Добавил:
Опубликованный материал нарушает ваши авторские права? Сообщите нам.
Вуз: Предмет: Файл:

Patterson, Bailey - Solid State Physics Introduction to theory

.pdf
Скачиваний:
1117
Добавлен:
08.01.2014
Размер:
7.07 Mб
Скачать

12Current Topics in Solid Condensed–Matter Physics

This chapter is concerned with some of the newer areas of solid condensed-matter physics and so contains a variety of topics in nanophysics, surfaces, interfaces, amorphous materials, and soft condensed matter.

There was a time when the living room radio stood on the floor and people gathered around in the evening and “watched” the radio. Radios have become smaller and smaller and thus, increasingly cheaper. Eventually, of course, there will be a limit in smallness of size to electronic devices. Fundamental physics places constraints on how small the device can be and still operate in a “conventional way”. Recently people have realized that a limit for one kind of device is an opportunity for another. This leads to the topic of new ways of using materials, particularly semiconductors, for new devices.

Of course, the subject of electronic technology, particularly semiconductor technology, is too vast to consider here. One main concern is the fact that quantum mechanics places basic limits on the size of devices. This arises because quantum mechanics associates a wavelength with the electrons that carry current and electrical signals. Quantum effects become important when electron wavelength becomes comparable to component size. In particular, the phenomenon of tunneling, which is often assumed to be of no importance for most ordinary microelectronic devices becomes important in this limit. We will discuss some of the basic physics needed to understand these devices, in which tunneling and related phenomena are important. Here we get into the area of bandgap engineering to attain structures that have desired properties not attainable with homostructures. Generally, these structures are nanostructures. A nanostructure is a condensed-matter structure having at least one minimum dimension between about 1 nm to 10 nm.

We will start by discussing surfaces and then consider how to form nanostructures on surfaces by molecular beam epitaxy. Nanostructures may be two dimensional (quantum wells), one dimensional (quantum wires), or “zero” dimensional (quantum dots). We will discuss all of these and also talk about heterostructures, superlattices, quantum conductance, Coulomb blockade, and single-electron devices.

Another reduced-dimensionality effect is the quantum Hall effect, which arises when electrons in a magnetic field are confined two dimensionally. As we will see, the ideas and phenomena involved are quite novel.

Carbon, carbon nanotubes, and fullerene nanotechnology may lead to entirely new kinds of devices and they are also included in this chapter, as the nanotubes are certainly nanostructures.

610 12 Current Topics in Solid Condensed–Matter Physics

Amorphous, noncrystalline disordered solids have become important and we discuss them as examples of new materials if not reduced dimensionality.

Finally, the new area of soft condensed-matter physics is touched on. This area includes liquid crystals, polymers, and other materials that may be “soft” to the touch. The unifying idea here is the ease with which the materials deform due to external forces.

12.1 Surface Reconstruction (MET, MS)

As already mentioned, the input and output of a device go through the surface, so physical understanding of surfaces is critical. Of course, the nature of the surface also affects crystal growth, chemical reactions, thermionic emission, semiconducting properties, etc.

One generally thinks of the surface of a material as being the top two or three layers. The rest can be called the bulk or substrate. The distortion near the surface can be both perpendicular (stretching or contracting) as well as parallel. Below we concentrate on that which is parallel.

If we project the bulk with its periodicity on the surface and if no reconstruction occurs we say the surface is 1 × 1. More likely the lack of bonding forces on the surface side will cause the surface atoms to find new locations of minimum energy. Then the projection of the bulk on the surface is different in symmetry from the surface. For the special case where the projection defines primitive surface vectors a and b, while the actual surface has primitive vectors aS = Na and bS = Mb then one says one has an N × M reconstruction. If there also is a rotation R of β associated with aS and bS primitive cell compared to the a, b primitive cell we write the reconstruction as

 

 

aS

 

 

 

 

 

bS

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

×

 

 

 

 

Rβ .

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

a

 

 

 

 

 

 

 

b

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

Note that the vectors a and b depend on whether the original (unreconstructed or unrelaxed) surface is (1, 1, 1) or (1, 0, 0), or in general (h, k, l). For a complete description the surface involved would also have to be included. The reciprocal lattice vectors A, B associated with the surface are defined in the usual way as

A aS = B bS = 2π ,

(12.1a)

and

 

A bS = B aS = 0 ,

(12.1b)

where the 2π now inserted in an alternative convention for reciprocal lattice vectors. One uses these to discuss two-dimensional diffraction.

Low-energy electron diffraction (LEED, see Sects. 1.2.7 and 12.2) is commonly used to examine the structure of surfaces. This is because electrons, unlike photons, have charge and thus, do not penetrate too far into materials. There are

12.2 Some Surface Characterization Techniques (MET, MS, EE) 611

theoretical techniques, including those using the pseudopotential, which are available. See Chen and Ho [12.12].

Since surfaces are so important for solid-state properties we briefly review techniques for their characterization in the next section.

12.2Some Surface Characterization Techniques (MET, MS, EE)

AFM: Atomic Force Microscopy–This instrument detects images of surfaces on an atomic scale by sensing atomic forces between the sample and a cantilevered tip (in one kind of mode, there are various modes of operation). Unlike STMs (see below), this instrument can be used for nonconductors as well as conductors.

AES: Auger Electron Spectroscopy–uses an alternative (to X-ray emission) decay scheme for an excited core hole. The core hole is often produced by the impact of energetic electrons. An electron from a higher level makes a transition to fill the hole, and another bound electron escapes with the left-over energy. The Auger process leaves two final-state holes. The energy of the escaping electron is related to the characteristic energies of the atom from which it came, and therefore chemical analysis is possible.

EDX: Energy Dispersive X-ray Spectroscopy–electrons are incident at a grazing angle and the energy of the grazing X-rays that are produced, are detected and analyzed. This technique has sensitivities comparable to Auger electron spectroscopy.

Ellipsometry–study of the reflection of plane-polarized light from the surface of materials to determine the properties of these materials by measuring the ellipticity of the reflected light.

EELS: Electron Energy Loss Spectroscopy–electrons scattered from surface atoms may lose amounts of energy dependent on surface excitations. This can be used to examine surface vibrational modes. It is also used to detect surface plasmons.

EXAFS: Extended X-ray Absorption Spectroscopy–photoelectrons caused to be emitted by X-rays are backscattered from surrounding atoms. They interfere with the emitted photoelectrons and give information about the geometry of the atoms that surround the original absorbing atom. When this technique is surface specific, as for detecting Auger electrons, it is called SEXAFS.

FIM: Field Ion Microscopy–this can be used to detect individual atoms. Ions of the surrounding imaging gas are produced by field ionization at a tip and are detected on a fluorescent screen placed at a distance, to which ions are repelled.

LEED: Low-Energy Electron Diffraction–due to their charge, electrons do not penetrate deeply into a surface. LEED is the coherent reflection or diffraction of

612 12 Current Topics in Solid Condensed–Matter Physics

electrons typically with energy less than hundreds of electron volts from the surface layers of a solid. Since it is from the surface, the diffraction is twodimensional and can be used to examine surface reconstruction.

RHEED: Reflection High-Energy Electron Diffraction–high-energy electrons can also be diffracted from the surface, provided they are at grazing incidence and so do not greatly penetrate.

SEM: Scanning Electron Microscopy–a focused electron beam is scanned across a surface. The emitted secondary electrons are used as a signal that, in a synchronous manner, is displayed on the surface of an oscilloscope. An electron spectrometer can be used to only display electrons whose energies correspond to an Auger peak, in which case the instrument is called a scanning Auger microscope (SAM).

SIMS: Secondary Ion Mass Spectrometry–a destructive but sensitive surface technique. Kiloelectron-volt ions bombard a surface and knock off or sputter ions, which are analyzed by a mass spectrometer and thus can be chemically analyzed.

TEM: Transmission Electron Microscopy–this is like SEM except that the electrons transmitted through a thin specimen are examined. Both elastically and inelastically scattered electrons can be examined, and high contrast is possible.

STM: Scanning Tunneling Microscopy–A sample (metal or semiconductor) has a sharp metal tip placed within 10 Å or less of its surface. A small voltage of order 1 V is established between the two. Since the wave functions of the atoms on the surface of the sample and the tip overlap, in equilibrium the Fermi energies of the sample and tip equalize and under the voltage difference a tunneling current of order nanoamperes will flow between the two. Since the current flow is due to tunneling, it depends exponentially on the distance from the sample to the tip. The exponential dependence makes the tunneling sensitive to sub-angstrom changes in distance, and hence it is possible to use this technique to detect and image individual atoms. The current depends on the local density of states (LDOS) at the surface of the sample and hence is used for LDOS mapping. The position of the tip is controlled by piezoelectric transducers. The apparatus is operated in either the constant-distance or constant-current mode.

UPS: Ultraviolet Photoelectron (or Photoemission) Spectroscopy–the binding energy of a core electron is measured by measuring the energy of the core electron ejected by the ultraviolet photon. For energies not too high, the energy distribution of emitted electrons is dominated by the joint density of initial and final states. An angle-resolved mode is often used since the parallel (to the surface) component of the k vector as well as the energy is conserved. This allows experimental determination of the energy of the initial occupied state for which k parallel is thus known (see Sect. 3.2.2). See also Table 10.3.

XPS: X-ray Photoelectron (or Photoemission) Spectroscopy–the binding energy of a core electron is measured by measuring the energy of the core electron ejected by the X-ray photon–also called ESCA. See also Table 10.3

12.3 Molecular Beam Epitaxy (MET, MS) 613

There are of course many other characterization techniques that we could discuss. There are many kinds of scanning probe microscopes, for example. There are many kinds of characterization techniques that are not primarily related to surface properties. Some ideas have already been discussed. Elastic and inelastic X-ray and neutron scattering come immediately to mind. Electrical conductivity and other electrical measurements can often yield much information, as can the many kinds of magnetic resonance techniques. Optical techniques can yield important information (see, e.g., Perkowitz [12.49], as well as Chap. 10 on optical properties in this book). Raman scattering spectroscopy is often important in the infrared. Spectroscopic data involves information about intensity versus frequency. In Raman scattering, the incident photon is inelastically scattered by phonons. Commercial instruments are available, as they are for FTIR (Fourier transform infrared spectroscopy), which use a Michelson interferometer to increase the signal-to-noise ratio and get the Fourier transform of the intensity versus frequency. A FFT (fast Fourier transform) algorithm is then used to get the intensity versus frequency in real time. Perkowitz also mentions photoluminescence spectroscopy, where in general after photon excitation an electron returns to its initial state. Commercial instruments are also available. This technique gives fingerprints of excited states. Considerable additional information about characterization can be found in Bullis et al [12.5]. For a general treatment see Prutton [12.52] and Marder (preface ref. 6, pp73-82).

12.3 Molecular Beam Epitaxy (MET, MS)

Molecular beam epitaxy (MBE) was developed in the 1970s and is by now a common technology for use in making low-dimensional solid-state structure. MBE is an ultrathin film vacuum technique in which several atomic and/or molecular beams collide with and stick to the substrate. Epitaxy means that at the interface of two materials, there is a common crystal orientation and registry of atoms. The substrates are heated to temperature T and mounted suitably. Each effusion cell, from which the molecular beams originate, are held at appropriate temperatures to maintain a suitable flux. The effusion cells also have shutters so that the growth of layers due to the molecular beams can be controlled (see Fig. 12.1). MBE produces high-purity layers in ultrahigh vacuum. Abrupt transitions on an atomic scale can be grown at a rate of a few (tens of) nanometers per second. See, e.g., Joyce [12.31]. Other techniques for producing layered structures include chemical vapor deposition and electrochemical deposition.

614 12 Current Topics in Solid Condensed–Matter Physics

Fig. 12.1. Schematic diagram of an ultrahigh vacuum, molecular beam growth system (adapted from Joyce BA, Rep Prog Physics 48, 1637 (1985), by permission of the Institute of Physics). Reflection high-energy electron diffraction (RHEED) is used for monitoring the growth

12.4 Heterostructures and Quantum Wells

By use of MBE or other related techniques, heterostructures and quantum wells can be formed. Heterostructures are layers of semiconductors with the same crystal structure, grown coherently, but with different bandgaps. Their properties depend heavily on their type. Two types are shown in Fig. 12.2: normal (example GaAlAs-GaAs) and broken (example GaSb-InAs). There are also other types. See Butcher et al [12.6 p. 15]. Ec is the conduction-band offset.

Two-dimensional quantum wells are formed by sandwiching a small-bandgap material between two large-bandgap materials. Energy barriers are formed that quantize the motion in one direction. These can be used to form resonant tunneling

12.5 Quantum Structures and Single-Electron Devices (EE) 615

devices (e.g. by depositing small-bandgap – large-bandgap – small-bandgap – large

– small, etc. See applications of superlattices in Sect. 12.6.1). A quantum well can show increased tunneling currents due to resonance at allowed energy levels in the well. The current versus voltage can even show a decrease with voltage for certain values of voltage. See Fig. 12.11. Diodes and transistors have been constructed with these devices.

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

EC

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

Gap

 

 

 

 

 

 

 

 

 

EC

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

Gap

 

Gap

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

EV

EV

Gap

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

Normal (type I)

 

Broken (type II)

Fig. 12.2. Normal and broken heterostructures

12.5Quantum Structures and Single-Electron Devices (EE)

Dimension is an important aspect of small electronic devices. Dimensionality can be controlled by sandwiching. If the center of the sandwich is bordered by planar materials for which the electronic states are higher (wider bandgaps), then threedimensional motion can be reduced to two, producing quantum wells. Similarly one can make linear one-dimensional “quantum wires” and nearly zerodimensional or “quantum dot” materials. That is, a quantum wire is made by laying down a line of narrow-gap semiconductors surrounded by a wide-gap one with the carriers confined in two dimensions, while a quantum dot involves only a small volume of narrow-gap material surrounded by wide-gap material and the carriers are confined in all three dimensions. With the quantum-dot structure, one may confine or exchange one electron at a time and develop single-electron transistors that would be fast, low power, and have essentially error-free signals. These three types of quantum structures are summarized in Table 12.1.

616 12 Current Topics in Solid Condensed–Matter Physics

 

Table 12.1. Summary of three types of quantum structures

 

 

Nanostructures

Comments

 

 

Quantum wells

Superlattices can be regarded as quantum well layers – alternating

 

layers of different crystals (when the wells are not too far apart)

Quantum wires

A crystal enclosed on two sides by another crystalline material, with

 

appropriate wider bandgaps

Quantum dots

A crystal enclosed on three sides by another crystalline material –

 

sometimes descriptively called a quantum box

Note: Nanostructures have a least one dimension of a size between approximately one to ten nanometers. See Sects. 12.6 and 7.4.

References: 1. Bastard [12.2]

2.Weisbuch and Vinter [12.65]

3.Mitin et al [12.47]

12.5.1Coulomb Blockade1 (EE)

The Coulomb blockade model shows how electron–electron interactions can give rise to effects that in certain circumstances are very easy to understand. It relates to the ideas of single-electron transistors, quantum dots, charge quantization leading to an energy gap in the density of states for tunneling, and is sometimes even qualitatively likened to a dripping faucet. For purposes of illustration, we consider a simple model of an artificial atom represented by the metal particle shown in Fig. 12.3.

Experimentally, the conductance (current per voltage bias) from source to drain shows large changes with gate voltage. We wish to analyze this with the Coulomb blockade model. Let C be the total capacitance between the metal particle and the rest of the system, which we will assume is approximately the capacitance between the metal particle and the gate. Let Vg be the gate voltage, relative to source, and assume the source, particle, and drain voltages are close (but sufficiently different to have the possibility of drawing current from source to drain). If there is a charge Q on the metal particle, then its electrostatic energy is

U = QVg +

Q2

.

(12.2)

2C

 

 

 

Setting ∂U/∂Q = 0, we find U has a minimum at

 

 

Q = Q0 = −CVg .

(12.3)

If N is an integer, let Q0 = −(N + η)e, where e > 0, so

 

CVg = (N +η)e .

(12.4)

1 See Kastner [12.32]. See also Kelly [12.33, pp. 300-305].

12.5 Quantum Structures and Single-Electron Devices (EE) 617

Fig. 12.3. Model of a single-electron transistor

Note that while Q0 can be any value, the actual physical situation will be determined by the integer number of electrons on the artificial atom (metal particle) that makes U the smallest. This will only be at a mathematical minimum if Vg is an integral multiple of e/C.

For −1/2 < η < 1/2, and Vg = (N + η)e/C, the minimum energy is obtained for N electrons on the metal particle. The Coulomb blockade arises because of the energy required to transfer an electron to (or from) the metal particle (you can’t transfer less than an electron). We can easily calculate this as follows. Let us consider η between zero and one half. Combining (12.2) and (12.4),

 

1

 

Q2

 

 

U =

 

Q(N +η)e +

2

.

(12.5)

 

 

C

 

 

 

 

 

 

 

 

Let the initial charge on the particle be Qi = −Ne and the final charge be Qf = −(N ± 1)e. Then for the energy difference,

 

 

 

 

 

 

U ± =U ±f

Ui ,

 

 

 

 

we find

 

 

 

 

 

 

 

 

 

 

 

 

 

 

+

 

e2

1

 

 

 

e2

1

 

U

 

=

 

 

 

η ,

U

 

=

C

 

 

+η .

 

C

2

 

2

 

 

 

 

 

 

 

 

 

 

We see that for η < 1/2 there is an energy gap for tunneling: the Coulomb blockade. For η = 1/2 the energies for the metal particle having N and N + 1 electrons are the same and the gap disappears. Since the source and the drain have approximately the same Fermi energy, one can understand this result from Fig. 12.4. Note ∆U + is the energy to add an electron and ∆U is the energy to take away an electron (or to add a hole). Thus the gap in the allowed states of the particle is e2/C. Just above η = 1/2, the number of electrons on the artificial atoms increases by 1 (to N + 1) and the process repeats as Vg is increased. It is indeed reminiscent of a dripping faucet.

618 12 Current Topics in Solid Condensed–Matter Physics

 

 

 

 

 

e2/2C =

U+

 

 

 

 

 

 

Particle

 

 

 

 

 

 

 

Particle

 

 

 

 

 

 

e2/2C =

U

 

 

 

 

 

 

 

 

 

 

Source

 

 

 

 

Drain

 

 

Source

 

 

 

Drain

 

 

 

 

 

 

η = 0

 

 

 

 

 

 

 

η = 1/2

Fig. 12.4. Schematic diagram of the Coulomb blockade at η = 0. At η = 1/2 the energy gap ∆E disappears

Conductance

Vg = e/C

Vg

 

Fig. 12.5. Periodic conductance peaks

The total voltage change from one turn on to the next turn on occurs, e.g., when η goes from 1/2 to 3/2 or

 

e

3

 

1

 

 

e

 

Vg =

 

N +

 

N +

 

 

=

 

.

 

2

2

C

 

C

 

 

 

 

A sketch of the conductance versus gate voltage in Fig. 12.5 shows periodic peaks. In order to conduct, an electron must go from source to particle, and then from particle to drain (or a hole from drain to particle, etc.).

Low temperatures are required to see this effect, as one must have

kT < e2 , 2C

so that thermal effects do not wash out the gap. This condition requires small temperatures and small capacitances, such as encountered in nanodevices. In addition the metal particle–artificial atom has discrete energy levels that may be observed in tunneling experiments by fixing Vg and varying the drain-to-source voltage. See Kasner op cit.

Соседние файлы в предмете Химия