You are on page 1of 14

Chapter 5

Nanomaterials: Characterization

5.1 Goals
• Overview of characterization techniques.

Read:

• P. S. Hale et al., J. Chem. Educ. 82 (5), 775 (2005). Growth kinetics and modeling of
ZnO nanoparticles.

5.2 X-Ray Diffraction


Waves of wavelength comparable to the crystal lattice spacing are strongly scattered (diffracted).
Analysis of the diffraction pattern allows to obtain information such as lattice parameter,
crystal structure, sample orientation, and particle size. We will only mention that lattice
parameters are obtained from the Bragg formula:

2d sin θ = nλ, (5.1)

where d is the lattice spacing.


In a typical set-up, a collimated beam of X-rays is incident on the sample. The intensity
of the diffracted X-rays is measured as a function of the diffraction angle 2θ (Fig. 5.1). The
intensities of the spots provide information about the atomic basis. The sharpness and shape
of the spots are related to the perfection of the crystal. The two basic procedures involve
either a single crystal or a powder. With single crystals, a lot of info about the structure

61
62 CHAPTER 5. NANOMATERIALS: CHARACTERIZATION

Figure 5.1: A table-top XRD set-up and results for a w-ZnS nanowire.

can be obtained. On the other hand, single crystals might not be readily available and
orientation of the crystal is not straightforward.
One disadvantage of XRD is the low intensity of diffracted beam for low-Z materials.

5.3 Optical Spectroscopy


Optical spectroscopy uses the interaction of light with matter as a function of wavelength
or energy in order to obtain information about the material. For example, absorption or
emission (photoluminescence or PL) experiments with visible and UV light tend to reveal
the electronic structure. Vibrational properties of the lattice (i.e., phonons) are usually in
the IR and studied either using IR absorption or Raman spectroscopy. Raman is an example
5.3. OPTICAL SPECTROSCOPY 63

of an inelastic process whereby the energy of the incoming light is changed. The others are
elastic processes where the intensity is changed. Typical penetration depth is of the order
of 50 nm. Optical spectroscopy is attractive for materials characterization because it is fast,
nondestructive and of high resolution.

5.3.1 UV-vis spectroscopy


This technique involves the absorption of near-UV or visible light. One measures both
intensity and wavelength. It is usually applied to molecules and inorganic ions in solution.
Broad features makes it not ideal for sample identification. However, one can determine the
analyte concentration from absorbance at one wavelength and using the Beer-Lambert law:

I
A = − log( ) = a × b × c, (5.2)
I0

where a = absorbance, b = path length, and c = concentration. A schematic of the technique


is shown in Fig. 5.2, together with a sample data.

Figure 5.2: Schematic of UV-vis spectrophotometer, an apparatus and a sample data on ZnO
nanoparticles. Absorption spectra for a colloid aged at 650 C. The spectra were recorded at
0, 1, 3, 5, 11, 16, 30, 60, 90, and 120 min after immersion in the aging water bath [9].
64 CHAPTER 5. NANOMATERIALS: CHARACTERIZATION

5.3.2 Absorption and photoluminescence


An example of PL data on semiconducting nanowires was previously given in Fig. 2.7. Ab-
sorption results are shown in Fig. 5.3.

Figure 5.3: Absorption data from bulk cubic ZnS and w-ZnS nanowire.

5.3.3 Raman spectroscopy


The Raman process describes the excitation of vibrational modes (phonons) in the sample
using light, whose frequency is then reduced due to energy conservation. Such a scattering
process is weak; experimentally, a laser is needed for good signal.
Raman spectroscopy has been used to characterize the chirality of carbon nanotubes
(Fig. 5.4).

5.4 Optical Microscope


The basic principles of image formation is illustrated in Fig. 5.5.

5.5 TEM and SEM


Electron beams can be used to produce images. The basic operation in a transmission
electron microscope (TEM) is for electrons to be generated from an electron gun, which are
then scattered by the sample, focussed using electrostatic lenses, and finally form images
(Fig. 5.6).
5.5. TEM AND SEM 65

Figure 5.4: Raman spectrum on CNT’s [10].

Figure 5.5: Principles of optical image formation.

A typical accelerating voltage is 100 kV for which the electrons have mean free paths
of the order of a few tens of nm for light elements and a few hundreds of nm for heavy
elements. These would be the ideal film thicknesses since much thinner films would lead
to little scattering and much thicker ones would lead to too many scattering of the same
electron resulting in a blurred image of low resolution.
The imaging mode can be controlled by the use of an aperture. If most of the unscattered
electron is allowed through, the resulting image is called a bright-field image. If specific
scattered beams are selected, the image is known as a dark-field image. Examples are shown
in Fig. 5.7. In addition to forming images, a TEM can be used for chemical analysis and
melting-point determination.
If the TEM is operated in scanning mode, it is known as a scanning electron microscope
or SEM. Electrons scattered from the sample are collected on a CRT to form the image. The
resolution is a few nm and magnification is from ∼ 10 to 500,000 times. One such SEM is
shown in Fig. 5.8.
66 CHAPTER 5. NANOMATERIALS: CHARACTERIZATION

Figure 5.6: TEM.

5.6 Scanning-Tunneling Microscope


The scanning-tunneling microscope (or STM) is one of the most powerful microscopes avail-
able. It provides atomic-scale resolution of surfaces and is also being developed to move
atoms on surfaces. According to its inventors, G. Binnig and H. Rohrer of IBM Zurich, it
was first operational in 1981. They won the 1986 Nobel Prize for this work.
The STM relies on a purely quantum-mechanical phenomenon: tunneling. The main
components are drawn in Fig. 5.9(a). The STM relies on the fact that electrons near surfaces
have wave functions which decay into the vacuum outside the surface boundary [Fig. 5.9(b)].
The microscope consists on a conducting tip connected to a current-measuring circuit. When
the tip is in close proximity to the surface (∼ 1 nm), the decaying wave function from the
surface could overlap with the tip, i.e., the surface electron has a finite probability of being
in the tip. Since the latter is conducting, the electron under a voltage field (∼ 1–10 V) can
5.6. SCANNING-TUNNELING MICROSCOPE 67

Figure 5.7: TEM. ZnS nanowires.


68 CHAPTER 5. NANOMATERIALS: CHARACTERIZATION

Figure 5.8: SEM.


5.6. SCANNING-TUNNELING MICROSCOPE 69

Figure 5.9: Principles of a scanning-tunneling microscope. (a) Set-up. (b) Physics.

then move creating a current. This current is known as a tunneling current whose magnitude
(as for all tunneling currents) is very sensitive to the surface–tip separation (the region in
between effectively acting as a barrier):

I ≈ V e−2κd , (5.3)

where κ is the wave number of the electrons. The current should also depend on the density
of electron states. A constant current would correspond to a constant altitude of the tip with
respect to the surface (Fig. 5.10). Hence a constant-current scan of the two-dimensional plane
reveals the surface structure. The tip motion can then be converted into a grayscale image.
The other mode of operation of an STM is in constant height mode (Fig. 5.10). Note that,
classically, the system consists of an open circuit, hence there should be no current.
The extreme sensitivity to the sample–probe separation translates into a very high pre-
cision in the vertical height (∼ 0.01 nm), with the separation being typically around 1 nm.
Indeed, the piezoelectric mechanism has extension coefficients of ∼ Å/V, resulting in very
accurate vertical movement of the tip for low voltages (typically, mV for metals and V for
semiconductors due to the latter’s band gap). The lateral resolution, on the other hand, is
determined by the tip size which can be down to a single atom and the best resolution is
∼ 0.1 nm. Prior to the STM, traditional microscopes have been based on light and electrons.
70 CHAPTER 5. NANOMATERIALS: CHARACTERIZATION

Figure 5.10: STM. Operating modes.

The resolution of the latter types of microscopes is limited by diffraction effects; basically,
the resolution ∼ wavelength λ/2. An example of an STM image is shown in Fig 5.11.

Example
λvisible ∼ 500 nm.
λe (1 keV) = hp = √2mE
h
∼ 0.1 nm.
While the electron microscope can display atomic resolution, the high energy required
means that the primary electron beam penetrates deep into the material and is not sensitive
to surface features. Both of the above types of microscopes require focussing elements. A
major new feature of the STM are the moving parts (controlled by piezoelectric materials)
and the shielding of the latter from external vibrations (using electromagnetic damping via
eddy currents).
A limitation of the STM is the requirement for a conducting sample. A variation on
the STM for insulators is the atomic-force microscope (AFM) which operates on the atomic
force between the sample atoms and the probe atoms (in contrast to the tunneling current
for the STM).

STM chemistry
The STM has also been used to carry out chemical reactions on surfaces. This relies on a
tip–sample interaction. The force can be either attractive or repulsive.
Thus Hla et al. [1] induced all the steps of the Ullmann reaction with the STM tip
5.7. ATOMIC FORCE MICROSCOPY 71

Figure 5.11: InAs surface structure.

(Fig. 5.12); this has the potential to lead towards single-molecule engineering. The reaction
is
2C6 H5 I + 2Cu → C12 H12 + 2CuI.

Steps Chemical STM


reaction reaction
dissociation of thermal electrons
iodobenzene ∼ 180 K 1.5 eV
2 phenyl diffusion tip dragging
groups
association molecular excitation
electrons

5.7 Atomic Force Microscopy


The AFM differs from the STM in that what is being measured is the force between the
sample and the tip. The AFM operates like a record player except that it has flexible
cantilevels, sharp tips, and a force feedback system (Fig. 5.13). The spring constant of the
cantilever is of the order of 0.1 N/m which is about ten times more flexible than a slinky.
72 CHAPTER 5. NANOMATERIALS: CHARACTERIZATION

Figure 5.12: STM-catalyzed chemical reaction.

Since no electric current is involved, the tip/sample does not have to be metallic. There
are two modes of operation: contact mode whereby the sample-tip distance is so small that
the important force is the core-core repulsive one, and noncontact mode where the force is
the van der Waals one. AFM’s can achieve a resolution of 10 pm.
Examples of AFM images are shown in Fig. 5.14.
5.7. ATOMIC FORCE MICROSCOPY 73

Figure 5.13: AFM. Schematic.


74 CHAPTER 5. NANOMATERIALS: CHARACTERIZATION

Figure 5.14: Patterns formed using AFM in polymers, 80 nm deep. From Nature Materials
2, 468472 (2003).

You might also like