
300 400 500
Wavelength (nm)
E=
hc
=
( )(
6.62 × 10−34 Js 2.998 × 108 m/s
≈ 2 × 10−18 J
)
−9
λ 100 × 10 m
3mε 0c03
τ0 = (5.2)
2Q 2πν20n
3(Nm)ε 0c03 1
τagg = = τ0 (5.3)
2(NQ )2 πν20n N
We can exploit the fact that the excited domain must exist
through a balance between attractive forces of the in-phase
oscillating dipoles (excited portion) and the thermal motion
tending to disrupt such a domain in order to find the num-
ber N of molecules contributing to the excited domain. It
turns out that the binding energy (–∆E/N) in a pair of oscil-
lators decreases the energy of the excited state, making it
214 Understanding Nanomaterials
−ΔE
= kT (5.4)
N
It has been found that ∆E = –0.24 eV. Restating Equation 5.4,
ΔE 0.24eV 1
N =− = ⋅ (5.5)
kT k T
and for room temperature,
3000K
N= = 10 (5.6)
300K
We can now use the number of molecules N contributing to
the excited domain, which we solved for in Equation 5.6, in order
to approximate the fluorescence lifetime from Equation 5.3.
T
τagg = τ0 (5.7)
3000K
300K
τagg = × 5 × 10−9 s = 5 × 10−10 (5.8)
3000K
We can also approximate the speed υ of an exciton through
a monolayer. In order to do so, we must consider the area A
covered by the exciton of width L during τagg:
A = Lexciton ⋅ v ⋅ τagg
(5.9)
A Za
v= =
Lexciton ⋅ τagg Lexciton ⋅ τagg (5.10)
Types and Uses of Some Nanomaterials 215
5.2 NANOWIRES
A nanowire is a nanostructure that has a diameter on the
scale of a nanometer and an unrestricted length. An example
of a nanowire is shown in Figure 5.5. Nanowires have been
synthesized to be as long as one millimeter (10 –3 m), but are
more commonly made to be about one micrometer (10 –6 m).
Regardless, the typical nanowire has a width-to-length ratio
Figure 5.5 A vertical silicon nanowire array on a silicon substrate. (Image provided by
Professor Hari Reehal, London South Bank University.)
216 Understanding Nanomaterials
h
R=
ne 2 (5.11)
5.2.2 Conductivity
As already mentioned, quantum mechanics has great control
over conductivity. Yet conductivity varies significantly
depending on the composition and the physical properties of
the nanowire (width and length). A nanowire’s extremely large
length-to-width ratio results in high levels of resistance. This
resistance is compared to that of a bulk phase wire, composed
of the same substances, whose resistance is much lower due to
the much smaller length-to-width ratio. The nanowire’s nar-
row width contributes in a number of ways to increasing resis-
tance. For one, decreased width increases the relative effect
of the wire’s surface molecules. The molecules at the surface
of the wire are much less tightly bound than molecules found
in the interior of the wire. Also, surface molecules are in con-
tact with far fewer of the wire’s molecules than those found in
the interior of the wire. The overall consequence is that mol-
ecules found at the surface of a wire have greater difficulty
propagating electrons to other wire molecules. Therefore, high
surface-area-to-volume ratios tend to increase the resistance.
The basic concept of width effects on resistance is related to
what is known as the mean-free path.
Mean-free path relates to the distance an electron travels
between subsequent collisions with other moving particles.
Collisions are not favorable for electrons in nanowires since
the electron can be deflected in any direction (Figure 5.6) and
impact the wire’s forward flow of electrons. Collisions occur
most frequently when the width of the nanowire is smaller
than the mean-free path, resulting in frequent collisions and
218 Understanding Nanomaterials
w<λ
(a)
w>λ
(b)
Figure 5.6 A schematic demonstrating the effects of mean-free path. The open circle
indicates the electron and the solid black circles represent larger atoms. w is the width of
the wire, and λ is the mean free path of the conduction electron. (a) Diffusive transport
occurs when the width of the wire is smaller than the mean-free path of the electron, and
(b) ballistic transport occurs when the width of the wire is greater than the mean-free path.
5.2.4 Summary
Nanowires have great potential. As the field grows, we will
begin to see new synthetic techniques capable of producing
A B
500 nm 50 nm
C D
10 nm 100 nm
Figure 5.7 Gold nanowires were synthesized using a solution-based bottom-up tech-
nique utilizing a copolymer template. Images were taken using a scanning electron micro-
scope. (Reprinted with permission from Chen, Jingy, Benjamin J. Wiley, and Younan Xia.
“One-Dimensional Nanostructures of Metals: Large-Scale Synthesis and Some Potential
Applications.” Langmuir 8, 2007: 4120–4129. © 2007 American Chemical Society.)
220 Understanding Nanomaterials
Zig-zag
Arm-chair
Ch
ma2
θ
a2
a1
Figure 5.8 Schematic diagram showing how a hexagonal sheet of graphite can be rolled
to form a carbon nanotube. The properties of the nanotube will depend on the chiral vector
and the chiral angle. The unit vectors are shown in the lower left-hand side of the diagram.
cosθ =
(2n + m)
(
⎡2 n2 + m2 + nm ⎤
⎣ ⎦ ) (5.13)
222 Understanding Nanomaterials
Metal catalyst
(a)
(b)
Figure 5.9 The two suggested models of carbon nanotube growth through chemical
vapor deposition onto a metal catalyst. In both models, the carbon grows around the edges
of the metal catalyst.
226 Understanding Nanomaterials
Inorganic core
Shell material
2–10 nm
Figure 5.10 Structure of a quantum dot. The size and nature of the inorganic core and
the shell material can be varied. Quantum dots are often capped with an organic thin film.
(c)
Intensity (a.u.)
(b)
1/e
(a)
0 t1 t2 50
Time (ns)
Figure 5.11 Comparison of the excited state decay curves (monoexponential models)
between quantum dots and common organic dyes. Line (a) indicates excitation. Lines (b) and
(c) indicate the decay curves for the organic dye and quantum dot, respectively. A measure of
the fluorescence lifetime is indicated by time t1 (for the dye) and t2 (for the quantum dot).
C NH
OH
H
C O
QD capping
OH
OH
O
C
O
ligand TOPO
O
C
NH
C
O
NH
O
C
C
O
NH
C O OH
O=P
C
O=P
P
O
O=
O
P
C OH
C OH
=
O
O
=P
P
PEG
O
O=
O O
C NH Quantum C NH (–CH2–CH2–O–)n
O=P
O=P
dot
O
O=
C O P O=P O
H O C CH
=
O
P
O
=
O=
O=
O=P
O
P
C
C
P
P
OH
OH O
O
C
C
Amphiphilic
NH O
NH
C
O OH polymer coating
O
C
O
OH
C OH
C
O
O
C
OH
HN
Figure 5.12 The structure of a multifunctional quantum dot probe. Schematic illustra-
tion showing the capping ligand TOPO, an encapsulating copolymer layer, tumor-targeting
ligands (such as peptides, antibodies, or small-molecule inhibitors), and polyethylene glycol.
(Reprinted from Gao, X.; Yang, L.; Petros, J. A.; Marshall, F. F.; Simons, J. W.; Nie, S. In vivo
Molecular and Cellular Imaging with Quantum Dots. Current Opinion in Biotechnology 2005,
16, 63–72. With permission from Elsevier.)
(a)
COOH
2
NH
H
CO
O
CO
O H
N
Quantum H EDAC Quantum
COOH COOH + H2N 2
dot dot
NH2
Antibody
CO
O H
O
CO
COOH
(b)
NH2
OH
CO
CO
OH
OH
NH2
CO
Antibody fragments
(c)
OH
CO
CO
OH
Quantum
Quantum Adaptor dot
+
dot protein
Antibody
CO
OH
OH
CO
(d)
COOH
Binding site
OH
CO
OH
CO
O O
O
COOH Quantum O Quantum
CONH +
dot N N
2– dot
His6
O
His-tagged
CO
OH
O
OH
COOH
CO
Ni-NTA peptide
Figure 5.13 Methods for conjugating QDs to biomolecules. (a) Traditional covalent
cross-linking chemistry using EDAC (ethyl-3-dimethyl amino propyl carbodiimide) as a
catalyst. (b) Conjugation of antibody fragments to QDs via reduced sulfhydryl-amine cou-
pling. SMCC, succinimidyl-4-N-maleimidomethyl-cyclohexane carboxylate. (c) Conjugation
of antibodies to QDs via an adaptor protein. (d) Conjugation of histidine-tagged peptides
and proteins to Ni-NTA-modified QDs, with potential control of the attachment site and
QD:ligand molar ratios. (Reprinted from Gao, X.; Yang, L.; Petros, J. A.; Marshall, F. F.; Simons,
J. W.; Nie, S. In vivo Molecular and Cellular Imaging with Quantum Dots. Current Opinion in
Biotechnology 2005, 16, 63–72. With permission from Elsevier.)
232 Understanding Nanomaterials
Air
Water
Π Π
Air
Water
Π Π
Air
Water
Π Π
Air
Water
Figure 5.14 A Langmuir film. Amphiphilic molecules in chloroform solvent are depos-
ited on water surface. The Langmuir film is formed by confining the amphiphiles at the air-
water interface with movable barriers.
Types and Uses of Some Nanomaterials 233
ΠA = k BT
(5.14)
Π (mN/m)
LC
LE
G
0.1 0.5
A (nm2/molecule)
Figure 5.15 A simplified Π-A isotherm for a simple long-chain fatty acid. Extrapolating
the slope of the S phase to zero pressure enables one to obtain the molecular area per mol-
ecule at zero pressure (see Example 5.2).
234 Understanding Nanomaterials
Use Figure 5.15 to determine the area per molecule for the amphiphile
described by the isotherm.
5.5.2 Langmuir-Blodgett Films
The LB film is subsequently formed by the immersion of a
substrate into the water to break the Langmuir film and
transfer the monolayer onto the substrate. Additional immer-
sions result in the fabrication of the multilayer film, whose
Types and Uses of Some Nanomaterials 235
Air
Substrate
Water
(b)
Air
Substrate
Air
Water
Water
(a) (c)
Figure 5.16 (a) The LB method is used to transfer a film at the air-water interface to a
solid substrate. (b) The downstroke results in the monolayer being transferred to the hydro-
phobic substrate. (c) Withdrawing the substrate generates a bilayer as a Y-type film.
5.6 POLYELECTROLYTES
Polyelectrolytes are polymers, or chains of molecules, that
contain free ions that make them electrically conductive.
Soluble in water, polyelectrolytes become charged when in
solution, and are often countered by a salt ion of opposite
charge. Figure 5.17 shows some common polycations and
polyanions. The amount of charge on a polyelectrolyte deter-
mines whether it is classified as strong or weak. Strong poly-
electrolytes are fully soluble, whereas weak polyelectrolytes,
with fractional charge, are only partially soluble. Typically,
n
HN
SO2
N
N
n
HO O
COO – N n
Na + CH2 CH
OH n SO3– Na+ H
(a) (b) (c) (d)
CH2 CH2
n
n
+ Cl–
N
NH3+ Cl– H3C CH3
(e) (f)
Figure 5.17 Some common polyelectrolytes. Polyanions are (a) PAZO (poly{1-[4-
(3-carboxy-4-hydroxyphenylazo)benzenesulfonamido}-1,2-ethanediyl, sodium salt]), (b)
PAA [poly(acrylic acid)], and (c) PSS [poly(styrenesulfonate)]. Polycations are (d) PEI
[poly(ethylenimine)], (e) PAH [poly(allylamine hydrochloride)], and (f) PDDA [poly(diallyldimethyl
ammonium chloride)].
238 Understanding Nanomaterials
Spontaneous
self-assembly
Bound counterions
– – – – – – + – –
Negatively charged + – + ++
––––––––––––––––– – + + +
+
+ + + + + + Electrostatic
Glass substrate –––––––––––––––––
attraction
– –
– –
– – –
Process –
repeated
–
-
– – – – – – –
- -- -
- - –
+ + + +
Multilayer
+ +
+
+ + + + + +
– –– –– –– – –– – – –– – – –
Entropy gain
Bilayer
Figure 5.19 Polyelectrolyte self-assembly. The polycation adsorbs from aqueous solution to the negatively charged substrate. This is followed by exposure of this film
to an aqueous solution of the polyanion. Adsorption of the polycation to the polyanionic film results in a bilayer film. The process can be repeated many times to produce a
multilayer nanofilm.
Understanding Nanomaterials
Types and Uses of Some Nanomaterials 241
Polyelectrolyte multilayers n
Hn
SO2
Fuzzy nanoassemblies
N
N
8 nm
(80 Å)
CO2–
OH
Polyanion PAZO
0.8 nm
(8 Å)
Polycation N+ n PEI
H H
Figure 5.20 Dimensions of a typical multilayer constructed using the polycation PEI
and the polyanion PAZO. The thickness of a bilayer as measured by ellipsometry is around 1
nm. A 10-bilayer film has a thickness between 100 and 150 nm. Interlayer interpenetration
is illustrated as the “fuzzy” nanoassembly on the right. The degree of interpenetration is
drastically underexaggerated.
100μ-1 nm pinhole
Hydrophobic
support
(a)
(b)
Figure 5.21 (a) Illustration of a black lipid membrane. (b) The formation of a folded
lipid bilayer. (Reprinted from Castellana, E. T. and Cremer, P. S. Solid Supported Lipid Bilayers:
From Biophysical Studies to Sensor Design. Surface Science Reports 2006, 61.10: 429–444.
Copyright 2006, with permission from Elsevier.)
Solid support
Pull
Push
Figure 5.23 Common methods for forming supported bilayers. (a) Langmuir-Blodgett
technique to deposit monolayer followed by pushing the substrate horizontally through
another lipid monolayer. (b) Vesicles in solution adsorb and fuse to the surface to form a
bilayer. (c) A combination of (a) and (b). (Reprinted from Castellana, E. T. and Cremer, P. S.
Solid Supported Lipid Bilayers: From Biophysical Studies to Sensor Design. Surface Science
Reports 2006, 61.10: 429–444. Copyright 2006, with permission from Elsevier.)
10
0
–10
∆f3 (Hz)
–20
–30
–40 Rinse
Rinse
–50
–60 Rinse
∆D3 (10–6)
2.5
2
1.5
1
0.5
0
0 5 10 15 0 50 100 150 0 5 10 15 0 5 10 15 20
Time (min)
A–DOPC/DOPS (1:2) B–DOPC/DOPS (1:1) C–DOPC/DOPS (4:1) D–DOTAP
Figure 5.24 Lipid deposition pathways measured by QCM-D on silica. (a) Vesicles do
not adsorb. (b) Vesicles adsorb and remain intact, forming a supported vesicular layer (SVL).
(c) Vesicles adsorb and remain initially intact. At high vesicular coverage an SLB is formed.
(d) Vesicles adsorb and rupture instantaneously to form an SLB. (Reprinted with permission
from Richter, R. P., Berat, R., and Brisson, A. R. Formation of Solid-Supported Lipid Bilayers:
An Integrated View. Langmuir 2006, 22.8 : 3497–3505. © American Chemical Society.)
Figure 5.25 Peripheral domains of transmembrane proteins can become immobilized and denatured on a solid support. A polymer cushion can help shield the protein
from the substrate. (Reprinted from Castellana, E. T. and Cremer, P. S. Solid Supported Lipid Bilayers: From Biophysical Studies to Sensor Design. Surface Science Reports
2006, 61.10: 429–444. Copyright 2006, with permission from Elsevier.)
251
252 Understanding Nanomaterials
Light
(a) (b)
(d)
(c)
Figure 5.26 Principle of FRAP. (a) The bilayer is uniformly labeled with a fluorescent
tag. (b) The label is selectively photobleached. (c) The intensity within the photobleached
region is monitored as a function of time. (d) Eventually uniform intensity is restored.
⎡ ⎛ 2τ ⎞ ⎛ 2τ ⎞ ⎤
()
f t = e −2τD /t ⎢ I 0 ⎜ D ⎟ + I 1 ⎜ D ⎟ ⎥ (5.15)
⎣ ⎝ t ⎠ ⎝ t ⎠⎦
f (t)
Figure 5.27 Illustration of a typical FRAP recover curve with corresponding images.
The solid line represents a fit to the data using a 2D diffusion model equation.
A B C D E
F G H I J
10 µm
Figure 5.28 FRAP images showing the fusion of two vesicles. The red and green color is shown in grayscale. (Reprinted from Lei, G. and MacDonald, R. Biophysical
Journal, 2003, 85, 1585–1599, Biophysical Society, Elsevier. Copyright 2003, with permission from Elsevier. See Lei and MacDonald for the color image.)
Understanding Nanomaterials
Types and Uses of Some Nanomaterials 255
Excitation Emission
Assembly Fluorophore
SiO2 layer
Silicon substrate
Figure 5.29 Basic principles of a FLIC experiment. The resulting intensity is a function
of the path length difference between the direct and reflected light paths, which is a func-
tion of the height of the fluorophore above the reflective plane.
for the excitation and the emission light, each of which must be
considered separately. The typical setup makes use of a silicon
wafer as the reflector. A transparent oxide (SiO2) layer acts as a
spacer. The system of interest (i.e., a lipid bilayer with fluores-
cent probes) is placed on top of the oxide layer.
FLIC can also be used to look at the topography of a second-
ary membrane. A secondary lipid bilayer can be assembled atop
a solid supported bilayer via the rupture of giant lipid vesicles
with diameters that approach tens of microns. The secondary
membrane is separated from the primary membrane by a con-
fined layer of water and is therefore free to exhibit nanometer-
scale height fluctuations.
dθ/dt = k (1 – θ) (5.16)
O SH
HSCH2(CH2)gCH2 OH Fe
O
SH
HS O CH3
HS
X X X X X
S S S S S S
Au Au
Figure 5.30 Upper: Some examples of thiols used to functionalize a gold surface.
Lower: Thiol bound to a gold substrate via a strong Au-S bond. The grey tube is typically a
hydrocarbon chain of specified length. X represents some surface functional group.
Types and Uses of Some Nanomaterials 257
Si
Cl
Cl Cl
Si
O O O OH
OH OH OH OH SiO2
SiO2
5.9 PATTERNING
As the field of nanotechnology expands, the demand for sur-
face manufacturing techniques that are cheaper, more flexible,
Types and Uses of Some Nanomaterials 259
O O
–OH or H+/H O
O 2
Si OH
O Si EtOH O
O O
Hydrolysis
O O
O O
O Si OH
OH OH Si
Si + Si O
O HO –H2O
O O
O O
Condensation
HO
HO HO HO O
O O
O O Si Si
O Si Si Si *
* Si
O
O O O
O O O O
Si
Si Si O O
O O O OH
OH O O *
O O Si Si
O Si Si Si
Si
* O
O O O
O O HO HO
Si
Si *
* *
*
Silica Network
Figure 5.32 The sol-gel process. Silanes are highly reactive to OH groups and
(R-O)3Si-OH molecules, and can self-react in a condensation process leading to the silica
network containing nanopores.
λ
α=k
φ (5.17)
Mask
Types and Uses of Some Nanomaterials
Substrate
Photoresist
Figure 5.33 Patterning a substrate using UV light. The light passes through a photomask and degrades exposed regions of the photoresist.
261
262 Understanding Nanomaterials
ODT in ethanol
PDMS
(a)
Figure 5.34 (a) A solution of ODT is placed over the PDMS stamp. The solvent is
removed and ODT molecules assemble on the stamp. (b) The PDMS stamp with the ODT “ink”
is placed on a gold substrate. When the stamp is removed, the ODT molecules in contact with
the gold surface chemisorb to the substrate, transferring the pattern from the stamp to the
gold via the ODT ink.
Types and Uses of Some Nanomaterials 263
as ink, the characteristics of the AFM tip, the write speed, and
the pressure used can all affect the pattern ultimately produced
on the surface. For example, altering the humidity can change
the size and structure of the meniscus where ink is deposited,
resulting in a larger or smaller area of the surface onto which
the same amount of ink should have been deposited.
“Fountain-pen” methods, in which the tip is placed within
a nanoscale tube that delivers a constant flow of “ink” mol-
ecules to the tip, are a related surface fabrication method with
similar advantages and drawbacks. The resolution here is
worse than dip-pen methods, as the fountain-pen methods can
produce only patterns with feature widths above 200 nm.
AFM tip hammering nanolithography is yet another varia-
tion on AFM nanolithography. In AFM tip hammering nano-
lithography, the AFM tip is used to imprint the surface. One
paper by Wang et al. (2009) gives an example of the technique
being used on a polystyrene-block-poly(ethylene/butylenes)-
block-polystyrene (SEBS) copolymer. The AFM tip deforms
the polystyrene sphere component of the copolymer, result-
ing in an imprinted area. Embossing as well as imprinting is
possible; one simply imprints the area around the area to be
embossed to get a raised structure.
This technique has several advantages. For example, the
surface imprint is reversible. Depending on the copolymer
used as the surface, heating it at high temperatures that allow
for reorganization can erase the imprint. This does not mean
that the imprint is unstable at all temperatures. At room tem-
perature, below the threshold energy required for reorganiza-
tion of many molecules, the patterns are fairly stable. For the
SEBS copolymer, it was found that the imprint is erased in
about 5 minutes at 70°C and that around 50% of the original
surface contrast remains after 70 days at 25°C. More stable
copolymers would require a higher temperature to reverse
the patterning and clear the imprint, but they would likely
also be more stable for longer at room temperature. Also,
the AFM tip hammering nanolithography has a very good
feature resolution and is currently able to produce patterns
with features as small as 13 nm for imprinting and 18 nm
for embossing. The combination of feature resolution, surface
stability, and reversibility makes the method attractive for
high-density data storage. The method also has drawbacks,
the primary one being surface specificity. The surface chosen
must be imprintable with a reasonable level of force, but also
stable enough that the imprinting will survive a desirable
amount of time.
266 Understanding Nanomaterials
5.9.5 Summary
Nanolithographic methods can be used individually or com-
bined with other fabrication methods to amplify the range of
possible surface structures and their uses. For example, pho-
tolithography can be used to create a pattern on a surface on
which nanowires can be grown. The result is multiple levels
of roughness, a surface property with practical applications in
superhydrophobicity. Nanolithography is thus a quickly devel-
oping field encompassing many techniques for the fabrication
of very useful nanoscale patterned surfaces.
(a) G
G G
G
A T C A T C A T C
T G T T G T T G T
G A G G A G G A G
C C A C C A C C A
(b)
G
G G
G
A T C A T C A T C
T G T T G T T G T
G A G G A G G A G
C C A C C A C C A
(c)
G
G
A T C A T C A T C
T G T T G T T G T
G A G G A G G A G
C C A C C A C C A
(d)
A B C
O O O O O O
LI LI LI LI LI LI
G G G G G G
O O O O O O
A A B A B C
Figure 5.35 Methods of fabrication. (a) In situ with masks. (b) In situ with mirrors.
(c) In situ printing. (d) Printing pre-synthesized oligonucleotides. (Reprinted from Dufva,
M. Fabrication of High Quality Microarrays. Biomolecular Engineering 2005, 22, 173–184.
Copyright 2006, with permission from Elsevier.)
Hydration Layer
Silica Substrate
100X
UV (184–275 nm) Zoom-in
(b) (d)
PBS 2 um
100 um
180 (e)
160
140
120
100
Types and Uses of Some Nanomaterials
80
(c) 60
PBS
Figure 5.36 Direct patterning of void arrays within bilayer membranes using UV photolithography. (a) A schematic diagram of the key process steps. (b) Bright-field image,
where bright squares are quartz and dark regions reveal the chrome background. (c) Epifluorescence images revealing resultant fluorescence patterns. (d) A high-magnification
fluorescence image of the sharp boundary between the UV-exposed and UV-protected regions. (e) Fluorescence intensity profile across an arbitrarily chosen line spanning four
alternating UV-exposed and unilluminated bilayer regions. (Reprinted with permission from Yee, C. K., Amweg, M. L., and Parikh, A. N. Direct Photochemical Patterning and
269
Refunctionalization of Supported Phospholipid Bilayers. Journal of the American Chemical Society 2004, 126.43:13962–13972. © American Chemical Society.)
270 Understanding Nanomaterials
5.10.3 Optimization
A great deal of optimization is required before a DNA microar-
ray can be used effectively. The quality of a DNA microarray
is greatly influenced by the method in which it was fabricated.
For example, spot density, the number of spots that can be
placed in a given area of a microarray, varies greatly among in
situ, non-contact printing and contact printing methods. Using
in situ methods, spot sizes with diameters less than 10 μm
have been created, whereas printing methods can create spot
sizes between 20 and 30 μm. Also affecting spot density is the
array geometry. It is important for spots to be arranged in a
geometry that maximizes spot density while preventing any
overlap; additionally, precise array geometry is significant for
future data analysis. Currently, our ability to create high spot
density arrays has surpassed our ability to accurately interpret
fluorescence. Spot sizes of approximately 30 μm are the small-
est that result in distinct fluorescence from separate spots. The
creation of higher resolution fluorimeters will be required for
further miniaturization of DNA microarrays.
Spot morphology is another parameter that must be moni-
tored. A homogeneous spot is preferred for data analysis. Poor
spot morphology is less of an issue with light-directed in
situ syntheses. Spot morphology of printing methods can be
improved by controlling humidity and temperature, as well as
by adjusting the spotting buffer.
Perhaps the most important factors to be controlled for
optimization are probe density and hybridized density. These
two factors are closely related, as probe density is the num-
ber of probes in a given spot, and hybridized density is the
fraction of these probes that hybridize when saturated with
target molecules. In order to get the greatest fluorescent signal,
hybridized density must be maximized; however, a number
of issues arise. If probe density is too high, probe molecules
may interact with other probe molecules on the surface and
become unable to hybridize. Additionally, high probe density
leads to high steric hindrance, which prevents targets from
hybridizing. A number of troubleshooting methods have been
developed to optimize these parameters. In many cases, the
microarray surface and the probe molecules can be modified
to maximize functional probe density.
Another method to optimize hybridized density uses linker
molecules to move the probes farther from the surface to
Types and Uses of Some Nanomaterials 271
5.10.4 Applications
DNA microarrays are excellent tools for studying gene expres-
sion and have valuable uses in both research and industry.
One of the first uses of DNA microarrays was in the study
of E. coli. Samples of stationary-phase E. coli and log-phase
E. coli were analyzed by DNA microarray. The data found from
microarray analysis led to the discovery of multiple growth-
regulating genes and an overall improved understanding of
this significant microorganism (Ye et al., 2001).
Industrially, DNA microarrays are of great significance in
optimizing biocatalysis. Particularly, fermentation processes are
often optimized using DNA microarrays. As biological hydrogen
fuel cells gain more attention, it is likely that DNA microarrays
will be of great importance in optimizing these systems.
DNA microarrays have great potential to impact healthcare.
Intelligent drug design has benefited from the use of DNA
microarrays as it allows drug developers to anticipate adverse
effects or, alternatively, to discover new drug targets. Other
kinds of microarrays are also effective in diagnosing diseases.
DNA microarrays have been designed to screen for inherited
genetic diseases; also, changes in the expression of certain
genes can be attributed to cancer and other chronic illnesses,
allowing for earlier detection. Perhaps the most exciting pos-
sibility with DNA microarrays is the development of personal-
ized medicine. Using a DNA microarray to better understand
an individual’s genotype will allow doctors to more effectively
treat each person, ultimately resulting in a healthier country.
0.50
0.25
Flow 0.00
outlet 1 2 3 4 5 67 8
100 µm
1.00
0.75
0.50
NBD
0.25
0.00
1 2 3 4 5 67 8
Corral (from left) Field-induced separation of
oppositely charged dye-lipids:
Texas Red (–) and DiD (+)
Captured Membrane micro-array
Microarray
substrate
Flow
channels
Silica substrate
Figure 5.38 Schematic of the CFM apparatus and closeup of the CFM print head in
contact with a silica substrate used for bilayer formation. (Reprinted with permission from
Smith, K. A., Gale, B. K., and Conboy, J. C. Micropatterned Fluid Lipid Bilayer Arrays Created
Using a Continuous Flow Microspotter. Analytical Chemistry 2008, 80.21: 7980–7987.
© American Chemical Society.)
274 Understanding Nanomaterials
5.10.6 Summary
The DNA and lipid microarrays are fairly new technologies
whose potential utility has yet to be reached. Widespread
use is currently limited by the length of time involved in
oligonucleotide synthesis and array fabrication techniques,
which make the cost of a microarray experiment rise signifi-
cantly. As technology advances, microarrays are likely to be
an essential tool in any biochemist’s or molecular biologist’s
arsenal.
CITED REFERENCES
Baddour, C. E.; Breins, C. Carbon Nanotube Synthesis: A Review.
Int. J. Chem. React. Eng. 2005, 3, 1–20.
Blodgett, K. B. Films Built by Depositing Successive Unimolecular
Layers on a Solid Surface. J. Am. Chem. SOC. 1935, 57,
1007–1022.
Castellana, E. T.; Cremer, P. S. Solid Supported Lipid Bilayers:
From Biophysical Studies to Sensor Design. Surf. Sci. Rep.
2006, 61(10), 429–444.
Crossland, E. J.; Ludwigs, S.; Hillmyer, M. A.; Steiner, U.
Freestanding Nanowire Arrays from Soft-Etch Block
Copolymer Templates. Soft Matter. 2007, 3, 94–98.
Dai, H. Carbon Nanotubes: Opportunities and Challenges. Surf.
Sci. 2002, 500, 218–241.
Drexhage, K. H.; Zwick, M. M.; Kuhn, H. Sensitized Fluorescence
after Light-Released Energy Transfer Through Thin Layers.
Ber. Bunsen-Ges. Phys. Chem. 1963, 67, 6267.
Dufva, M. Fabrication of High Quality Microarrays. Biomol.
Eng. 2005, 22, 173–184.
Gao, X.; Yang, L.; Petros, J. A.; Marshall, F. F.; Simons, J. W.; Nie,
S. In Vivo Molecular and Cellular Imaging with Quantum
Dots. Curr. Opin. Biotechnol. 2005, 16, 63–72.
Types and Uses of Some Nanomaterials 275
n1 sin θ1 = n2 sin θ2
(A.1)
⎛n ⎞
θcritical = arcsin ⎜ 2 ⎟ (A.2)
⎝ n1 ⎠
EVANESCENT WAVES
A description of total internal reflection using classical phys-
ics says that the incident light is completely reflected from the
surface between the two substances. In reality, however, some
of the energy from the incident light actually penetrates the
second medium to a small extent. This “portion” of the light
that enters the other medium is called an evanescent wave. The
evanescent wave (or evanescent field) decays rapidly, moving
281
282 Appendix
− x /d p
E x = E 0e (A.3)
λ
dp =
2
2 ⎛ n2 ⎞ (A.4)
2πn1 sin θincidence − ⎜ ⎟
⎝ n1 ⎠
283
284 Glossary
Free electron model: In this model, the electron does not exist
as a discrete particle moving along the line. Rather, it
resembles a standing wave whose exact form depends
on the value of n.
Gravimetry: The measure of the strength of gravitational
fields.
H-type aggregate: One-dimensional molecular assembly in
which the dipole moments are aligned parallel to each
other but perpendicular to the line joining their cen-
ters. This is sometimes referred to as the “face-to-face
arrangement.”
Hard sphere model: A way to determine atomic radius by
assuming atoms in a solid are hard spheres and pack
closely together.
HOMO: Highest occupied molecular orbital.
Hydrodynamic radius: The radius of a sphere that diffuses at
the same rate as the molecule. Because most molecules
are not spherical, this radius is often smaller than the
effective rotational radius.
Hydrogen bonding: Attractive interaction between a hydrogen
atom and another electronegative atom.
Hydrophobic effect: Tendency of non-polar molecules to aggre-
gate in polar solvents to reach a thermodynamically
favored energy state.
Induced dipole interactions: Force existing between a perma-
nent dipole and a neighboring induced dipole.
Interface: The two-dimensional region of space at which two
different phases contact each other.
Interferometry: The study of the ways in which light waves
interact or interfere with each other.
Intermolecular force (F): Forces acting between sets of mol-
ecules, such as hydrogen bonding or dipole−dipole
interactions.
Interpenetration: In polyelectrolyte multilayer films, this
is the tendency of polycation and polyanion to com-
mingle to form highly homogeneous assemblies rather
than distinct, stratified layers. This process shields
the excess charge within the distinct layers, allowing
for tighter packing.
Ion-ion forces: Attractive or repulsive interaction between
ionic species.
J-type aggregate: One-dimensional molecular assembly in which
the dipole moments of the individual monomers are
aligned parallel to the line joining their centers. This is
sometimes referred to as the “end-to-end arrangement.”
288 Glossary
“The writing ... is very fluid. The problems and figures are good. Overall, I
learned a great deal….”
—Professor Lisa Klein, Rutgers University
“I believe the textbook will serve students well in their goal to gain a greater
understanding of why nanoscaled systems are of great interest, how they are
fabricated, and how they are characterized using a wide variety of analytical
instrumentation very commonly found in university and industrial settings.”
—Professor Marcus D. Lay, University of Georgia
Author Malkiat S. Johal earned his Ph.D. from the University of Cambridge in England. He later served
as a post-doctoral research associate at Los Alamos National Laboratory, New Mexico, where he
worked on the nonlinear optical properties of nanoassemblies. Dr. Johal is currently a professor and
researcher at Pomona College in Claremont, California. His work focuses on the use of self-assembly
and ionic adsorption processes to fabricate nanomaterials for optical and biochemical applications.
73109