06.03.2015 Views

Carbon Nanostructures

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Critical Reviews in Solid State and Materials Sciences, 27(3/4):227–356 (2002)<br />

<strong>Carbon</strong> <strong>Nanostructures</strong><br />

O.A. Shenderova, 1,2 V.V. Zhirnov, 1,3 and D.W. Brenner 1<br />

1<br />

North Carolina State University, Raleigh, North Carolina; 2 International Technology Center, Research Triangle<br />

Park, North Carolina; 3 Semiconductor Research Corporation, North Carolina<br />

ABSTRACT: An overview of the various carbon structures with characteristic sizes in the nanoscale region is<br />

presented, with special attention devoted to the structures and properties of ‘nanodiamond’ and carbon nanotubes.<br />

The term ‘nanodiamond’ is used broadly for a variety of diamond-based materials at the nanoscale ranging from<br />

single diamond clusters to bulk nanocrystalline films. Only selected properties of carbon nanotubes are discussed,<br />

with an aim to summarize the most recent discoveries. Current and potential applications of carbon nanostructures<br />

are critically analyzed.<br />

Table of Contents<br />

I. Introduction ............................................................................................................................. 228<br />

A. Historical Overview ........................................................................................................... 228<br />

B. <strong>Carbon</strong> Family at the Nanoscale ....................................................................................... 230<br />

II. Stability of <strong>Carbon</strong> Phases at the Nanoscale ....................................................................... 233<br />

A. Phase Diagram of <strong>Carbon</strong> at the Nanoscale ..................................................................... 236<br />

B. Theoretical Studies on the Relative Stability of Different Forms of <strong>Carbon</strong> at<br />

the Nanoscale ..................................................................................................................... 240<br />

III. Selected <strong>Carbon</strong> <strong>Nanostructures</strong>, Their Synthesis, and Properties .................................. 249<br />

A. Diamond at the Nanoscale (‘Nanodiamond’) ................................................................... 249<br />

1. Types of ‘Nanodiamond’ ........................................................................................... 249<br />

2. Ultradispersed Diamond ............................................................................................ 266<br />

a. Synthesis and Post-Synthesis Treatment............................................................. 266<br />

b. Experimental Characterization of Ultradispersed Diamond ............................... 270<br />

3. Atomistic Simulation on Diamond <strong>Nanostructures</strong>................................................... 283<br />

a. Diamond Clusters: Structural Properties............................................................. 283<br />

b. Diamond Clusters: Electronic Properties ............................................................ 290<br />

c. Diamond Nanorods .............................................................................................. 294<br />

B. <strong>Carbon</strong> Nanotubes ............................................................................................................. 298<br />

1. Synthesis and Properties ............................................................................................ 298<br />

2. Mechanical Properties ................................................................................................ 302<br />

3. Assemblies of Nanotubes and Nanodiamond............................................................ 313<br />

IV. Applications of <strong>Carbon</strong> <strong>Nanostructures</strong> ............................................................................... 319<br />

A. Diamond-Based Nanostructured Materials for Macroscopic Applications ...................... 320<br />

1. Applications of Ultradispersed Diamond .................................................................. 322<br />

2. Applications of Ultrananocrystalline Diamond Films............................................... 326<br />

3. Applications of Carbide-Derived Diamond-Structured <strong>Carbon</strong>................................ 327<br />

B. <strong>Carbon</strong> Nanotubes in Advanced Electron Sources ........................................................... 329<br />

C. <strong>Carbon</strong> Nanotubes as Nanoelectronics Components ........................................................ 338<br />

D. Medical Applications of Fullerene-Based Materials ........................................................ 343<br />

E. Atomic Modeling of <strong>Carbon</strong> <strong>Nanostructures</strong> as a Tool for Developing New<br />

Materials and Technologies ............................................................................................... 343<br />

V. Conclusions and Future Outlook........................................................................................... 347<br />

1040-8436/02/$.50<br />

© 2002 by CRC Press, Inc.<br />

227


I. INTRODUCTION<br />

Much of the discussion of nanotechnology<br />

perspectives is currently centered around carbonbased<br />

nanostructures. While the popularity of<br />

carbon nanostructures to a large extent is due to<br />

fullerenes and nanotubes, other members of the<br />

nanocarbon family are also attracting steadily<br />

increasing attention. For confirmation we refer to<br />

recent reviews 1,2 and a book 3 on nanodiamond<br />

materials. Accordingly, structures, properties, and<br />

numerous applications of nanostructured graphite,<br />

which belongs to a broad group of so called<br />

new carbon materials, has been summarized recently<br />

in a book by Inagaki. 4 In parallel, new<br />

carbon allotropes are being discovered such<br />

as, for example, carbolite, an esoteric chain-like<br />

crystalline form of carbon. 5 Clearly, carbon<br />

nanoscience, a discipline studying properties of<br />

all groups of carbon entities at the nanoscale within<br />

a unified framework, including interrelationships<br />

between the various forms of nanocarbon, conditions<br />

under which one form transforms to another,<br />

and the possibility of combining nanocarbon<br />

entities to hierarchical structures, is becoming a<br />

field onto itself. There is also a new tendency to<br />

bring together at scientific forums researchers from<br />

different carbon communities, graphite fibers,<br />

fullerenes, and diamond, which were developing<br />

before rather independently. 6 While there are excellent<br />

reviews 7–10 that discuss several classes of<br />

carbon nanostructures in one general scheme, in<br />

particular graphite-related and fullerene materials,<br />

nanodiamond has remained out of the scope<br />

of most recent discussions. This is despite a number<br />

of recent discoveries demonstrating that layered<br />

graphite nanoparticles can transform to<br />

nanodiamond and vice versa, in other words, intimate<br />

interrelations between different forms of<br />

nanocarbon exist. Moreover, a variety of interesting<br />

research on the thermodynamics and kinetics<br />

of carbon at the nanoscale have been published<br />

recently. We review this topic in Section II.<br />

In the present work we try to generalize results<br />

on the currently known nanocarbon materials.<br />

After a short historical overview on carbon<br />

nanostructures, we complete Section I by classifying<br />

the carbon family at the nanoscale. In Section<br />

III we discuss classes of nanodiamond and<br />

particularly, in more detail, ultradispersed diamond<br />

obtained by detonation synthesis, the topic<br />

of research that is popular in Russia and eastern<br />

countries and less known in the U.S. research<br />

community. Some properties of nanotubes, selected<br />

according to the authors’ interests, are also<br />

discussed in Section III. In Section IV a critical<br />

analysis of current and perspective applications<br />

of selected nanocarbon structures is provided.<br />

A. Historical Overview<br />

While the history of synthetic graphite begins<br />

in the 19th century, 11 artificial diamonds were not<br />

synthesized until the middle of the 20th century.<br />

Since then, both graphite- and diamond-related<br />

groups of carbon materials have experienced several<br />

waves of renewed interest in scientific communities<br />

when new types of materials or synthesis<br />

techniques had been discovered. Within the<br />

graphite-based group, new materials (“new carbon”<br />

4 ), such as carbon fibers, glass-like carbons,<br />

pyrolitic carbons, etc., were developed in the early<br />

1960s, and found broad industrial applications. 4<br />

The most significant relatively recent application<br />

of this class of carbon material is probably lithium<br />

ion rechargeable batteries that use nanostructured<br />

carbon anodes, which have made possible portable<br />

electronic devices. 4 Within this group of<br />

‘new carbon’ materials, texture on a nanometer<br />

scale based on preferred orientation of anistropic<br />

hexagonal layers play an important role in their<br />

properties. Some of the ‘new’ graphitic materials<br />

contain nanostructural units within a complex<br />

hierarchical structure such as, for example,<br />

carbon fibers consisting of carbon nanotubes in<br />

their cores.<br />

A new era in carbon materials began when in<br />

the mid-1980s the family of buckminsterfullerenes<br />

(“buckyballs”) were discovered 12 followed by the<br />

discovery of fullerene nanotubules (“buckytubes”) .13<br />

The discovery of these structures set in motion a<br />

new world-wide research boom that seems still to<br />

be growing. As mentioned above, fullerene<br />

nanotubules and graphite-based materials are inherently<br />

connected, and researchers who produced<br />

carbon filaments had been unknowingly growing<br />

nanotubes decades before Iijima’s publication. 13<br />

228


Some of the related topics, predictions, and discoveries<br />

of the carbon cage structures are summarized<br />

in Table 1 in chronological order.<br />

Diamond was synthesized from graphite by<br />

high-pressure/high-temperature methods in the<br />

1950s, and low-pressure chemical vapor deposition<br />

(CVD) of diamond polycrystalline films had<br />

been developed at the beginning of 1960s. The<br />

area of the CVD of diamond films experienced<br />

several shifts of scientific and funding activity,<br />

with the last peak taking place in the United States<br />

in the mid-1990s. The interest in diamond thin<br />

films has increased in the last few years as research<br />

activities related to nanotechnology have<br />

.<br />

229


grown world-wide, and new synthesis methods of<br />

nanocrystalline diamond films have been discovered.<br />

2,14<br />

Diamond powder was synthesized by shock<br />

waves in the beginning of the 1960s by Du Pont<br />

de Nemour & Co, a leader in explosive technology<br />

previously applied to other materials. Du<br />

Pont produced diamond using shock wave compression<br />

induced by solid explosive detonation of<br />

carbon materials (graphite, carbon black) mixed<br />

with metal powder (Ni, Cu, Al, Co) placed in a<br />

capsule (which was destroyed after the process).<br />

Produced polycrystalline diamond particles of<br />

micron size (1 to 60 µm) (tradename Mypolex TM )<br />

consist of nanometer-sized diamond grains (1 to<br />

50 nm). This material has been used for highprecision<br />

polishing applications for a long time.<br />

In 1999, the DuPont Corporation was acquired by<br />

Spring Holding, a Swiss holding company and<br />

parent company of Mypodiamond Inc. The production<br />

volume of the company is about 2 million<br />

carats per year (less than half a ton). 15<br />

Another approach for producing diamond<br />

powder by a more effective means with a reusable<br />

detonation capsule is the conversion of carboncontaining<br />

compounds into diamond during firing<br />

of explosives in hermetic tanks. 17 The history<br />

of the discovery of this type of nanodiamond, also<br />

known as ultradispersed diamond (UDD) or detonation<br />

diamond, is much less known and has been<br />

described in a recently published book by<br />

Vereschagin. 3 This method was initiated in Russia<br />

in the earlier 1960s soon after Du Pont’s work<br />

on shock wave synthesis and was a very active<br />

area of research in the 1980s, where it was studied<br />

independently by different groups of researchers<br />

(Table 2). Publications in this area at that time<br />

were very scarce, with some reports appearing<br />

decades after the actual discoveries had been<br />

made. 17 The first work on nanodiamond produced<br />

by detonation in the United States was published<br />

in 1987, where the method of synthesis was described.<br />

18 In 1983 the NPO “ALTAI” was founded<br />

in Russia, the first industrial company to commercialize<br />

the process of detonation diamond<br />

production in bulk quantities (tons of the product<br />

per year). 3 According to a USSR government report<br />

(1989) on UDD production, it was planned to<br />

increase UDD production by up to 250 million<br />

carats per year. 3 At the present time the production<br />

of detonation diamond by “ALTAI” is limited.<br />

Currently, there are several commercial centers<br />

in the world producing UDD, particularly<br />

in Russia (e.g., the ‘Diamond Center’ in<br />

S.Petersburg), Ukraine (e.g., “Alit”), Belorussia,<br />

Germany, Japan, and China. A center for the production<br />

of UDD is being organized in India.<br />

B. <strong>Carbon</strong> Family at the Nanoscale<br />

In principle, different approaches can be used<br />

to classify carbon nanostructures. The appropriate<br />

classification scheme depends on the field of application<br />

of the nanostructures. For example, a classification<br />

can be based on an analysis of the<br />

dimensionalities of the structures, 4,19 which in turn<br />

are connected with the dimensionality of quantum<br />

confinement and thus is related to nanoelectronic<br />

applications. The entire range of dimensionalities<br />

is represented in the nanocarbon world, beginning<br />

with zero dimension structures (fullerenes, diamond<br />

clusters) and includes one-dimensional structures<br />

(nanotubes), two-dimensional structures<br />

(graphene), and three-dimensional structures<br />

(nanocrystalline diamond, fullerite). In a different<br />

approach, the scale of characteristic sizes can be<br />

introduced as the major criterion for classification.<br />

This scheme more naturally allows the consideration<br />

of complicated hierarchical structures of carbon<br />

materials (carbon fibers, carbon polyhedral<br />

particles). A summary based on different shapes<br />

and spatial arrangements of elemental structural<br />

units of carbon caged structures also provides a<br />

very useful picture of the numerous forms of carbon<br />

structures at the nanoscale. 20 Regarding the<br />

last approach, the spatial distribution of penta- and<br />

hexa-rings within structures also can provide a<br />

basis for classification. 21<br />

In terms of a more fundamental basis for the<br />

classification of carbon nanostructures, it would<br />

be logical to develop a classification scheme based<br />

on existing carbon allotropes that is inherently<br />

connected with the nature of bonding in carbon<br />

materials. Ironically, there is no consensus on<br />

how many carbon allotropes/forms are defined at<br />

present. From time to time publications appear<br />

proposing new crystalline forms or allotropic<br />

230


modifications of carbon. Whether fullerenes or<br />

carbynes are considered as new carbon allotropes<br />

depends to a large extent on the corresponding<br />

scientific community. 5,22,24–26 Sometimes the<br />

‘fullerene community’ appears to ignore the<br />

carbynes, 24 which were discovered in the 1960s. 5<br />

However, because it can be produced only in<br />

nanoscopic quantities, it was very difficult to<br />

measure its physical properties. Similarly, the<br />

‘carbyne community’ does not classify fullerenes<br />

as an allotrope. 25 Until the 1960s when ‘new carbon’<br />

materials were synthesized, only two allotropic<br />

forms of carbon were known, graphite and<br />

diamond, including their polymorphous modifications.<br />

Until recently, ‘amorphous carbon’ had<br />

been considered as a third carbon allotrope. Pres-<br />

231


ently, however, the structure of amorphous and<br />

quasiamorphous carbons (such as carbon blacks,<br />

soot, cokes, glassy carbon, etc.) is known to approach<br />

that of graphite to various degrees. 4,5 In<br />

this context, it should be noted that within the<br />

diamond community the same term ‘amorphous<br />

carbon’ is used for diamond like carbon thin<br />

films. 27<br />

An interesting discussion of carbon allotropy<br />

and a scheme for classifying existing carbon forms<br />

is provided in Ref. 22. The classification scheme<br />

is based on the types of chemical bonds in carbon,<br />

with each valence state corresponding to a certain<br />

form of a simple substance. Elemental carbon<br />

exists in three bonding states corresponding to<br />

sp 3 , sp 2 , and sp hybridization of the atomic orbitals,<br />

and the corresponding three carbon allotropes<br />

with an integer degree of carbon bond hybridization<br />

are diamond, graphite, and carbyne. 22 All<br />

other carbon forms constitute so-called transitional<br />

forms that can be divided to two big groups.<br />

The first group comprises mixed short-range order<br />

carbon forms of more or less arranged carbon<br />

atoms of different hybridization states, for example,<br />

diamond-like carbon, vitreous carbon, soot,<br />

carbon blacks, etc., as well as numerous<br />

hypothetical structures like graphynes and<br />

‘superdiamond’. The second group includes intermediate<br />

carbon forms with a non-integer degree<br />

of carbon bond hybridization, sp n . The subgroup<br />

with 1


FIGURE 1. Ternary “phase” diagram of carbon allotropes. P/H corresponds to the ratio of pentagonal/hexagonal<br />

rings. (Reprinted from Ref. 22, Copyright (1997), with permission from Elsevier<br />

Science.)<br />

blies of the structural units, ranging from simple<br />

forms, such as multiwall nanotubes (MWNT) or<br />

carbon onions (Table 3) to more complicated<br />

carbon architectures such as carbon black,<br />

schwarzites, and agglomerates of nanodiamond<br />

particles with fractal structure. An example of a<br />

carbon structure with a complex architecture combining<br />

two structural units are recently discovered<br />

graphite polyhedral nano- and microcrystals<br />

with axial carbon structures having nanotube cores,<br />

nanotube-structured tips, and graphitic faces. 31<br />

Finally, at the upper micro/macroscopic scale there<br />

is diamond, graphite, carbolite, fullerite and recently<br />

discovered single wall nanotube (SWNT)<br />

strands of macroscopic sizes. 32 While the described<br />

scheme corresponds to the bottom-up approach of<br />

molecular synthesis, it is also necessary to add to<br />

the scheme for completeness the nanostructures<br />

obtained by top down approaches using different<br />

nanopatterning techniques such as, for example,<br />

fabrication of diamond nanorods of 40 nm diameter<br />

by plasma etching of diamond films. 33 Obviously,<br />

structural units from different families can<br />

be combined to form hybrid nanostructures.<br />

From this section, it should be obvious that<br />

one of the approaches to classification of carbon<br />

nanostructures is based on combinations of the<br />

type of hybridization of carbon bonds within the<br />

structure and characteristic size of the structure.<br />

II. STABILITY OF CARBON PHASES AT<br />

THE NANOSCALE<br />

It is well known that the most stable carbon<br />

phase on the macroscale is graphite and that diamond<br />

is metastable. The energy difference between<br />

the two phases is only 0.02 eV/atom. How-<br />

233


FIGURE 2. Classification of carbon nanostructures. The mark ‘sp n ‘ indicates intermediate carbon forms with a noninteger<br />

degree of carbon bond hybridization.<br />

ever, because of the high activation barrier for a<br />

phase transition (~0.4 eV/atom), very high temperatures<br />

and pressures and/or the use of a catalyst<br />

are required to realize the phase transformation.<br />

Several factors, however, prompted a number<br />

of researchers to reconsider the issue of carbon<br />

phase stability at the nanoscale at the end of the<br />

1980s. 72,73 It was necessary to explain the homogeneous<br />

nucleation of diamond at low pressures<br />

from the gas phase, 74 the formation of nanometersized<br />

diamond particles during detonation of explosives,<br />

18 and the observation of nanometer-sized<br />

interstellar diamond in meteorites, which was<br />

hypothesized to form from metastable carbon<br />

condensates. 58 While there can be alternative explanations<br />

for these observations, such as kinetically<br />

hindered formation of graphite from the gas<br />

phase, 75 or nucleation of diamond at the high<br />

pressure-temperature conditions that occur during<br />

detonation that correspond to the equilibrium<br />

diamond region, the idea that nanoscale diamond<br />

can be more stable than graphite was put forward.<br />

72,73 Several researchers have since addressed<br />

this question by atomistic modeling at various<br />

levels of sophistication. 76–81 Other research areas<br />

related to the stability of carbon forms at the<br />

nanoscale are simulations, performed back in the<br />

early 1980s, of very small carbon clusters (less<br />

than 20 to 30 atoms) to understand interstellar<br />

carbon as well as studies of the energetics of<br />

fullerene and nanotube formation after these species<br />

had been discovered.<br />

Very interesting transformations between carbon<br />

forms at the nanoscale had been discovered<br />

in the mid-1990s. After annealing at around 1300<br />

to 1800 K, nanodiamond particles transform to<br />

carbon onions with a transformation temperature<br />

that depends on the particle size. 60 Moreover, it<br />

234


235


was discovered that carbon-onions transform to<br />

nanocrystalline diamond under electron irradiation.<br />

61 Several groups performed atomistic simulations<br />

to understand both these phenomena. 82–84<br />

Recently, considering the nanodiamond-onion<br />

transformation, Barnard and colleagues included<br />

fullerenes in the ‘traditional’ analysis of the relative<br />

stability of diamond and graphite at the<br />

nanoscale and defined a size region of diamond<br />

stability. 81 As the system size is increased the<br />

most stable carbon form at the nanoscale changes<br />

from fullerene — to nanodiamond — to graphite.<br />

The crossover from fullerenes to closed nanotubes<br />

has also been analyzed recently. 85 In principle, a<br />

relatively large amount of accurate simulation<br />

results have been generated that create a general<br />

concept of the stability of carbon forms at the<br />

nanoscale. It is desirable, however, that the simulations<br />

be done using the same computational<br />

approach (of ab initio level) so that a quantitative<br />

comparison of energetics reported by different<br />

groups is possible.<br />

Another important area to achieving an understanding<br />

of carbon behavior at the nanoscale is<br />

a reexamination of the carbon phase diagram by<br />

introducing in addition to pressure and temperature<br />

a third parameter — cluster size. 86–88<br />

Although very important, phase diagrams<br />

and analysis of relative stabilities of different<br />

carbon forms at zero temperature are not enough<br />

for a general understanding of the complex relationship:<br />

initial carbon material – application<br />

of certain external conditions – final nanocarbon<br />

structure. At finite temperature, the stability<br />

depends on the transition probabilities among<br />

the possible configurational states of the system<br />

and is directly related to the height of the<br />

energy barrier separating the particular states. 89<br />

Recently, the potential energy surface controlling<br />

the dynamics of the graphite-diamond phase<br />

transformation has been investigated along a<br />

model reaction path using first principles and<br />

semiempirical total energy calculations on finite<br />

carbon clusters. 80,82,90 In general, the activation<br />

barrier is size dependent and increases<br />

as the size of the cluster is increased, achieving<br />

a value of the order of several tens of eV for the<br />

largest clusters (~3 nm) as found in bulk diamond.<br />

80,90<br />

Below we discuss a recent analysis of the<br />

phase diagram for nanocarbon, including the<br />

nonequilibrium phase diagram of carbon under<br />

irradiation as well as recent atomistic simulations<br />

of the stability of carbon forms at the nanoscale.<br />

A very important related topic for a general<br />

understanding of carbon behavior at the nanoscale<br />

is diamond nucleation from the gas phase, 75 which<br />

is not considered in the present review, however.<br />

A. Phase Diagram of <strong>Carbon</strong> at the<br />

Nanoscale<br />

The phase diagram of carbon has been reconsidered<br />

several times; a recent version 91 is included.<br />

It includes, for example, an indication of<br />

regions of rapid solid phase graphite to diamond<br />

conversion, fast transformation of diamond to<br />

graphite, hexagonal graphite to hexagonal diamond<br />

synthesis, shock compression of graphite to<br />

hexagonal or cubic diamond synthesis, and other<br />

phase transitions recently observed experimentally.<br />

Low pressure–high–temperature regions of<br />

the diagram have also been tentatively assigned<br />

for carbyne formation (another example including<br />

carbyne in the phase diagram is Ref. 5).<br />

Fullerenes and carbon onions were also considered<br />

in one of the schematic versions of the diagram.<br />

92<br />

Estimates for the displacement of the phase<br />

equilibrium lines for small carbon particles containing<br />

from several hundred to several tens of<br />

thousands of atoms had been made recently. 80,86,87<br />

In the expressions for the Gibbs free energy per<br />

atom of a cluster of n atoms in a given phase, the<br />

surface energy contribution is added to the bulk<br />

free energy:<br />

G i (T,P,n)=dE i n –1/3 + G i (T,P), (1)<br />

where dE i is the n-atom cluster surface energy of<br />

the i-th phase (it is assumed 70, 40, and 1 kcal/<br />

mol for diamond, graphite, and liquid carbon,<br />

respectively 80 ). Then the phase equilibrium lines<br />

for an n-atom cluster is defined by equating the<br />

Gibbs energies of the corresponding phases (Figure<br />

3). The authors report better agreement with<br />

calculations for experimental shock pressure-vol-<br />

236


FIGURE 3. Approximate phase diagram for 1000 atom carbon clusters. Shadowed<br />

region corresponds to estimated uncertainties in location of equilibrium<br />

lines derived from available experimental data. (Reprinted from Ref. 87, Copyright<br />

2001, with permission from the American Institute of Physics.)<br />

ume and temperature data than those obtained<br />

with a bulk carbon equation of state. The results<br />

also suggest that carbon particles, of the order of<br />

10 3 to 10 4 atoms, can exist in the liquid state at<br />

lower temperatures than bulk carbon.<br />

Figure 4a illustrates a three-dimensional phase<br />

diagram where particle size is an additional parameter.<br />

88 It was derived based on the published<br />

data on properties of detonation diamond. The<br />

lower horizontal plane corresponds to the phase<br />

diagram for bulk phases. The vertical axis corresponds<br />

to the particle size. The diamond phase<br />

diminishes at a particle size 1.8 nm or below (the<br />

point T 1a at the figure) that corresponds to the<br />

experimentally observed minimum particle size.<br />

At particle sizes below 3 nm the diamond phase<br />

is considered the most stable phase, so at the<br />

upper part of the state diagram only the diamond<br />

phase is present. The striped vertical plane is<br />

drawn based on the fact that spherical diamond<br />

particles with an average size 4 nm are produced<br />

from the liquid state at a temperature of 3000 K. 55<br />

Therefore the triple point has been shifted to this<br />

point accordingly. Unfortunately, the positions of<br />

the critical points for construction of the diagram<br />

along the pressure axis were not discussed in Ref.<br />

88. In addition, the position of the triple point in<br />

Figure 3 and Figure 4a is different for small particle<br />

sizes. The triple point displaces toward higher<br />

pressures in Figure 3 and toward lower pressures<br />

in Figure 4a as the particle size is decreased.<br />

While calculations to derive Figure 3 are quite<br />

accurate, they do not explain the higher stability<br />

of diamond particles over graphite at the nanometer<br />

size scale. Thus, we suggest one more variant<br />

of the 3-D phase diagram based on the results<br />

reported in Ref. 80, but, in addition, tentatively<br />

introduce a change of the slope of the diamond/<br />

237


238<br />

FIGURE 4. Three-dimensional phase diagrams for carbon. Phase diagram constructed<br />

according to published data on detonation diamond properties (a). 88 Schematic 3-D<br />

phase diagram, including fullerenes (b). 92 (Reprinted from Ref. 88 and Ref. 92 with<br />

permission.)


graphite equilibrium line as particle size is decreased<br />

(Figure 5). This change results in a higher<br />

stability of nanodiamond over nanographite at<br />

ambient conditions.<br />

According to the fact that at sizes below 1.8<br />

nm other carbon forms are abundant, such as<br />

fullerenes and onions, it was suggested to assign<br />

the corresponding region of the state diagram to<br />

fullerenes and onions as shown schematically in<br />

Figure 4b.<br />

Although the diagrams illustrated in Figures<br />

3 to 5 are rather tentative, it is a good starting<br />

point for constructing the nano-phase carbon diagrams<br />

using more accurate methods. One of the<br />

critical points is accurate surface energies of the<br />

corresponding particles as well as realism of the<br />

related structural models (mostly surface structure).<br />

Fortunately, rapidly increasing the number<br />

of the ab initio-based works addressing these issues<br />

for relatively big systems will provide a<br />

deeper understanding of the phase equilibrium of<br />

nanocarbon.<br />

As mentioned above, under the nonequilibrium<br />

conditions of intense irradiation, the phase<br />

equilibrium between graphite and diamond can<br />

be reversed so that graphite can be transformed<br />

into diamond even if no external pressure is applied.<br />

61 A detailed quantitative study of the<br />

nonequilibrium phase diagram of carbon under<br />

irradiation was presented in Ref. 93. The theoretical<br />

treatment is based on the master equation<br />

approach for incoherent phase transformation involving<br />

motion of the interface between two<br />

phases. In addition to the thermally activated exchange<br />

of atoms across the interface, under irradiation<br />

conditions atoms are ‘ballistically’ displaced<br />

from the lattice positions (with different<br />

threshold energies for different phases) to interstitial<br />

positions within the interface. Then, depending<br />

on the temperature conditions, interstitial<br />

atoms will relax toward particular bulk lattice<br />

sites. A nonequilibrium effective free energy is<br />

defined 93 that governs the phase stability under<br />

irradiation and yields quantitative predictions of<br />

FIGURE 5. Schematic 3-D phase diagram for carbon illustrating the change in<br />

the position of the triple point as a function of particle size drawn according to Ref.<br />

87. As shown, the nanodiamond phase is the most stable phase at ambient<br />

conditions.<br />

239


the interface velocity that can be directly compared<br />

to the experimental observations obtained<br />

by TEM. As a result of this work, the nonequilibrium<br />

phase diagram (irradiation intensity vs temperature)<br />

is illustrated in Figure 6. The kinetics of<br />

a reverse transition, from nanodiamond to onions<br />

during annealing, was described in Ref. 94.<br />

B. Theoretical Studies on the Relative<br />

Stability of Different Forms of <strong>Carbon</strong> at<br />

the Nanoscale<br />

There is a relatively large number of the theoretical<br />

studies devoted to the relative stability of<br />

diamond/graphite at the nanoscale. Within the<br />

earliest studies, the energy advantage of<br />

londsdaleite over graphite for very small particles<br />

elongated along the c axis was reported based on<br />

the comparison of the number of dangling bonds<br />

of graphite and londsdaleite. 76 To explain the observation<br />

of a 5 nm-diamond found in meteorites,<br />

Nuth compared the surface energies of graphite<br />

and diamond, but the uncertainty was too large to<br />

make a reasonable conclusion. Based on scaling<br />

the enthalpies of hydrocarbon molecules to larger<br />

scales, Badziag et al. 73 suggested that diamondlike<br />

clusters with sizes up to 3 nm are more stable<br />

than aromatic structures with comparable hydrogen<br />

to carbon ratios (the transitions are predicted<br />

to occur at about H/C=0.24). A similar result was<br />

obtained for relaxed hydrogenated graphitic and<br />

diamond-like structures using bond-order interatomic<br />

potential functions for which the transition<br />

in stability for diamond/graphite occures at<br />

H/C ratios of ~0.3. 95 Gamarnik 78 compared the<br />

cohesive energies of bare surface diamond clusters<br />

and 3-D graphite clusters using an empirical<br />

description of interatomic interactions. He also<br />

calculated the temperature dependence of the critical<br />

size of stable diamond clusters, including the<br />

entropy contribution to free energy: dF=T(Sg-<br />

Sd). The difference in entropy of graphite and<br />

diamond was chosen as Sg-Sd=3.37 kJ/mol at<br />

lower temperatures and 4.59 kJ/mol at 800 to<br />

1100 o C. A more extended discussion of these<br />

values can be found in Ref. 96. According to the<br />

prediction, 78 diamond nanocrystals formed, for<br />

example, at 1100 o C should be less than 5 nm in<br />

size and the maximum size of stable diamond<br />

FIGURE 6. Experimentally and theoretically (solid line) determined<br />

nonequilibrium phase diagram (irradiation intensity vs temperature) for irradiation<br />

with 1250 keV electrons. Open circles: diamond growth, black squares:<br />

graphite growth. (Reprinted from Ref. 93, Copyright 2000, with permission<br />

from American Physical Society.)<br />

240


crystals should not exceed 10 nm at room temperature.<br />

There are also several studies based on the<br />

‘classic’ thermodynamic functions modified to<br />

account for the size of the cluster. 96,97,98 Hwang et<br />

al. in their study of homogeneous diamond vs.<br />

graphite nucleation outlined a chemical potential<br />

model 98 and a charged cluster model, where the<br />

critical size for charged particles was evaluated as<br />

350 atoms. 97 The additional pressure contribution<br />

due to the curvature of the nanometer-sized particles<br />

was considered. 96,97 It was assumed that due<br />

to the additional pressure, the external pressure<br />

necessary for the transition of nanographite to<br />

nanodiamond decreases. 96 After introducing this<br />

contribution to the pressure-temperature analytical<br />

expression for the diamond/graphite equilibrium<br />

line in the bulk phase diagram, the authors<br />

obtained a size-dependent phase diagram. The<br />

reported critical cluster size at room temperature<br />

was ~6 nm. A general theory of dynamic and<br />

static nanodiamond formation from different solid<br />

carbon forms was suggested by Lin. 109 The suggested<br />

cluster transformation mechanism is based<br />

on the concept of vibrational interactions between<br />

objects with an anomalously high Debye temperature.<br />

Results from the first ab initio restricted<br />

Hartree-Fock calculations on the phase stability<br />

of a graphene sheet and cubic diamond clusters<br />

were reported in Ref. 77. Both bare and hydrogenated<br />

surfaces were considered. The cohesive<br />

energies obtained by extrapolation to bulk values<br />

of diamond and graphite were rather different<br />

from experimental values, indicating the low accuracy<br />

of the method.<br />

The general scheme for calculation of the<br />

heats of formation for graphene sheets and diamond<br />

clusters was outlined in Refs. 73, 77, 79.<br />

Below we follow the outline of Ref. 79.<br />

The dependence of the total energy of a<br />

nanocarbon system on the number of carbon atoms<br />

can be expressed as:<br />

E tot = N C E C + N db E db , (2)<br />

where N C is the number of carbon atoms and N db<br />

is the number of dangling bonds. For two-dimensional<br />

graphite sheets N C =6N 2 and N db =6N, where<br />

N is number of carbon atoms along one edge.<br />

This can be also estimated from the formula for<br />

symmetric polycyclic aromatics C 6m2 H 6m , where<br />

m is the number of rings along a single edge. 79 For<br />

a diamond octahedral with edges of n atoms, the<br />

cluster contains N C =N(4N 2 -1)/3 and N db =4N 2 . 73<br />

The parameters E db and E C are the energy per<br />

dangling bond and per carbon atom, respectively.<br />

Atomistic simulations provide a set of values of<br />

E tot for a particular system size and configuration<br />

such that the parameters E C and E db can be extracted<br />

from a least squares fit of E tot as a function<br />

of N db /N C for the graphite and carbon clusters. The<br />

cohesive energy of a cluster is defined as the<br />

difference between the total energy per atom in a<br />

cluster and the carbon free energy (e.g., it is E tot /<br />

N C ). The surface energy contribution becomes<br />

less and less important as cluster size increases,<br />

so that for an infinite cluster extrapolation of the<br />

cluster cohesive energy should be close to the<br />

bulk carbon cohesive energy, E C,, according to the<br />

expression:<br />

E tot /N C = E C + N db /N C E db . (3)<br />

With such tests, the accuracy of the method<br />

describing interatomic interactions can be evaluated<br />

for both diamond and graphite systems.<br />

Equally, instead of dangling bonds, hydrogen termination<br />

of a cluster surface can be considered<br />

within the above scheme. Such an evaluation was<br />

done by Winter and Ree for the density functional<br />

approach, using ab initio restricted Hartree Fock<br />

and semiempirical AM1 and PM3 methods 79 for<br />

both bare and hydrogenated clusters. It was concluded<br />

that the best accuracy was obtained utilizing<br />

semiempirical methods (which is not surprising<br />

because the enthalpies of formation of organic<br />

compounds are within the fitted data set).<br />

Analysis of the heat of formation in terms of<br />

bond energies provides more physical meaning to<br />

the parameters rather than the empirical expressions<br />

above. Each carbon atom in graphite forms<br />

three intralayer sp 2 bonds and experiences a weak<br />

interlayer interaction. In the diamond lattice each<br />

carbon atom forms identical sp 3 bonds. The heats<br />

of formation for sp 2 and sp 3 carbon clusters can be<br />

expressed in terms of the CC and CH bond energies<br />

and the atomic heats of formation as follows:<br />

241


0 2<br />

∆Hf<br />

( sp ) 3 2 N<br />

2<br />

sp H sp 1 2<br />

sp O<br />

= ECC<br />

+<br />

⎛<br />

ECH<br />

− ECC<br />

+ ∆Hf<br />

( H)<br />

⎞<br />

NC<br />

2 N ⎝<br />

C 2<br />

⎠<br />

O 1 disp<br />

+ ∆Hf<br />

( C)<br />

+ ECC<br />

(2.4a)<br />

2<br />

0 3<br />

∆Hf<br />

( sp )<br />

3 N<br />

3<br />

sp H sp 1 3<br />

sp O<br />

= 2ECC<br />

+<br />

⎛<br />

ECH<br />

− ECC<br />

+ ∆Hf<br />

( H)<br />

⎞<br />

NC<br />

N ⎝<br />

C 2<br />

⎠<br />

O<br />

+ ∆H<br />

( C)<br />

(2.4b)<br />

f<br />

where N H is the number of hydrogen atoms, E CC<br />

and E HC are the energies of a C-C and H-C bond,<br />

respectively, and E disp is the C-C pair energy due<br />

to the interlayer dispersion (1.66 kcal/mol 79 ).<br />

∆H<br />

0 f ( H ) =171.29 kcal/mol and ∆H 0 f ( C )=52 kcal/<br />

mol are the experimental values of the standard<br />

heat of formation of carbon and hydrogen, respectively,<br />

at room temperature. Equating the<br />

intercepts to experimental cohesive energies of<br />

graphite and diamond (left part of the equation at<br />

big N C ), bond energies can be estimated. Thus, an<br />

outcome of the scheme is an analytical expression<br />

for atomic heats of formation of hydrogenated<br />

nanodiamond and hydrogenated nanographite as<br />

functions of the number of carbon atoms. Figure<br />

7 illustrates an example of such dependences<br />

obtained by Winter and Ree using a semiempirical<br />

approach (PM3 cluster calculations). The curve<br />

crossing corresponds approximately to 33,000<br />

atoms (~7 nm particle size).<br />

Analysis of the stability of hydrogenated<br />

nanodiamond and nanographite is relatively<br />

straightforward. What is also important is that the<br />

relaxation of hydrogenated nanodiamond surfaces<br />

is in general comparable to bulk diamond, 99,100 so<br />

that bond energies of bulk diamond can be used<br />

for rough estimations.<br />

However, analysis of the optimized geometries<br />

of the bare nanodiamond surfaces showed that the<br />

picture is quite complicated. The first analysis of<br />

nonhydrogenated octahedral nanodiamond clusters<br />

FIGURE 7. Comparison of the cluster size dependence of the heat of formation<br />

∆Hf(sp3) and ∆Hf(sp 2 ) determined by the PM3 HF method. 80 The fits to the sp 3<br />

(open circles) and sp 2 (crosses) data are given by the dashed and solid lines,<br />

correspondingly. (Reprinted from Ref. 80, Copyright 1999, with permission from<br />

Elsevier Science.)<br />

242


and graphene sheets using a semiempirical approach<br />

was done by Winter and Ree. 79 If the surface<br />

orbitals are left uncapped, a finite graphene<br />

sheet distorts due to bond formation between adjacent<br />

pairs of adjacent planar dangling bonds (Figure<br />

8). For the example in Figure 8, 12 out of 18<br />

dangling bonds form 6 in-plane π bonds. In the<br />

case of nanodiamond, the number of dangling bonds<br />

for the smallest octahedral molecule C 10 is also<br />

reduced (Figure 8). Optimized structures of all<br />

other octahedral molecules with larger sizes resemble<br />

onion-like carbon with a diamond core and<br />

flattened outer layers of the octahedral cluster to<br />

form π bonds from the unbonded orbitals (Figure 8).<br />

The bonds between core and surface atoms are<br />

elongated by up to 1.7 to 3.0 Å, followed by round-<br />

FIGURE 8. Configurations of a bare graphene sheet (a) and diamond octahedral clusters (b-d) after geometry<br />

optimization. 79 In the case of the graphene sheet (a) six additional in-plane p bonds (1.31 Å long) formed from 12<br />

of the original 18 dangling bonds and the locations of the remaining six open-shell orbitals. The number of dangling<br />

bonds reduces from 16 to 8 for C10 (adamante-related) diamond cluster (b). The “buckification” – formation of the<br />

sp 2 shell over a diamond core begins with a C 35 cluster (1sp3 atom in the core) (c) and persist as a crystal size is<br />

increased (d). The atoms in the core of C 165 cluster are highlited by dashed lines. (Reprinted from Ref. 79, Copyright<br />

1998, with permission from Kluwer Academic Publishers.)<br />

243


ing of the cluster surface; this bond elongation is<br />

more pronounced for larger clusters. No cohesive<br />

energies had been reported for the reconstructed<br />

diamond clusters in Ref. 79.<br />

The preferential exfoliation of the diamond<br />

(111) surface over other low-index faces was<br />

considered by Kuznetsov et al. 82 using standard<br />

semiempirical method (MNDO) calculations of a<br />

two-layer cluster model for the (111) and (110)<br />

surfaces and physical arguments to excluding the<br />

(100) surface. It was shown that the activation<br />

barrier for (111) plane exfoliation is much lower.<br />

A comprehensive analysis of the stability of<br />

nanodiamond clusters of three specific morphologies<br />

was done by Barnard et al. using ab initio<br />

Density Functional Theory with the Generalized-<br />

Gradient Approximation. 99–101 The cohesive energies<br />

of the diamond clusters calculated by Barnard<br />

et al. are summarized in Table 4. The results on<br />

octahedral clusters are similar to those by Winter<br />

and Ree. 79 Cuboctahedral clusters, the surface area<br />

of which is comprised of 40% of (111) and 60%<br />

of (100)-oriented surfaces, exhibited transitions<br />

from sp 3 to sp 2 bonding only on the (111) planes.<br />

The (100) surfaces initially reconstructed (thereby<br />

increasing the (111) surface area), followed by a<br />

reorientation of the surface dimers to form curved<br />

graphite-like (111) cages (Figure 9). 99 The results<br />

of relaxation of the C 29 nanodiamond showed a<br />

transformation to the C@C28 endofullerene<br />

(Figure 9). The ‘buckification” of spherical<br />

nanodiamond clusters (which mostly consist of<br />

(111) and (100) surfaces) was also observed in ab<br />

initio DFT as well as Quantum Monte Carlo simulations<br />

90 that is discussed in Section III.A.3. Three<br />

cubic (100) structures that were considered contained<br />

28, 54, and 259 atoms. In the case of the 28<br />

atoms cluster, the final structure was a metastable<br />

amorphous structure and not a C 28 fullerene, as<br />

had been expected. Additional analysis did not<br />

reveal the transformation path of this structure to<br />

the fullerene. Preliminary conclusion might be<br />

that the lack of a (111) surface may influence the<br />

onset of graphitization and thus the transformation<br />

into fullerene-like structures. The two larger<br />

cubic nanodiamonds were found to have surface<br />

244


(a)<br />

(b)<br />

FIGURE 9. Cubic (a) and cuboctahedral (b) nanodiamond crystals before<br />

and after relaxation. Note surface reconstruction of the (100) surface and<br />

‘buckification’ of the (111) surfaces. As an extreme case of ‘buckification’<br />

a C28 fullerene with an endohedral C atom is formed for the C 29 cluster.<br />

(Reprinted from Ref. 99 with permission.)<br />

245


econstructions and relaxations comparable to bulk<br />

diamond (Figure 9). 99<br />

Thus, ab initio simulations demonstrate that<br />

within the size range 1 nm 99–101 up to 3 nm for<br />

spherical clusters in Ref. 90, the crystal morphology<br />

plays a very important role in cluster stability.<br />

While the surfaces of the cubic crystals exhibit<br />

structures similar to bulk diamond, the<br />

surfaces of the octahedral and cuboctahedral clusters<br />

showed transition from sp 3 to sp 2 bonding.<br />

The preferential exfoliation of the (111) surfaces<br />

begins for clusters in the subnanometer size range<br />

and promotes the cluster transition to endofullerences<br />

for small clusters (~ tens of atoms)<br />

and onion-like shells with diamond cores for larger<br />

clusters. 99<br />

In principle, a analysis similar to that outlined<br />

above for hydrogenated carbon clusters can be<br />

done for bare reconstructed surfaces if one replaces<br />

parameters for hydrogen with those for<br />

dangling bonds. This was done by Barnard et al. 81<br />

Plotting the total energy per atom as a function of<br />

the fractional ratio of dangling bonds for nonbucky<br />

clusters from Table 4, and extrapolation of<br />

N db /N tot → 0 gave a linear fit of cohesive energy<br />

for the atoms within the inner region of the diamond<br />

cluster of 7.71 eV/atom. As can be seen<br />

from Table 4, the cohesive energies of diamond<br />

clusters are below 7 eV/atom due to the highly<br />

defective surface atoms. However, due to bond<br />

shortening and the high freedom for relaxation<br />

inner atoms can gain an additional energy so that<br />

cohesive energy becomes 7.71 eV/atom. At the<br />

same time, extrapolation using this procedure for<br />

unrelaxed clusters with dangling bonds gives a<br />

cohesive energy of 7.39 eV for bulk diamond. In<br />

principle, as the crystal size is increased, the ‘inner’<br />

cohesive energy of diamond crystals should<br />

asymptotically decrease toward the bulk diamond<br />

value. With more data points for larger clusters, it<br />

might be possible that three (or more) groups of<br />

diamond clusters with specific morphologies and<br />

corresponding specific energetics (cohesive energy<br />

of the inner core) can be defined.<br />

Barnard et al. 81 also predicted the relative<br />

stability of nonhydrogenated diamond clusters,<br />

fullerenes and hydrogenated graphite within the<br />

same study. At a system size of up to 1127 atoms,<br />

fullerenes are the most stable structures (corresponding<br />

size of the cubic diamond cluster ~1.9<br />

nm). At a system size 1127 < N C < 24,389 atoms,<br />

diamond is the most stable form (‘bucky’ diamonds<br />

were excluded from the analysis). The size<br />

of a diamond cubic cluster corresponding to 24389<br />

atoms is ~5.2 nm. Finally, at a system size larger<br />

than ~24,000 atoms, graphite is the most stable<br />

phase. In principle, it would be interesting also to<br />

include in the analysis nonhydrogenated graphite<br />

clusters, which have not yet been thoroughly investigated.<br />

Results obtained in Ref. 81 are also<br />

important for the construction of the 3-D phase<br />

diagrams discussed above.<br />

Recently, a comparison of the stability of<br />

members of two other carbon families, fullerenes<br />

and closed nanotubes, was made 85 using first principles<br />

pseudopotential calculations for carbon clusters<br />

of C N (60 ≤ n ≤ 540). The analytical expressions<br />

obtained for stabilities are based on strain<br />

energy contributions due to the curvature effects<br />

as well as a contribution due to the presence of<br />

pentagons, which introduces nonplanarity of the<br />

graphitic sheet incurring incompleteness of the π<br />

bonding, as had been shown earlier by Adams and<br />

co-workers. 102 The model 85 predicts that a<br />

nanotube of ~13 Å in diameter (e.g., a (9,9) or<br />

(10,10)) is the energetically most stable form<br />

among various single-walled nanotubes and<br />

fullerenes (Figure 10), consistent with many experimental<br />

observations. The curve of stability of<br />

fullerenes in Ref. 85 is different from the previous<br />

predictions 102 for cluster sizes above ~200 atoms.<br />

There are several studies that investigated the<br />

stability of nanotubes relative to graphene. For<br />

example, graphene is the least stable structure<br />

until about 6000 atoms, where it becomes more<br />

stable than the (10,0) and (5,5) nanotubes. 108 Energy<br />

gain due to the arrangement of SWNT into<br />

arrays caused by van der Waals interactions, as<br />

well as faceting (flattening) of nanotube walls at<br />

a larger nanotube diameter, has been addressed in<br />

Refs. 380–383 (see also Table 3). Similarly, the<br />

stability of nanotubes arranged in multi-shell structures,<br />

as well as their faceting with increasing<br />

nanotube size were studied in Refs. 377–379<br />

(Table 3). The authors 107,373–376 discussed the energy<br />

characteristics of carbon onions and related<br />

faceting issues. In addition, the molecular dynamics<br />

method was used to estimate the barriers to<br />

246


FIGURE 10. Dependence of excess energy (relatively to an infinite<br />

graphene sheet) on the system size for (n,n) capped nanotubes and<br />

fullerenes (Reprinted from Ref. 85, Copyright 2002, with permission<br />

from American Physical Society.)<br />

relative shell rotation as well as the temperature<br />

dependence of the mutual arrangement of the<br />

shells. 107<br />

Regarding “small” carbon clusters, it was<br />

known long ago from mass spectra data of molecular<br />

beams 39 that the 11, 15, 19, and 23 atoms<br />

clusters are abundant, while between 28 and 36<br />

atoms the abundance of clusters is very low. Above<br />

36 atoms and up to hundreds of atoms, only evennumbered<br />

clusters are produced. An extensive<br />

study of the stability of “small” carbon clusters<br />

for a wide size range and several types of configurations<br />

were carried out by Tomanek and Schluter<br />

within a tight binding method. 103 A recent study<br />

using a density function formalism showed similar<br />

tendencies and some differences in detailed<br />

behavior. 104 In general, the current studies 104,105<br />

show three regions for the stability of small carbon<br />

clusters; below 20 atoms the most stable<br />

geometries are one-dimensional ring clusters;<br />

between 20 and 28 atoms clusters with quite different<br />

types of geometry have similar energetics;<br />

for larger clusters fullerenes should be more stable<br />

(Figure 11). Thermodynamic estimates for the<br />

efficiency of formation of spheroidal, flat clusters,<br />

and linear carbon chains depending on the<br />

charge of the reacting particles was analyzed in a<br />

recent work. 106 It was shown that charge influences<br />

both geometry and stability of flat clusters,<br />

promoting folding to curved structures as well as<br />

dissociation. The hierarchy of the stabilities of<br />

carbon forms at the nanoscale are summarized in<br />

Figure 12.<br />

A number of atomistic studies dealt with<br />

carbon onions. In particular, concentricshell<br />

fullerenes were generated from diamond<br />

nanoparticles of 1.2 nm to 1.4 nm diameters by<br />

means of molecular dynamics simulations based<br />

on approximate Kohn-Sham equations. 83 The diamond-to-concentric-shell<br />

fullerene transformation<br />

were observed at temperatures from 1400 K to<br />

2800 K and started at the surface of the diamond<br />

particle. The final structure consisted of two concentric<br />

graphitic shells with the intershell spacing<br />

distinctly below the interlayer distance of graphite<br />

with sp 3 -like cross links between the shells.<br />

Simulated irradiation accelerated the transformation<br />

and reduced the number of cross links. In a<br />

subsequent study, 84 structural transformations of<br />

carbon nanoparticles were studied by means of<br />

molecular dynamics using a density-functionalbased<br />

tight-binding method. The starting particles<br />

consisted of 64 to 275 atoms arranged on a graphitic<br />

or diamond lattice. At elevated tempera-<br />

247


FIGURE 11. The calculated binding energy of carbon clusters as a function of the<br />

number of atoms in a cluster. For the smallest sizes, the rings are the most stable<br />

structures. For the largest size, the fullerene structure is the most stable. For 20 and 24<br />

atoms, ring, bowl, and fullerene clusters have similar energies. (Reprinted from Ref.<br />

104, Copyright 2001, with permission from Kluwer Academic Publishers.)<br />

FIGURE 12. Schematic representation of the most stable carbon phase, depending on the size of<br />

the carbon structure.<br />

248


tures (1400 to 2800 K), the particles transformed<br />

into spherical or elongated closed cages, concentric<br />

shell fullerenes, carbon nanotips, and<br />

spiraloidal and irregularly shaped clusters. The<br />

type of the final cluster depended essentially on<br />

the size and the atomic order of the starting particles,<br />

and on the temperature applied. An atomic<br />

mechanism of transformation of nanodiamond to<br />

carbon onions was considered in Ref. 82. In particular,<br />

the observed the transformation of three<br />

diamond plains to two graphite planes was explained<br />

by a “zipper”-like migration mechanism,<br />

with the carbon atoms of the middle diamond<br />

layer being distributed equally between the two<br />

growing graphitic sheets.<br />

Besides pure thermodynamic considerations<br />

of the carbon phase stability outlined above kinetic<br />

considerations play an equally important<br />

role, 110 which are, however, not discussed here.<br />

In summary, it had been demonstrated by the<br />

highest level of sophistication computational approaches<br />

that nanodiamond is thermodynamically<br />

stable over graphite when the particle size is less<br />

than 5 to 10 nm, in contrast to the macroscale where<br />

diamond is metastable. At the same time, the computational<br />

methods indicate that nanodiamond stability<br />

is restricted by the smallest sizes of ~1.9 nm,<br />

below which fullerene structures are more stable.<br />

III. SELECTED CARBON<br />

NANOSTRUCTURES, THEIR SYNTHESIS<br />

AND PROPERTIES<br />

A. Diamond at the Nanoscale<br />

(“Nanodiamond”)<br />

1. Types of Nanodiamond<br />

Below is briefly summarized experimental<br />

evidence for diamond structures with critical<br />

sizes at a nanoscale as well as related methods of<br />

synthesis. The term ‘nanodiamond’ is used to<br />

identify a variety of structures that include diamond<br />

crystals present in interstellar dust and<br />

meteorites, isolated diamond particles nucleated<br />

in the gas phase or on a surface, and nanocrystalline<br />

diamond films. Methods of nanodiamond synthesis<br />

are diverse, ranging from gas phase nucleation<br />

at ambient pressure to high-pressure high-temperature<br />

graphite transformation within a shock<br />

wave. Nanodiamond materials possess different<br />

degrees of diamond purity, with a wide variety of<br />

functional groups/elements at the surface of diamond<br />

particles or within grain boundaries in<br />

nanocrystalline diamond films. The structures can<br />

be as-grown or compacted from previously synthesized<br />

diamond particles.<br />

There are two major groups of nanodiamond<br />

structures observed experimentally. The first group<br />

includes nanodiamond, which is the final product<br />

of one of the methods of synthesis (such as<br />

ultradispersed diamond, diamond obtained by<br />

transformation of carbon onions under electron<br />

beam irradiation, ultrananocrystalline diamond<br />

films). Another group includes nanodiamond nuclei<br />

that have been observed when conventional<br />

processes of micro- and macro-diamond growth<br />

had been interrupted at early stages to study a<br />

diamond nucleation process. A summary of the<br />

representative experimental observations of<br />

nanosized diamond is provided in Table 5.<br />

The information on nanodiamond observations<br />

is arranged according to the dimensionality<br />

of the diamond constituents. We discuss systems<br />

of increasing complexity beginning with the zerodimensional<br />

structures in the form of isolated<br />

particles, particles on a surface, and particles embedded<br />

in a matrix of another material. Some of<br />

the nanodiamond particles from this group had<br />

been observed during fundamental studies of diamond<br />

nucleation or transformation of carbon<br />

phases in specially designed experiments. Other<br />

types of nanodiamond such as ultradispersed diamond<br />

obtained by the shock wave process are<br />

produced in bulk quantities. The least represented<br />

group is one-dimensional diamond structures at<br />

the nanoscale, for which we discuss only three<br />

examples. Finally, three-dimensional assemblies<br />

of diamond nanocrystals grown as thin films or<br />

compacted from UDD powder to preformed bulk<br />

shapes are reviewed in more detail due to their<br />

importance in current technological applications.<br />

a. Zero-Dimensional Nano-Diamond<br />

Structures<br />

Representative observations of isolated diamond<br />

particles are summarized in Table 5a, with<br />

249


250


selected high-resolution transmission electron<br />

microscopy (HRTEM) images illustrated in Figures<br />

13 to 19. Most of the HRTEM images of<br />

single particles are obtained from particles etched<br />

out from a matrix of foreign materials or isolated<br />

from particle agglomerates, which are typical, for<br />

example, of UDD particles during storage. In<br />

general, these observations provide valuable information<br />

on diamond stability, morphology, polymorphic<br />

modifications, and lattice defects at the<br />

nanoscale, which, in turn, can be related to the<br />

nucleation mechanism and growth conditions.<br />

However, the influence of treatment conditions<br />

on ND surface states during ND extraction and<br />

purification for sample preparation is difficult to<br />

address. Examples of observations of ND unaltered<br />

by sample preparation are in situ experiments<br />

on the phase transformation during annealing<br />

60 and electron irradiation in a high-voltage<br />

electron microscope. 61<br />

i. Nanodiamond Nucleated in a Gas<br />

Phase<br />

A rather limited number of experiments have<br />

been conducted to examine homogeneous nucle-<br />

251


252<br />

FIGURE 13. HRTEM images of nonlinear multiply-twinned nano-diamond exhibiting Σ=3{111} coherent<br />

twin boundaries are present in (a-b) Murchison X, (c-d) vapor grown particles using substrate-free MPCVD<br />

flow reactor, and (e-f) detonation soot residues. (Reprinted from Ref. 117, Copyright 1996, with permission<br />

from Elsevier Science.)


FIGURE 14. High-resolution TEM image of the cluster with the diamond lattice<br />

observed among the clusters captured on the silica membrane for 60 s with a gas<br />

ratio of C 2 H 2 :O 2 =1.04 during the flame deposition of diamond. (Reprinted from<br />

Ref. 116, Copyright 2002, with permission from Elsevier Science.)<br />

FIGURE 15. HRTEM images of ultradispersed diamond particles obtained by explosive detonation<br />

synthesis. (After Ref. 183.)<br />

253


FIGURE 16. Multiple twin particles of a presolar diamond. (Courtesy of T.<br />

L. Daulton, Naval Research Laboratory.)<br />

FIGURE 17. Using particle beams, a “carbon onion,” a structure consisting of<br />

nested fullerene-like balls, can be converted into a diamond. A growing diamond<br />

is seen inside concentric graphitic layers. The diamonds can grow up to 100<br />

nanometers in diameter. (Reprinted from Ref. 61, with permission from Nature,<br />

copyright 1996. Macmillan Magazines, Inc.)<br />

254


FIGURE 18. HRTEM images of nanodiamond particles extracted from a graphite matrix where they had been formed<br />

by ion irradiation at room temperature. 122 (Courtesy T. L. Daulton, Naval Research Laboratory, after Ref. 122,<br />

Copyright 2001, with permission from Elsevier Science.)<br />

FIGURE 19. HRTEM image of a diamond crystallite (diameter ~6 nm) grown directly<br />

on Si with a random alignment. (After Ref. 140.)<br />

255


ation of diamond in the gas phase at atmospheric<br />

and subatmospheric pressures. 43,74,111,112 Theoretical<br />

arguments based on classic nucleation theory<br />

that homogeneous nucleation of diamond is possible<br />

have been presented by Fedoseev and coworkers.<br />

43 They also reported the formation of<br />

diamond particles in the gas phase using a wide<br />

variety of methods that are summarized in Ref.<br />

74. The experimental procedures included (1) an<br />

electric discharge between graphite or nickel electrodes<br />

immersed in a liquid hydrocarbon, (2) a<br />

drop of an organic liquid exposed to a focused<br />

high intensity IR laser beam, and (3) quenching<br />

acetylene pyrolysis by either injection of water or<br />

expansion of the gas in a nozzle. In the first two<br />

cases diamond/lonsdaleite (a hexagonal diamond<br />

phase) powder with average grain size 0.1 µm<br />

was obtained, and in the last method the particle<br />

size was about 20 nm. The decomposition of<br />

methane in a radio-frequency electric field is another<br />

approach used 112 in the synthesis of submicron<br />

size diamond/lonsdaleite particles.<br />

Frenklach and co-workers 74 studied nucleation<br />

and growth of diamond powder directly from the<br />

vapor phase in a substrate-free low-pressure microwave-plasma<br />

CVD reactor. The particles were<br />

collected downstream of the reaction zone on a<br />

filter within the tubular flow reactor and subjected<br />

to wet oxidation to remove nondiamond<br />

carbon. The homogeneous diamond nucleation<br />

took place when a dichloromethane- and trichloroethylene-oxygen<br />

mixture was used as a source<br />

material. The particles formed had crystalline<br />

shapes with an average particle size around 50<br />

nm. A mixture of diamond polytypes was observed<br />

in the powder.<br />

Frenklach et al. 111 also studied the effects of<br />

heteroatom addition on the nucleation of solid carbon<br />

in a low-pressure plasma reactor. The addition<br />

of diborane (B 2 H 6 ) resulted in substantial production<br />

of diamond particles, 5 to 450 nm in diameter,<br />

under the same conditions that show no diamond<br />

formation without diborane present. The observed<br />

yield of the oxidation-resistant powder produced in<br />

boron-containing mixtures reached 1.3 mg/h.<br />

Atomic resolution lattice images of CVD<br />

nanodiamond with ~5-nm particles sizes obtained<br />

the technique developed by Frenklach et al. had<br />

been thoroughly analyzed by Daulton et al. 117 (Figure<br />

13c-d). It was found that nanodiamonds in the<br />

CVD residue have an abundance of linear twins<br />

and star-twin microstructures consistent with radial<br />

(isotropic) gas phase growth conditions. Studies<br />

of diamond nucleation directly from an activated<br />

gas phase have important implications in<br />

revealing mechanisms of interstellar dust formation,<br />

which is discussed below.<br />

Another example of homogeneous diamond<br />

nucleation in the gas phase is laser-induced decomposition<br />

of C 2 H 4 at low pressures and temperatures<br />

113,114 that results in diamond powder<br />

formation with grain diameters 6 nm to 18 um.<br />

According to the authors, 113,114 high-purity homogeneously<br />

nucleated diamond nanoparticles had<br />

spherical and faceted morphologies. Homogeneous<br />

nucleation of diamond particles at low pressure<br />

by a DC plasma jet 115 has also been reported.<br />

A drastically different nucleation and growth<br />

mechanism, the so-called charged cluster model,<br />

was suggested by Hwang et al. 98 The model predicts<br />

that stabilization of diamond originates from<br />

charge rather than hydrogen. The authors suggest<br />

that charged carbon clusters of a few nanometers<br />

are generated in the CVD diamond and are suspended<br />

like colloidal particles in the gas phase<br />

and then subsequently deposited as diamond films.<br />

In an extended series of studies to confirm the<br />

model, the authors thoroughly analyzed the dependence<br />

of the nanodiamond nucleus formation<br />

on the deposition environment. In the most recent<br />

experiments they observed carbon clusters by TEM<br />

after capturing them on a grid membrane during<br />

oxyacetylene flame synthesis (Figure 14). 116 It<br />

was found that the captured clusters of ~1.5 nm in<br />

a gas mixture with an acetylene-to-oxygen ratio<br />

of 1.04 were mostly amorphous with a few having<br />

a diamond lattice. Clusters larger than 5 nm captured<br />

at a gas ratio of 1.09 were mostly graphite<br />

with a minor fraction of diamond. 116 The authors<br />

also emphasize, following the observation of several<br />

hundreds atoms clusters by HRTEM that, in<br />

principle, the crystallinity can be altered by interaction<br />

with the substrate. So far, these are the<br />

smallest sizes of observed diamond clusters. Regarding<br />

the charged cluster model, first principles<br />

atomistic simulations of charged diamond cluster<br />

stability can provide deeper insight on the reliability<br />

of the model.<br />

256


2. Ultradispersed Diamond, or Detonation<br />

Diamond<br />

A technologically important class of<br />

nanodiamond materials is ultradispersed diamond.<br />

UDD synthesis is performed by the detonation<br />

of solid explosives in an inert atmosphere.<br />

The product obtained in detonation synthesis,<br />

called detonation soot, contains the diamond phase,<br />

which is separated by chemical treatment based<br />

on moderate temperature oxidation of the impurities<br />

by nitric acid under pressure. 1 Therefore, the<br />

final morphology of the particles is influenced by<br />

the treatment conditions. Images of UDD particles<br />

are shown in Figure 13 and Figure 15. A<br />

very well-developed facet on a particle can be<br />

seen in Figure 15. Daulton et al. 117 performed<br />

extensive studies on the morphology of UDD<br />

particles that is described in the next subsection.<br />

iii. Interstellar Diamond<br />

Astronomical observations suggest that as<br />

much as 10 to 20% of the interstellar carbon is in<br />

the form of nanodiamonds. 125 Diamond nanograins<br />

are the most abundant but less understood interstellar<br />

matter found in metorites, where it comprises<br />

up to 0.15%, equivalent to about 3% of the<br />

total carbon. 126 The conclusion about the presolar<br />

nature of meteoritic nanodiamonds is based mainly<br />

on isotope composition measurements of trace<br />

elements, including noble gases. 127 There are<br />

uncertainties in the fraction of the presolar<br />

nanodiamond population in meteorites because<br />

the amount of trace elements is very small. 118 In<br />

general, questions on the places and times of the<br />

origin of nanodiamond particles remain open.<br />

The first presolar grains discovered were nanometer-sized<br />

diamonds isolated from meteorite<br />

Allende 58 (Figure 16). The diameter of presolar<br />

diamond particles in Allende ranged between 0.2<br />

to 10 nm, with an average size near 2.7 nm.<br />

The two main theories for presolar diamond<br />

formation are vapor condensation, similar to CVD<br />

processes, in the outer envelopes of carbon and<br />

red-giant stars 58 and shock-induced metamorphism<br />

in a supernovae. 121 It has been also suggested that<br />

vapor condensation of diamond could also occur<br />

in a supernovae. 119 The coexistence of the two<br />

mechanisms of ND formation (vapor condensation<br />

and shock-induced metamorphism) is assumed<br />

in the ND synthesis by explosion of a mix of<br />

nanocarbon species and explosives. 3 Another indication<br />

of the possible coexistence of the mechanisms<br />

is the presence of a bimodal distribution in<br />

the sizes of ND particles produced by shock wave<br />

compression, where particles within the 1 to 4 nm<br />

range are assumed to be condensed from the vapor<br />

phase. 3<br />

To discriminate among the most likely formation<br />

mechanisms, high-pressure shock-induced<br />

metamorphism or low-pressure vapor condensation,<br />

the microstructures of presolar diamond crystallites<br />

were compared to those of (terrestrially)<br />

synthesized nano-diamonds by Daulton et al. 117<br />

Nano-diamonds isolated from acid dissolution<br />

residues of primitive carbonaceous meteorites<br />

(Allende and Murchison) were studied using<br />

HRTEM. The synthesized diamonds used for<br />

comparison in this study were produced by explosive<br />

detonation (e.g., UDD particles, because it<br />

was assumed that their morphology should be<br />

close to that produced by shock wave transformation<br />

of carbon species) and by direct nucleation<br />

and homoepitaxial growth from the vapor phase<br />

in CVD processes (obtained from Frenklach’s<br />

group). Microstructural features were identified<br />

that appear unique to both explosive detonation<br />

synthesis (shock metamorphism) and to nucleation<br />

from the vapor phase. Diamonds produced<br />

by CVD have abundant twin forms with star-like<br />

morphologies (similar to Figure 16) indicative of<br />

isotropic formation conditions. Shock-produced<br />

diamonds have mostly planar twin forms, presence<br />

of dislocations, and other features consistent<br />

with transformation behind a planar shock front 117<br />

(Table 6). A comparison of these features to the<br />

microstructures found in presolar diamonds indicates<br />

that the predominant mechanism for presolar<br />

diamond formation is a vapor deposition process,<br />

suggesting a circumstellar condensation origin. 117<br />

It is also possible that a subpopulation of presolar<br />

diamonds were shock formed, because presolar<br />

diamonds are also linked to anomalous trapped<br />

specific Xe isotope (on average one presolar diamond<br />

out of a million contain a Xe atom). This<br />

isotope can only be produced in a supernovae<br />

where shock metamorphism conditions would<br />

prevail. 117,118<br />

257


Another theory of presolar diamond formation<br />

suggests that graphite grains irradiated by<br />

energetic particles could be transformed into diamond,<br />

for instance, around supernovae. 120 So far,<br />

there has been experimental confirmation of graphite<br />

transformation to diamond by MeV electron<br />

and ion irradiation at elevated 123 and even at<br />

ambient 122 temperatures. The observation of<br />

nanodiamonds in fine-grained uranium-rich carbonaceous<br />

materials from the Precambrian period<br />

124 also supports the theory of diamond nucleation<br />

by irradiation of carbonaceous materials by<br />

energetic particles.<br />

The comprehensive studies of nanodiamond<br />

morphology by the astrophysical community that<br />

reveal its relevance to the mechanism of interstellar<br />

diamond formation demonstrates again the<br />

importance of an interdisciplinary approach in<br />

nanoscience from unexpected, from a first sight,<br />

perspectives.<br />

iv. Direct Transformation of <strong>Carbon</strong><br />

Solids to Nanodiamond<br />

Recent experiments have shown that heavy<br />

ion or electron irradiation induces the nucleation<br />

of diamond crystallites inside concentric nested<br />

carbon fullerenes. 61,93 Other carbon materials can<br />

also be transformed to nanodiamond by using<br />

laser pulses, MeV electrons or ion beams.<br />

Nanodiamond particles had been synthesized<br />

from fine particles of carbon black exposed to<br />

intense laser irradiation. 128,129 Similarly, the transformation<br />

from carbon nanotube to carbon onion<br />

258


and then to nanodiamond as a result of laser irradiation<br />

has been reported recently in Ref. 130.<br />

High-energy electron irradiation (1.2 MeV,<br />

>10 24 e/cm 2 ; ~100 dpa) was used successfully to<br />

convert the cores of concentric-shell graphitic<br />

onions into nanometer-size diamonds at irradiation<br />

temperatures above 900 K (Figure 17 ). These<br />

experiments were performed in situ in an electron<br />

microscope, which allowed continuous observation<br />

of the formation process. A strong compression<br />

in the interior of the onion was inferred by<br />

the observed reduction in the spacing between<br />

adjacent concentric shells during irradiation.<br />

Ion beam irradiation of carbon solids also<br />

resulted in formation of nanodiamond. Irradiation<br />

with Ne+ (3 MeV, 4 × 10 19 cm –2 ; ~600 dpa) and<br />

temperatures between 700˚C and 1100˚C converted<br />

graphitic carbon soot into nanometer-size<br />

diamonds. 123 Again the diamonds were found to<br />

nucleate in the cores of graphitic onions that developed<br />

under irradiation. The increased diamond<br />

yield when compared with electron beam irradiation<br />

is explained by the higher displacement cross<br />

section, the higher energy transfer, and the higher<br />

total beam current on the specimen.<br />

ND nucleation occurs inside graphite under ion<br />

irradiation at ambient temperature when implanted<br />

by Kr + ions (350 MeV, 6 × 10 12 cm –2 ). 122 The residue<br />

of the ion-irradiated graphite was found to contain<br />

nanodiamonds with an average diameter of 7.5 nm.<br />

Nanodiamond particles extracted from graphite irradiated<br />

by ions are illustrated in Figure 18.<br />

Another example of nanodiamond formation<br />

is the irradiation of highly oriented pyrolytic graphite<br />

surfaces using a highly charged ion (HCI). 131 In<br />

contrast to the above work, the transformation of<br />

spherical carbon onions to diamond by low-temperature<br />

heat treatment at 500 o C in air without<br />

electron and ion irradiation was reported. 136<br />

HRTEM images showed that diamond particles<br />

several tens of nanometers in diameter as well as<br />

the carbon onions coexist after the heat treatment<br />

in air. From detailed HRTEM and electron energyloss<br />

spectroscopy studies, the authors 136 suggest<br />

that sp 3 sites in the onions and the presence of<br />

oxygen during the heat treatment play important<br />

roles in the transformation without irradiation.<br />

Diamond nucleation resulting from the impact<br />

of energetic species corresponding to the<br />

conditions of a bias-enhanced CVD process 132 or<br />

diamond film growth using direct ion beam bombardment<br />

133 have been also studied extensively.<br />

The biased enhanced nucleation method, in which<br />

the substrate is negatively biased and ions are<br />

extracted from the CVD plasma, 132 provides the<br />

most versatile tool for the control of the density of<br />

the nucleus. These synthesis conditions create a<br />

rather harsh environment that decreases the probability<br />

of any nucleus that is formed of surviving<br />

in a gas phase or on a surface; therefore, a new<br />

explanation of the nucleation mechanism was<br />

required. 129 Diamond nuclei 5 to 10 nm in diameter<br />

have been observed recently in amorphous<br />

carbon films grown using bias-enhanced CVD 139<br />

or exposed to an ion beam. 130 A proposed model 139<br />

for diamond nucleation by energetic species corresponding<br />

to these two conditions involves the<br />

spontaneous bulk nucleation of a diamond embryo<br />

cluster (several tens of atoms) in a dense,<br />

amorphous hydrogenated carbon matrix; stabilization<br />

of the cluster by favorable interface conditions<br />

of nucleation sites and hydrogen termination;<br />

and, as a final step, ion bombardment induced<br />

growth through a preferential displacement mechanism.<br />

134 The preferential displacement mechanism<br />

has been used also by Banhart 135 to explain the<br />

transformation of graphite to diamond by MeV<br />

electron impact. The displacement energy of sp 2<br />

bonded atoms is considerably lower than that of<br />

sp 3 bonded atoms, so that electron bombardment<br />

leaves the diamond atoms intact but displaces the<br />

graphitic atoms to sp 3 or diamond-like positions.<br />

On the basis of this mechanism, the explanation<br />

of the diamond embryo growth under continuous<br />

bombardment during bias enhanced CVD is due<br />

to preferential displacement and transformation<br />

of amorphous carbon to diamond. The described<br />

mechanism has wide implications in understanding<br />

diamond nucleation and growth in/from other<br />

carbon-containing phases when sufficient activation<br />

energy is provided by means of highly activated<br />

species/radiation.<br />

v. Diamond Nucleation on a Substrate<br />

In the development of diamond films by CVD,<br />

studies of the nucleation and early growth mechanisms<br />

on a wide variety of substrates had been<br />

259


quite intensive (see, for example, review Ref.<br />

137). This is due to the fact that the nucleation<br />

process is critical in determining the films properties,<br />

morphology, homogeneity, defect formation,<br />

and adhesion.<br />

Because diamond heterogeneous nucleation<br />

has been a topic of numerous papers, 137,138 we<br />

include one representative study here. HRTEM<br />

images of the nucleation sites responsible for<br />

epitaxial growth of diamond by CVD on silicon 140<br />

revealed 2 to 6 nm diamond clusters at steps on<br />

the Si substrate (Figure 19).<br />

b. 1D Nanodiamond Structures<br />

The least represented group is one-dimensional<br />

diamond structures at the nanoscale. Aligned<br />

diamond whiskers have so far been formed only<br />

by a ‘top down’ approach by air plasma etching of<br />

polycrystalline diamond films, particularly of asgrown<br />

diamond films and films with molybdenum<br />

deposited as an etch-resistant mask 33 (Figure<br />

20). As for the as-grown diamond films,<br />

nanowhiskers were found to form preferentially<br />

at grain boundaries of diamond crystals. Dry etching<br />

of diamond films with Mo deposits created<br />

well-aligned whiskers 60 nm in diameter that were<br />

uniformly dispersed over the entire film surface<br />

with a density of 50/µm 2 .<br />

Micron-diameter filaments formed by colloidal<br />

assemblies of UDD particles have been observed<br />

in Ref. 64 (Figure 21). After extracting<br />

and drying, the filaments were similar to glass<br />

fibers, but no measurements of mechanical properties<br />

had been performed. Koscheev et al. also<br />

succeeded in the synthesis of submicron-diameter<br />

filaments consisting of UDD particles obtained<br />

by laser ablation of pressed nanodiamond pellets<br />

64 (Figure 21). In contrast to the dense filaments<br />

in colloids every laser-ablated fiber is a<br />

network of nanoparticle chains. Studies of elemental<br />

composition, IR, and Raman spectra of<br />

filaments confirmed that they consisted of origi-<br />

FIGURE 20. Magnified SEM micrograph of diamond nanowhiskers<br />

(2.5 µm across the picture). (Reprinted from Ref. 33, Copyright<br />

2000, with permission from Elsevier Science.)<br />

260


FIGURE 21. Diamond filaments grown by self-assembly of UDD particles in a colloidal suspension (a) and filaments<br />

obtained by laser oblation of pressed UDD pellets (b). 63 (Images courtesy of A. Koscheev.)<br />

nal nanoparticles that retained a diamond structure.<br />

After extraction from the vacuum chamber,<br />

the whole assembly behaved like an aerogel. In<br />

both examples of UDD-based filaments, the filament<br />

networks were rather tangled.<br />

c. 3D Nanodiamond Structures<br />

i. Ultra-Nanocrystalline Diamond (UNCD)<br />

Films<br />

A novel diamond-based material, called<br />

ultrananocrystalline diamond, with 2 to 5 nm<br />

grains (both H-containing and H-free) has been<br />

synthesized recently at the Argonne National<br />

Laboratory by Gruen and colleagues 2,62,141–143 using<br />

a microwave plasma assisted chemical vapor<br />

deposition (MPCVD) process.<br />

UNCD thin films were synthesized using argon-rich<br />

plasmas instead of the hydrogen-rich<br />

plasmas normally used to deposit microcrystalline<br />

diamond. By adjusting the noble gas/hydrogen<br />

ratio in the gas mixture, a continuous transition<br />

from micro- to nanocrystallinity was achieved.<br />

The controlled continuous transition from the<br />

micro- to nanoscale is a unique capability of the<br />

method. 2<br />

The use of small amounts of carbon-containing<br />

source gases (C 60 , CH 4 , C 2 H 2 ) with argon leads to<br />

the formation of C 2 -dimers, which is the growth<br />

species for all UNCD thin films. The nanocrystallinity<br />

is the result of a new growth and nucleation mechanism<br />

that involves the insertion of C 2 into the<br />

π-bonds of the nonhydrogenated reconstructed (100)<br />

surface of diamond. Unattached carbon atoms then<br />

react with other C 2 molecules from the gas phase to<br />

nucleate new diamond crystallites. 2 This results in<br />

an extremely high heterogeneous nucleation rate<br />

(10 10 cm –2 s –1 , which is 10 6 times higher than from<br />

conventional CH 4 /H 2 plasmas). UNCD grown from<br />

C 2 precursors consists of ultrasmall (2 to 5 nm)<br />

grains and atomically sharp grain boundaries.<br />

More subtle control of the properties of UNCD<br />

films can be accomplished via the addition of<br />

supplementary gasses to the plasma (N 2 , H 2 , B 2 H 6 ,<br />

PH 3 ) and growth conditions (biasing, power). For<br />

instance, the addition of hydrogen leads to highly<br />

insulating films with large columnar grains. The<br />

addition of nitrogen, however, yields films that<br />

are much more electrically conductive than UNCD<br />

made with pure CH 4 /Ar plasmas. The added nitrogen<br />

leads to the formation of CN in addition to<br />

C 2 in the plasma. The presence of CN results in<br />

decreased renucleation rates during growth, which<br />

leads to larger grains and grain boundary widths.<br />

Up to 10% of the total carbon in the<br />

nanocrystalline films is located within 2 to 4 atomwide<br />

grain boundaries. Because the grain boundary<br />

carbon is π-bonded, the mechanical, electrical,<br />

and optical properties of nanocrystalline<br />

diamond are profoundly altered compared with<br />

more conventional grain structures.<br />

261


UNCD films are superior in many ways to<br />

traditional microcrystalline diamond films. They<br />

are smooth, dense, pinhole free, phase-pure and<br />

can be conformally coated on a wide variety of<br />

materials and high-aspect-ratio structures. 2<br />

Here it should be emphasized that, although<br />

the term ‘ultra-nanocrystalline diamond’ and<br />

‘ultradispersed diamond’ have the same prefix<br />

‘ultra-’, these two materials are completely different.<br />

The latter are single diamond clusters that<br />

form agglomerates with loose bonds between<br />

particles within a powder or in solution where<br />

they are stored, and the former is a bulk material<br />

with strong chemical bonds and atomically sharp<br />

grain boundaries between nanodiamond crystals.<br />

ii. Crystalline Diamond-Structured <strong>Carbon</strong><br />

Films<br />

Recently, a completely different method from<br />

what had been discussed so far for the synthesis<br />

of diamond-structured carbon in bulk quantities<br />

has been developed by Gogotsi et al. 14 The method<br />

is based on extracting silicon from silicon carbide<br />

or metal carbide in chlorine-containing gases at<br />

ambient pressure and temperatures not exceeding<br />

1000 o C. Nanocrystalline diamond with an average<br />

crystallite size of 5 nm is formed after the<br />

extraction of silicon from the carbide. Figure 22<br />

shows diamond nanocrystals surrounded by amorphous<br />

carbon regions formed near the SiC/carbon<br />

phase interface. However, if no hydrogen was<br />

added to the gas, nanocrystalline diamond slowly<br />

transformed to the graphite phase during the longterm<br />

treatment at 1000°C, and only amorphous<br />

and graphitic carbon at a distance of more than<br />

3 µm from the SiC/carbon interface was observed.<br />

Following continued heat treatment at low hydrogen<br />

content, a typical film microstructure consisting<br />

of a nanocrystalline diamond layer (Figure<br />

22b) several microns wide near the SiC/carbon<br />

interface, followed by a region of diamond<br />

nanocrystals surrounded by carbon onions and<br />

disordered carbon was observed. The third region,<br />

closest to the surface layer, consists of carbon<br />

onions as well as curved graphite sheets,<br />

some planar graphite, and porous and disordered<br />

amorphous carbon. 145<br />

The specific feature of diamond-structured<br />

carbon is multiple diamond structures including<br />

FIGURE 22. High resolution TEM micrographs showing the structure of the carbon coating within a micrometer of<br />

the SiC/carbon interface (a), where nanocrystals of diamond are surrounded by graphitic carbon, including onionlike<br />

structures. The sample was treated in Ar/3.5% Cl 2 . Typical TEM micrograph of nanocrystalline layer of diamondstructured<br />

carbon produced with hydrogen present in the gas (b): Ar / 2.77% Cl 2 / 1.04% H 2 . (Reprinted from Ref.<br />

14, with permission from Nature, copyright 1996. Macmillan Magazines, Inc.)<br />

262


cubic, hexagonal (londsdalite) structures as well<br />

as a variety of other diamond polytypes. 145 For a<br />

stable conversion of silicon carbide to the diamond<br />

phase, the presence of hydrogen in the gas<br />

mixture to saturate dangling bonds on the surface<br />

of diamond particles is key. Nanocrystalline diamond<br />

films grown up to 50 µm thick at high<br />

hydrogen content have been demonstrated. 146 Once<br />

the process is optimized, the linear reaction kinetics<br />

allows transformation to any depth, so that the<br />

entire silicon carbide sample can be converted to<br />

nanocrystalline diamond. A specific feature of<br />

the carbide-derived coating is the possibility of<br />

varying the pores size from angstrom to a few<br />

nanometers, depending on the carbide precursor<br />

type leading to the growth of nanoporous carbon<br />

or nanoporous diamond. Thus, the morphological<br />

difference between carbide-derived and CVD-produced<br />

nanodiamond is a wide range of diamond<br />

poly types and the nanoporosity of the former.<br />

Because random orientation of diamond<br />

nanocrystals is typical for carbide-derived carbon,<br />

it was assumed that the growth of nanocrystalline<br />

diamond occurred mainly from highly disordered<br />

sp 3 carbon produced by selective etching of SiC. In<br />

most cases, SiC was first converted to amorphous<br />

sp 3 carbon, and then the formation of diamond<br />

occurred within nanometers of the SiC/carbon interface.<br />

The role of hydrogen is primarily to stabilize<br />

the dangling bonds of carbon on the surface of<br />

diamond nanocrystals. Correspondingly, if the starting<br />

material is SiC powder, a powder of diamond<br />

structured carbon can be synthesized.<br />

Figure 23 shows diamond particles with an<br />

average size of 5 nm embedded in an amorphous<br />

matrix formed during chlorination of TiC. 144 It is<br />

worth noting the absolutely round particle that are<br />

formed as illustrated in Figure 23.<br />

iii. Nanocomposite Material from UDD<br />

Bonded by Pyrocarbon<br />

Another interesting bulk form of nanodiamond<br />

particles is the so-called nanodiamond composite<br />

(NDC), 146 which consists of UDD particles connected<br />

by a pyrocarbon (PyC) matrix. Nanodiamond<br />

FIGURE 23. High-resolution TEM image of diamond nanocrystals embedded in amorphous<br />

carbon in a carbide derived carbon produced by chlorination of TiC. 144 (Reprinted from Ref. 144<br />

with permission from Y. Gogotsy.)<br />

263


powder is placed in a container of a predetermined<br />

shape, which is then bonded together by pyrocarbon<br />

formed by means of methane decomposition<br />

through the entire volume of the diamond powder.<br />

146 This material is characterized by a high<br />

porosity (50 to 70%). The reported pores size is not<br />

greater than 20 to 30 nm with average radius of 4.5<br />

nm. The Young’s modulus is 30 GPa.<br />

Due to the high density of nanopores, the<br />

material posses a high sorption activity, particularly<br />

for large biomolecules (such as tripsin). 146<br />

The production of the material is realized at Skeleton<br />

Technologies, Inc.<br />

iv. ND by Shock-Wave Process<br />

Polycrystalline diamond powder can be produced<br />

by shock synthesis. 16 Under suitable conditions,<br />

explosively produced shock waves can create<br />

high pressure (~140 GPa) high-temperature<br />

conditions in confined volumes for a sufficient<br />

duration to achieve partial conversion of graphite<br />

into nanometer-sized diamond grains that compact<br />

into micron-sized, polycrystalline particles.<br />

There are two characteristic ranges within the size<br />

distribution of the primary diamond nano-crystals:<br />

1…4 nm and 10…160 nm. 3<br />

An essential component that is mixed with<br />

graphite utilized in shock wave synthesis is copper,<br />

which provides fast heat dissipation in order<br />

to avoid the transformation of the diamond back<br />

to graphite at the high temperatures that are reached<br />

during the explosion.<br />

There are several modifications of the shockwave<br />

process. Particularly, graphite (or other carbonaceous<br />

materials such as carbon black, coal,<br />

etc.) can be loaded into explosives 57 vs. loading in<br />

a fixture vessel external to the explosives.<br />

The shock-wave process was commercialized<br />

by Du Pont de Nemour & Co to produce polycrystalline<br />

diamond particles of micron size (1 to<br />

60 µm) that is more friable than monocrystalline<br />

diamond microparticles (natural or produced by<br />

HPHT) and is widely used in fine polishing applications.<br />

Recently, low dynamic pressures up to 15<br />

GPa have been reported to be enough to produce<br />

diamond from ordered pyrolytic graphite (with<br />

voids between particles) using planar shock waves<br />

parallel to the basal plane of the graphite. 147 Diamond<br />

particles consisting of crystallites with grain<br />

sizes of several tens of nanometres were observed<br />

in the upper and middle regions of the post-shock<br />

sample by HREM.<br />

An interesting set of experiments was done<br />

by applying dynamic shock wave pressures (50<br />

GPa) on samples containing carbon nanotubes<br />

and polyhedral nanoparticles. 148 HRTEM studies<br />

of the samples recovered from the soot revealed<br />

that layers of the outer shells of the<br />

nanotubes break and transform into curled graphitic<br />

structures, and the inner nanotube walls<br />

and bulk material display structural defects.<br />

Therefore, no one-dimensional nanodiamond is<br />

produced by this method. The shock-wave compression<br />

of polyhedral particles, present in the<br />

starting material, resulted in nanodiamond production<br />

(Figure 24). In this context, it should be<br />

mentioned that the book by Vereschagin 3 contains<br />

a reference to work where aggregates of<br />

needle-like diamond crystals and particles with<br />

sharp edges of about 10 nm in size were produced<br />

by a modified shock wave process. Therefore,<br />

in principle, the production of diamond<br />

nanowhiskers by explosive detonation might be<br />

possible.<br />

Interesting results on diamond production by<br />

shock-wave compression of carbyne materials<br />

have been discussed by Heimann. 149 Linear carbon<br />

allotropes were found to transform readily<br />

into diamond at comparatively weak dynamic<br />

pressures below 5 GPa (when compared with pressures<br />

above 100 GPa for graphite conversion)<br />

without external thermal activation. Because the<br />

diamond particle sizes were not reported, we do<br />

not discuss this method in more detail, but in<br />

analogy with other carbon precursors we presume<br />

that their characteristic sizes should be within the<br />

nanometer scale range. These data emphasize the<br />

importance of the carbon precursor material for<br />

diamond production.<br />

v. High-Pressure High-Temperature<br />

Process<br />

We are not aware of publications reporting<br />

the production of nanosize diamond by traditional<br />

HPHT methods, because the method was tradi-<br />

264


FIGURE 24. An HRTEM image showing nanocrystals of diamond in the post-shock<br />

sample that are the results of carbon polyhedral particles transformation under shock<br />

compression. The inset is a higher magnification of the region marked by the solid arrow.<br />

(Reprinted from Ref. 148, Copyright 1998, with permission from Elsevier Science.)<br />

tionally aimed at the growth of high-quality macroscopic<br />

diamonds and the identification of<br />

nanodiamonds would require special equipment.<br />

There are no obvious obstacles to the production<br />

of nanodiamonds by this method if the volume of<br />

the carbon precursor material was small enough<br />

and these small volumes are well separated in<br />

space to prevent growth to microscopic sizes.<br />

Another issue is the practicality of the method —<br />

there are many other more economical ways to<br />

produce nanodiamonds.<br />

As there is no special emphasis on the size of<br />

diamond produced by HPHT methods, here we<br />

only refer to recent reports on HPHT production<br />

of diamond from exotic precursor materials such<br />

as fullerenes (Refs. 150, 154, and recent review<br />

151) as well as carbon nanotubes 152 that allow<br />

much lower temperatures and externally applied<br />

pressures when compared with graphite. For example,<br />

the transformation of buckyballs to diamond<br />

at high static pressure can be done at room<br />

temperature and does not require a catalyst. 154<br />

Alternatively, the production of diamond from a<br />

metallofullerite matrix at high temperature and<br />

ambient pressure has also been reported. 153 Another<br />

group of authors reported the conversion of<br />

fullerenes to diamond under ‘moderate’ conditions<br />

5.0 to 5.5 GPa and 1400˚C. 150<br />

<strong>Carbon</strong> nanotubes have been converted to<br />

diamond at 4.5 GPa and 1300˚C using NiMnCo<br />

catalyst. 152 According to their HRTEM observations,<br />

the authors suggest that under HPHT conditions<br />

the tubular structures collapse and broken<br />

graphitic shells curl up and close into spheroidal<br />

networks to eliminate the dangling bonds at the<br />

edges. The high curvature of the formed nested<br />

graphitic shells and cross-links between the layers<br />

in the onion-like structures that are formed<br />

lead to an increased fraction of sp 3 bonds that<br />

facilitate the formation of diamond.<br />

In Ref. 155 multiwalled carbon nanotubes<br />

were heated in a diamond anvil cell by a laser<br />

above 17 GPa and 2500 K. The recovered product<br />

consisted of nano-sized octahedral crystals (diamond)<br />

of less than 50 nm. The tubular structure<br />

completely changed to granular and grain sizes<br />

corresponded to the diameter of nanotubes. The<br />

grain size of the diamond suggests that the trans-<br />

265


formation took place by direct conversion of<br />

nanotubes that might provide a control over diamond<br />

size by the choice of the MWNT size.<br />

The aim of the present section was to summarize<br />

reports on the synthesis/observation of<br />

nanodiamond. In general, it is formed under a<br />

wide variety of nonequilibrium conditions. The<br />

methods of diamond synthesis reported in this<br />

section are summarized in a tentative scheme<br />

(Figure 25).<br />

2. Ultradispersed Diamond<br />

As has already been mentioned above,<br />

ultradispersed diamond — a material consisting<br />

of isolated diamond particles — is not well known<br />

in the U.S., while it has been on the market and<br />

used widely in different technologies in the Former<br />

Soviet Union for more than a decade. The areas of<br />

application of UDD are completely different from<br />

those, for example, for UNCD films, the material<br />

being actively developed now in the U.S. While<br />

UNCD as grown is a very hard, very smooth, and<br />

chemically inert coating, a single particle of UDD<br />

consists of a chemically inert diamond core and<br />

chemically active surface, with controllable properties.<br />

In the macro-world they exist in the form<br />

of powders (where UDD form agglomerates) and<br />

suspensions (aggregated or isolated particles).<br />

Below we provide more details on the synthesis<br />

and properties of UDD.<br />

a. Synthesis and Post-Synthesis Treatment<br />

At the beginning we would like to emphasize<br />

the differences between the major methods for<br />

diamond synthesis using explosives, because the<br />

materials that are produced are very different, and<br />

it is necessary to be aware of this difference when<br />

making a choice of the material for particular<br />

applications. There are three major methods that<br />

had been commercialized. 3 The first is the transformation<br />

of carbon precursors (e.g., graphite,<br />

coal, and carbon black) to diamond in a capsule<br />

compressed by a shockwave (~ 140 GPa) generated<br />

outside the capsule (the ‘Du Pont method’).<br />

FIGURE 25. Tentative scheme summarizing the methods of synthesis of diamond nanostructures.<br />

266


To prevent diamond regraphitization, a mixture<br />

of graphite (6 to 10%) with metallic powder (Cu,<br />

Al, Ni) is used. The diamond yield is about 60<br />

mass percent of the carbon phase or about 5% of<br />

the initial mixed material loaded into a capsule.<br />

The synthesized nanodiamond had a bimodal size<br />

distribution; the first maximum was within the<br />

1 to 4 nm size range and the second was within<br />

the 10 to 150 nm range. The product of the synthesis<br />

always contained the londsdalite phase due<br />

to the martensitic type of transformation graphiterombohedral<br />

graphite-londsdelite–diamond. 3<br />

There are several modifications of this method, 3<br />

however.<br />

The second method of ND production is based<br />

on the detonation of a mixture of carbon-containing<br />

material with explosives. In this case diamond<br />

formation takes place both within the carboncontaining<br />

particles as well by condensation of<br />

carbon atoms contained in the explosives. The<br />

explosion can be done in air or in an inert atmosphere<br />

relative to the product of synthesis. 3 In the<br />

latter case, a diamond cubic phase not more than<br />

20 nm in size is formed. When the formation is in<br />

an atmosphere of air, the samples always contain<br />

londsdalite and the particle size is smaller —<br />

about 8 nm. The diamond yield constitutes up to<br />

17% of the mass of the initial carbon material or<br />

about 3.4% of the mass of the explosives.<br />

The characteristic feature of the diamond<br />

obtained by these two methods based on the shock<br />

wave compression of the initial graphite phase is<br />

formation of polycrystalline material with particle<br />

sizes comparable to the particle sizes of the<br />

precursor carbon material. The particles consist<br />

of nanocrystalline diamond grains. After purification<br />

the particles produced by detonation are classified<br />

by fractions (from 1 to 60 µm ) for use in<br />

polishing applications. 3<br />

In the third method of using explosion energy<br />

for diamond production, diamond clusters are<br />

formed from carbon atoms contained within explosive<br />

molecules themselves, so only the explosive<br />

material is used as a precursor material (Figure<br />

26). A wide variety of explosive materials can<br />

be used, including products for military applications<br />

(so that UDD synthesis is a good way of<br />

utilizing of the stock pile materials). A typical<br />

explosive is a mixture of TNT (2-methyl-1,3,5-<br />

trinitrobenzene) and hexogen (in proportion 60/<br />

FIGURE 26. Schematic illustration of the steps in the controlled detonation synthesis of nanodiamond from carboncontaining<br />

explosives (1). During explosion (2) the highly dispersed carbon medium condenses from free explosive<br />

carbon in a fraction of a microsecond. Depending on the detonation conditions, the resulting detonation soot (3)<br />

contains 40 to 80 wt.% of ultradispersed diamond. After disposal, the product has to undergo several stages of<br />

purification. Big industrial detonation reactors are able to deliver tons of detonation diamond per month. The<br />

composition of detonation soot is provided based on 163 (pictures of the detonation process courtesy of PlasmaChem<br />

GmbH, Mainz/Germany).<br />

267


40) composed of C, N, O, and H with a negative<br />

oxygen balance (i.e., with the oxygen content<br />

lower than the stoichiometric value), so that ‘excess’<br />

carbon is present in the system. A negative<br />

oxygen balance in the system is an important<br />

condition for UDD formation. The explosion takes<br />

place in a nonoxidizing medium (Figure 26) of<br />

either gas (N 2 , CO 2 , Ar, or other medium — that<br />

can be under pressure) or water (ice) — ‘dry’ or<br />

‘wet’ synthesis, correspondingly, that plays the<br />

role of a coolant. To prevent the UDD formed in<br />

the detonation wave from transforming into graphite<br />

at the high temperature generated by the detonation,<br />

the cooling rate of the reaction products<br />

should be no less than 3000 K/min. 3 The initial<br />

shock from a detonator compresses the high-explosive<br />

material, heating it and causing chemical<br />

decomposition that releases enormous amounts<br />

of energy in a fraction of a microsecond (Figure<br />

26). As the detonation wave propagates through<br />

the material it generates high temperatures (3000<br />

to 4000K) and high pressures (20 to 30 GPa) that<br />

correspond to conditions of thermodynamic stability<br />

for diamond. During detonation, the free<br />

carbon coagulates into small clusters, which grow<br />

larger by diffusion. 80,86,87 The product of detonation<br />

synthesis, called detonation soot or diamond<br />

blend, contains 40 to 80 wt.% of the diamond<br />

phase depending on the detonation conditions.<br />

The carbon yield is 4 to 10% of the explosive<br />

weight for the most effective of the three industrial<br />

methods of diamond production using explosives.<br />

In summary, there are two major technical<br />

requirements for UDD synthesis using explosives.<br />

First, the composition of the explosives must provide<br />

the thermodynamic conditions for diamond<br />

formation, and second the composition of the gas<br />

atmosphere must provide the necessary quenching<br />

rate (by appropriate thermal capacity) to prevent<br />

diamond oxidation. The diamond yield depends<br />

to a large extent on the explosive mixture<br />

(Figure 27). 86 The shape of the explosive also<br />

influences the yield; the ideal shape is spherical,<br />

but for convenience a cylindrical shape is used<br />

regularly. The relationship between the mass of<br />

the explosives and the mass of the surrounding<br />

media influences also yield (e.g., 5 kg of explo-<br />

FIGURE 27. Diamond weight fraction recovered from soot as a function of the<br />

composition of the explosive mixture. The abbreviation for explosives is as<br />

follow: TNT — 2-methyl-1,3,5-trinitrobenzene; HMX — octahydro-1,3,5,7-<br />

tetranitro-1,3,5,7-tetrazocine; TATB — 2,4,6-trinitro-1,3,5-benzenetriamine;<br />

RDX — hexahydro-1,3,5-trinitro-1,3,5-triazine. (Reprinted from Ref. 86, Copyright<br />

2000, with permission from American Institute of Physics.)<br />

268


sive requires ~11 m 3 of detonation camera with<br />

gas media at ambient pressure to provide the necessary<br />

quenching rate). 3 The mechanisms of diamond<br />

formation as well as factors influencing<br />

yield during explosive detonation have been discussed<br />

in numerous publications by different<br />

Russian research groups (summarized in Ref. 3).<br />

In the U.S., fundamental studies of carbon particle<br />

phase transformation in detonation waves<br />

have been performed at Laurence Livermore<br />

National Laboratory, primarily by Francis Ree<br />

and colleagues 80,86,87 and Los Alamos National<br />

Laboratory primarily by Sam Shaw and colleagues.<br />

158 Both the thermodynamics of chemically<br />

reactive mixtures and the kinetics of carbon<br />

coagulation have been modeled. 86,87,158 The kinetics<br />

of the formation and growth of finite carbon<br />

particles contribute significantly to the detonation<br />

properties of carbon-rich CHNO explosives. As<br />

detonation proceeds, the most important products<br />

behind the reaction zone are carbon dioxide, water,<br />

nitrogen, and carbon residues that undergo<br />

phase changes. The relatively slow growth rate of<br />

carbon particles gives rise to an extended reaction<br />

zone. As a consequence, the detonation is more<br />

sensitive to the system configuration, with the<br />

observed Chapman-Jouguet pressure reflecting the<br />

condition of a partially reacted state rather than<br />

that of the final detonation product. <strong>Carbon</strong> clustering<br />

is a slow reaction in the detonation regime<br />

as demonstrated by Shaw and Johnson at the end<br />

of 1980s 158 using a diffusion-limited model. <strong>Carbon</strong><br />

particles build up from random collisions,<br />

while the hot dense background fluid maintains<br />

the equilibrium temperature, allowing the clusters<br />

to anneal to compact spherical objects. The final<br />

particle size was estimated to be 10 4 to 10 5 atoms<br />

(50 Å), which explained the experimental particle<br />

sizes. The model did not distinguish between<br />

the various forms of carbon present during the<br />

detonation process. The model has been reconsidered<br />

taking into account the diamond/graphite<br />

transitions within the carbon particles. 80,86,87 Estimates<br />

of the displacement of the phase equilibrium<br />

lines for small carbon particles containing<br />

from several hundred to several tens of thousands<br />

of atoms 87 was also included in the analysis of the<br />

kinetics of carbon coagulation. A simplified hydrodynamics<br />

model yielded a time and pressuretemperature<br />

path-dependent value for the<br />

nonequilibrium diamond fraction of the soot mixture.<br />

86<br />

i. Purification<br />

The diamond blend in addition to UDD contains<br />

graphite-like structures (35 to 45 wt.%), and<br />

incombustible impurities (metals and their oxides<br />

— 1 to 5wt.%). 1 Using X-ray diffraction and<br />

small angle X-ray scattering, it was shown that an<br />

UDD cluster in detonation soot has a complex<br />

structure consisting of a diamond core of about<br />

4.3 nm in size and a shell made up of sp 2 coordinated<br />

carbon atoms (Figure 27). 164 The shell structure<br />

and thickness vary with the cooling kinetics<br />

of the detonation products (dry vs. wet synthesis)<br />

and conditions of chemical purification of theUDD<br />

clusters from the detonation soot. 164 According to<br />

the X-ray diffraction spectra and the results of<br />

electron microscopy, UDD of the highest degree<br />

of purification contain no intermediate amorphous<br />

phase, graphite, or londsdalite. 161,162<br />

UDD purification is performed by mechanical<br />

and chemical methods. After mechanical removal<br />

of process admixtures, the diamond-carbonic<br />

powder is subjected to thermal oxidation<br />

with nitric acid under pressure to separate the<br />

diamond phase. 1 The method of acid purification<br />

at elevated temperatures is the most efficient purification<br />

method at the present time because it<br />

comprehensively influences all admixtures: metals<br />

are dissolved and non-diamond carbon is oxidized<br />

simultaneously. The diamond should be<br />

flushed with water after separation from the acidic<br />

media. After a typical purification step, powders<br />

of UDD can be considered as a composite consisting<br />

of different forms of carbon (80 to 89%),<br />

nitrogen (2 to 3%), hydrogen (0.5 to 1.5%), oxygen<br />

(up to 10%), and an incombustible residue<br />

(0.5 to 8%). 1 The carbon consists of a mix of<br />

diamond (90 to 97%) and non-diamond carbon<br />

(3 to 10%).<br />

In detonation synthesis of nanodiamonds, the<br />

impurity content is higher when compared with<br />

other artificial diamonds (i.e., HPHT diamonds<br />

contain no less than 96% carbon). Therefore, the<br />

effect of impurities is more pronounced for UDD<br />

compared with other diamonds. 1 Impurities in<br />

269


UDD, in principle, can be divided into the following<br />

groups: 1 (1) water-soluble ionized species (free<br />

electrolytes); (2) chemically bonded to diamond<br />

surfaces and prone to hydrolysis and ionization<br />

(salt forms of functional surface groups); (3) water<br />

insoluble; and (4) incorporated into the diamond<br />

lattice and encapsulated. The impurities of the<br />

first and second groups are formed in the chemical<br />

purification of UDD by acid stage. 1 The major<br />

water-soluble admixtures (first group) are removed<br />

by washing the UDD with water. The surface<br />

functional groups (second group) can be efficiently<br />

removed by ion-exchange resins that involve demineralization<br />

of surface groups. The water-insoluble<br />

impurities represent both individual<br />

microparticles of metals, oxides, carbides, salts,<br />

and metal oxides that do not dissociate. For their<br />

removal, UDD particles are treated with acids. By<br />

using different methods of UDD purification, one<br />

can remove 40 to 95% of the impurities of these<br />

three groups. It is practically not possible to remove<br />

the impurities of the fourth group by chemical<br />

methods. The UDD of the highest purity available<br />

at ‘Diamond Center’, Inc. contains 98.5% of<br />

diamond.<br />

Commercial products of purification have<br />

the following grades: a water suspension of diamond<br />

and powder obtained from suspensions by<br />

drying and grinding of UDD are a gray powder<br />

containing up to 99.5 wt.% of a pure diamond<br />

(without counting adsorbed gases). 1 To remove<br />

non-carbon impurities, the chemically purified<br />

product is subjected in some cases to additional<br />

purification using ion-exchange and membrane<br />

technologies.<br />

In general, the UDD production consists of<br />

detonation synthesis, chemical purification and<br />

acid washing of UDD, product conditioning, and<br />

modifying of diamond.<br />

b. Experimental Characterization of<br />

Ultradispersed Diamond<br />

The properties of UDD particles are mainly<br />

defined by their nanometer-scale sizes (4 to 6 nm<br />

in diameter) that are within a transitional size<br />

range between macromolucules and crystalline<br />

solids. About half of all atoms in such particles<br />

are on the surface, and thus are unavoidably bound<br />

to adsorbed atoms, molecules, and functional<br />

groups. These adsorbed atoms, which can exceed<br />

the number of atoms in the diamond particle, can<br />

strongly affect the physical and chemical properties<br />

of the particles.<br />

To minimize surface energy, the primary UDD<br />

particles with diameters of ~4 nm form larger<br />

clusters 20 to 30 nm in size that, in turn, form<br />

larger weakly bound aggregates (of an order of<br />

magnitude of hundreds of nanometers). The UDD<br />

aggregates have a fractal nature. 1 This hierarchy<br />

of nanodiamond blocks needs to be taken into<br />

account in the interpretation of all experimental<br />

results.<br />

The smallest amount of diamond matter, which<br />

can be prepared and studied in isolation are primary<br />

particles of UDD. In addition, the particles<br />

of UDD are thermodynamically stable at ambient<br />

conditions, due to their very small size (


While by traditional dynamic synthesis from<br />

graphite diamond particles of both cubic and hexagonal<br />

(a=0.252 nm) structures are produced, UDD<br />

obtained from explosives has only the cubic structure<br />

without any traces of the hexagonal phase. 1<br />

Raman Spectroscopy<br />

All experimental Raman studies of ultradisperse<br />

diamond quote practically the same frequency position<br />

of the peak within 1321–1322 cm –1 . 167 Typical<br />

Raman spectra of UDD and bulk diamond are<br />

shown in Figure 30. 167 The observed spectrum can<br />

be explained by the phonon confinement model<br />

(see references in Ref. 167). The authors of Ref.<br />

167 calculated from this spectrum based on the<br />

phonon confinement model that the size of the<br />

diamond nanoparticles is 5.5 nm (the X-ray size<br />

obtained by the same authors was 4.3 nm).<br />

Aleksenskii et al., 165 using the same Raman<br />

technique, obtained two numbers for the size of<br />

the diamond crystallites using two different sets<br />

of constants in the phonon confinement model:<br />

3.6 and 4.3 nm. The authors 165 noted that the<br />

estimates of crystalline size based on the photon<br />

confinement model are very sensitive to the choice<br />

of the dispersion relation for the photon modes. 165<br />

Annealing Effects and Diamond-Graphite<br />

Phase Transition<br />

The narrow size range of UDD nanoparticles<br />

of 4 to 5 nm was quoted in a number of studies. 165<br />

The usual explanation is that nanodiamond is more<br />

stable than graphite (see Section II). However, the<br />

thermodynamic stability reasoning, while explaining<br />

the upper bound does not give the lower bound<br />

for the size, thus the narrow size distribution still<br />

needs to be clarified.<br />

Aleksenskii et al. 165 used Raman spectroscopy<br />

to study the annealing effects on the<br />

nanodiamond structure. Starting from an anneal-<br />

271


FIGURE 28. Model structures of detonation-produced carbon. Depending on purification<br />

conditions, different layers of the UDD shell can be ‘stripped-off’. (Reprinted from Ref. 164,<br />

with permission from American Institute of Physics.)<br />

ing temperature of T=720 K, remarkable changes<br />

in the Raman spectra were observed. The peak<br />

due to diamond nanoparticles decreased significantly<br />

in intensity while not exhibiting any shift<br />

in frequency. With further increases of T ann<br />

a monotonic decrease of intensity of the<br />

nanocrystalline diamond line at 1322 cm –1 was<br />

observed, without any change in position. At<br />

T ann =1200 K this line was no longer observed.<br />

The constancy of the position of the 1322 cm –1<br />

Raman line up to T ann =1000 K indicates that no<br />

changes occur in the structure of the diamond<br />

nuclei at such annealing temperatures. X-ray diffraction<br />

data indicated the formation of a graphitic<br />

phase for T>1200 K. 165 The authors 165<br />

pointed out that because UDD exists in aggregated<br />

form, the amorphous phase is apparently<br />

distributed inside the aggregates on the surface of<br />

diamond nuclei. The presence of an amorphous<br />

phase was supported by X-ray diffraction, which<br />

yielded about 1.5 nm for the size of amorphous<br />

particles.<br />

The diamond-graphite phase transition temperature<br />

observed in UDD (T pt >1200 K) is considerably<br />

lower than the phase transition temperature<br />

of bulk single-crystal diamond (T pt >1900 K).<br />

An interesting result was reported by<br />

Obraztsova et al. 169 The authors developed a special<br />

annealing procedure that provided an effective<br />

means of reducing the size of diamond<br />

nanoparticles while preserving the diamond crystal<br />

structure. The procedure allowed them to reduce<br />

the size of diamond particles step by step<br />

from the initial 4.5 nm down to 2 nm by changing<br />

the annealing temperature. Obraztsova et al. found<br />

that for the smaller size the complete particle<br />

transformation into the onion-like carbon structure<br />

takes place. The size and phase transformations<br />

were characterized by Raman spectroscopy.<br />

The authors also observed a strong photolumines-<br />

272


FIGURE 29. X-ray diffraction from UDD. (Reprinted from Ref. 167,<br />

Copyright 1995, with permission from the American Institute of Physics.)<br />

cence (PL) simultaneously with the Raman signal<br />

of diamond. The maximum of the PL shifted from<br />

2.14 to 2.3 eV, while the annealing temperature<br />

was increased over the range from 1100 to 1600 K,<br />

and correspondingly the particle size changed from<br />

4.5 to 2 nm. The PL signal disappeared simultaneously<br />

with the disappearance of the Raman<br />

signal of diamond. These authors estimated the<br />

size of diamond crystals from the phonon confinement<br />

model to be 4.5 nm.<br />

X-ray Spectroscopy<br />

Quantum confinement effects are expected in<br />

nanocrystalline semiconductors when the size of<br />

the particle is smaller than some critical value.<br />

However, details of the electronic structure of<br />

nanodiamonds are still unknown. Chang et al.<br />

studied CVD diamond films with various crystallite<br />

sizes by X-ray absorption spectroscopy. 170<br />

The size of the diamond crystallites in the films<br />

ranged from 3.5 nm to 5 µm. Figure 31 shows the<br />

exciton state and conduction band edge of the<br />

nanodiamonds as a function of the crystallite<br />

size. 170 According to this plot, the conduction<br />

band edge shifts to higher energy with a decrease<br />

in the crystallite size. A strong change in energy<br />

structure, which is believed to be due to quantum<br />

confinement, takes place for sizes smaller than 10<br />

nm. The authors applied a standard analytical<br />

approximation for quantum confinement shift of<br />

energy levels in a one-dimensional quantum box<br />

of size a ∆E ≈ π 2 h<br />

2<br />

* 2<br />

and derived the value of<br />

2ma<br />

effective mass for nanodiamond of about 0.1m 0 .<br />

The above results and their interpretation raise<br />

several questions. First, the objects of this study<br />

were not isolated diamond crystallites but a complex<br />

composite material consisting of diamond<br />

crystallites of different sizes, grain boundaries,<br />

amorphous and possibly graphitic inclusions, etc.<br />

Second, Ley et al. 171 questioned the validity of the<br />

273


FIGURE 30. Raman spectra from UDD and bulk diamond. (Reprinted from Ref. 167,<br />

Copyright 1995, with permission from the American Institute of Physics.)<br />

method the authors used for the extraction of the<br />

values of energy shifts. Third, the calculated effective<br />

mass of nanodiamond of 0.1m 0 is considerably<br />

smaller than the bulk value. Note that in<br />

the limiting case of a→0, when the particle approaches<br />

one atom, m*→m 0 , that is, the effective<br />

mass increases. Fourth, the reported widening<br />

energy gap, remarkable already at a diamond<br />

particle size of about 10 nm, is in contradiction<br />

with theoretical results predicting the critical size<br />

for the band structure size-effect onset to be 2 to<br />

2.5 nm. 172 In 2002, Raty et al. presented theoretical<br />

results suggesting that the widening of the<br />

energy gap is remarkable only for the size of the<br />

diamond particle


FIGURE 31. The exciton state and conduction band edge of the<br />

nanodiamonds as a function of the crystallite size. (Reprinted from Ref.<br />

170, Copyright 1998, with permission from the American Physical Society.)<br />

to obtain bonding information. For example, it<br />

clearly shows sp 3 or sp 2 bonding of carbon atoms.<br />

In principle, the EELS sensitivity in compositional<br />

analysis can be extended to the single-atom<br />

limit, as was shown by Suenaga et al. 173 Because<br />

EELS is usually associated with high-resolution<br />

TEM, it is possible to collect information about<br />

bonding with very high spatial resolution within<br />

a nanoparticle. EELS in a TEM can in principle<br />

improve the accuracy of size measurements. In<br />

Ref. 174 the approximate diameters measured in<br />

a TEM were later adjusted by fitting theoretical<br />

and experimental EELS spectra. The fitting procedure<br />

was based on the spherical model shown<br />

in Figure 33, and both the amorphous layer thickness<br />

and crystalline diamond core diameter were<br />

varied to obtain the best fit. Table 8 shows measured<br />

diameters of diamond particles of three different<br />

sizes compared with calculated values derived<br />

by matching experimental EELS data.<br />

It should be noted that the assumption of a<br />

spherical shape for the diamond nanoparticles led<br />

to a considerable difference between the direct<br />

TEM measurements and EELS adjusted values.<br />

After more detailed investigations of TEM images,<br />

it was found that the particles were not<br />

really spherical. Taking into account the<br />

nonspherical shape resulted in better agreement<br />

between the EELS predicted and TEM measured<br />

diameters. Also, from their analysis the authors<br />

conclude that theoretical predictions for the case<br />

of crystalline diamond surrounded by a surface<br />

layer of sp 2 (e.g., graphitic) material were definitely<br />

not capable of reproducing the experimental<br />

results. 174<br />

Shape of Diamond Nanoparticles<br />

The shape of UDD particles 3 to 5 nm in size<br />

is often assumed to be spherical. As a model, a<br />

275


FIGURE 32. Histogram of diamond crystallites of various sizes directly measured in a HREM lattice image<br />

of a CVD nanodiamond film. (After Ref. 2.)<br />

spherical shape is easier for calculations. At the<br />

same time, there are experimental reports that<br />

most of the diamond particles have very spherical<br />

shape. 174,175,176 This led to speculation that the<br />

diamond nanoparticles are formed from liquid<br />

drops. Chen and Yun 175 argued that from the point<br />

of view of the carbon phase diagram, thermodynamic<br />

conditions created by explosive detonation<br />

(temperature range 3000 to 3500 K, pressure 30<br />

to 35 GPa do not correspond to the liquid state<br />

region of diamond crystal). The authors 175 suggested<br />

that this contradiction could be resolved<br />

by assuming that while bulk diamond cannot melt<br />

under the detonation conditions, for the nanometer-sized<br />

diamond particles the effective melting<br />

point might decrease. Thus, Chen and Yun concluded<br />

that the formation mechanism of nanosized<br />

diamond is through coagulation of liquid<br />

droplets. 175 One should mention that the statements<br />

about the spherical shape of diamond<br />

nanoparticles are usually based on TEM pictures<br />

without considerable magnification, where the<br />

particles appears so small (as for example, in<br />

Ref. 175) that it is impossible to make any assessments<br />

about shape. Sometimes particles are<br />

shown in TEM with larger magnification, but<br />

without isolation from each other and from the<br />

amorphous matrix so again it is difficult to assess<br />

the shape.<br />

Vereschagin and Sakovich also argued in favor<br />

of a spherical shape for UDD nanocrystals<br />

and the liquid carbon drop formation mechanism.<br />

168 Also, they compared the density of<br />

nanodiamonds calculated from the lattice parameters<br />

(3.527 g/cm 3 ) with the experimental values<br />

of the helium pycnometric density of UDD, which<br />

276


FIGURE 33. A model for a diamond nanoparticle consisting of a<br />

sphere of crystalline diamond with diameter d c , surrounded by a<br />

layer of amorphous carbon with thickness t a . The resulting diameter<br />

of the particle is d=d c +t a .<br />

277


is 3.05 to 3.10 g/cm 3 . Their conclusion was that<br />

the difference between calculated (X-ray) and<br />

pycnometric densities supports the liquid phase<br />

model of nanodiamond formation. Moreover, the<br />

authors 168 suggested that diamond nanoparticles<br />

formed by rapid crystallization are hollow spheres!<br />

They estimated from the density difference that<br />

the cavity inside 4 nm particle should be 1.77 nm<br />

in diameter.<br />

Surface Properties<br />

Maillard-Schaller et al. investigated the surface<br />

and electronic properties of nanometer-sized<br />

diamond particles (UDD Ndp2) by X-ray photoelectron<br />

spectroscopy (XPS) and UV photoelectron<br />

spectroscopy (UPS). 177 In these experiments,<br />

nanodiamonds were deposited on flat Si(100)<br />

substrates by electrophoresis. Figure 34 shows<br />

the XPS spectra of a nanodiamond powder film<br />

on Si(100). According to the XPS analysis, the<br />

UDD powder did not contain any detectable impurities<br />

except nitrogen. The N content was estimated<br />

to be 1 to 2 at. %. The as-deposited UDD<br />

films showed a strong oxygen peak. After treatment<br />

in a hydrogen microwave plasma and transfer<br />

in air to the XPS system, the nanodiamond<br />

films were found to be almost free from oxygen<br />

contamination, as can be seen in Figure 34.<br />

HeI (hν=21. 2 eV) and HeII (hν=40.8 eV)<br />

UV photoelectron spectroscopy measurements<br />

have been performed on the as-deposited and<br />

hydrogen plasma-treated nanodiamond films.<br />

Based on the UPS spectra (Figure 35), the energy<br />

band diagram of nanodiamond has been proposed<br />

(Figure 36). The electron affinity χ of UDD films<br />

was calculated using the emission width W and<br />

the low kinetic energy cut-off in the experimental<br />

UPS spectra (Figure 35): χ=hν-E g -W.<br />

The emission width of the as-deposited sample<br />

was 15.0 eV with a low-energy cut-off at 3.1 eV,<br />

which results in a positive value of electron affinity<br />

of +0.7 eV, assuming that the band gap E g =5.5<br />

eV. After H 2 plasma treatment, the samples showed<br />

an emission width of 15.9 eV with a very sharp<br />

low-energy cut-off at 2.1 eV, indicating a negative<br />

electron affinity of –0.2 eV (Figure 36).<br />

Electrical Characterization<br />

There are only a very few experimental reports<br />

on electrical and electronic properties of<br />

UDD. Some measured collective electrical characteristics<br />

of powders and suspensions are given<br />

FIGURE 34. Survey XPS spectra of the UDD film (a) before and (b) after the H 2<br />

plasma treatment. (Reprinted from Ref. 177, with permission from Elsevier Science.)<br />

278


FIGURE 35. HeI and HeII UPS spectra of as deposited and H 2 plasma-treated<br />

UDD surfaces. (Reprinted from Ref. 177, with permission from Elsevier Science.)<br />

FIGURE 36. Energy band diagram of the hydrogenated nanodiamond powder film. (Reprinted<br />

from Ref. 177, with permission from Elsevier Science.)<br />

279


in Table 7. Gordeev et al. reported a resistivity for<br />

bulk nanodiamond composites obtained by pressing<br />

UDD powders. 178 Such materials have very<br />

high porosity typically in the range 60 to 70 vol.%. 178<br />

The room temperature resistivity of such UDD<br />

composites was 1.2 × 10 9 W-m, consistent with the<br />

resistivity reported by Dolmatov (Table 7).<br />

Belobrov et al. reported results of studies of<br />

UDD powders with different surface modifications<br />

by electron parmagnetic resonance and<br />

nuclear magnetic resonance (NMR) spectroscopies.<br />

179 An example of a nanodiamond EPR<br />

spectrum is shown in Figure 37. The authors 179<br />

found that the nanodiamond EPR signal is independent<br />

from the chemical modification of the<br />

nanodiamond surface. The g-factor for UDD was<br />

found to be 2.0027(5). 179 Belobrov et al. concluded<br />

that the paramagnetic properties of the<br />

nanodiamond are determined only by the sp 3 core<br />

of the diamond particle.<br />

A 13 C NMR spectrum for nanodiamond is<br />

shown in Figure 38. According to the authors, 179<br />

the resonance line is assymetric and well decomposed<br />

into two Gaussian components,<br />

whose numerical values are presented in the<br />

table in Figure 38. The authors interpreted the<br />

narrow line with δ=35.1 ppm as relating to<br />

diamond carbon, while the wide line with δ=34.2<br />

ppm is caused by distortion of the tetrahedral<br />

coordination. The authors 179 concluded that only<br />

30% of bonds in nanodiamond are nondistorted<br />

sp 3 bonds, while the remaining 70% of carbon<br />

bonds are distorted but still with sp 3 hybridization.<br />

For comparison, natural jewel-quality diamonds<br />

have a characteristic chemical shift δ=50<br />

ppm.<br />

Zhirnov et al. characterized UDD particles<br />

using electron field emission measurements by<br />

putting small a amount (0.2 µm in thickness) of<br />

UDD on metal or silicon tips by electrophoresis.<br />

180,181 In these experiments, the field emission<br />

characteristics of tips with UDD coatings were<br />

compared to characteristics of bare tips. The<br />

emission experiments showed that the emission<br />

characteristics differ significantly for different<br />

UDD coating conditions. The different UDD<br />

coating conditions were obtained from one original<br />

UDD powder (marked as Nd) using different<br />

physical and chemical treatments. The modifications<br />

differed in concentration of impurities,<br />

pH of water suspension, and density as shown in<br />

Table 9. 180<br />

FIGURE 37. Central part of the EPR spectrum of nanodiamond with the Li standard (g=2.0023).<br />

The scan and the modulation are 50 and 0.01 mT (After Ref. 179, with permission from P.<br />

Belobrov.)<br />

280


FIGURE 38. 13 C spectrum for nanodiamond. The numerical values of chemical shifts, intensities, and line widths after<br />

decomposition into Gaussian components are given in the table. (After Ref. 179, with permission of P. Belobrov.)<br />

281


A summary of the emission characteristics<br />

for different UDD coatings is given in Table 10.<br />

The emission characteristics of UDD are discussed<br />

in more detail in Ref. 180. An interesting result is<br />

the difference in the emission characteristics of<br />

UDD with modifications NdP and NdP1 (Figure<br />

39). NdP1 was prepared using the heavier (bottom)<br />

fraction of the original NdP modification,<br />

and the main difference between the two was the<br />

average crystallite size (larger for NdP1). This<br />

result indicates critical role of diamond crystal<br />

size in the nanoscale regime on the emission properties<br />

of UDD-coated tips.<br />

Characterization of Isolated Nanodiamond<br />

Particles<br />

At this point, most of reported results on characterization<br />

of nanodiamond describe collective<br />

properties of multiparticle systems such as CVD<br />

nanodiamond films, agglomerates of UDD, etc.<br />

Such systems are composite materials, consisting<br />

of sp 3 -bonded diamonds often surrounded by<br />

amorphous carbon, sp 2 -bonded graphitic inclusions,<br />

etc. Correspondingly, it is often difficult to<br />

extract specific information about 3 to 6 nm primary<br />

nanodiamond crystallites. Moreover, in some<br />

cases the properties of nanodiamonds can be<br />

modified due to the presence of surrounding materials.<br />

For example, the surface energy, and correspondingly<br />

shape of particles, is very sensitive<br />

to their surroundings. The characterization of isolated<br />

nanodiamond particles is an important but<br />

challenging task. There are few reports of characterization<br />

of isolated diamond particles 40 to 100<br />

nm in size (see, for example, the next section).<br />

However, the smallest nanodiamond crystallites<br />

with a size of 3 to 6 nm are very difficult to<br />

characterize in isolation, due to their natural trend<br />

to form agglomerates.<br />

Tyler et al. recently reported a new technique<br />

for isolating individual nanodiamond particles by<br />

depositing them on sharp (radius 10 to 50 nm)<br />

metal tips by pulsed electrophoresis from alcohol<br />

suspensions of UDD. Apparently, the very high<br />

electric fields near the sharp tips breaks the agglomerates,<br />

and thus it is possible to manipulate<br />

individual nanodiamond particles. 182 Examples of<br />

metal tips with nanometer-size diamonds are<br />

shown in Figures 15 and 40. The typical size of<br />

diamond nanoparticle as measured in TEM was<br />

about 3 nm (Figure 40). It should be noted that in<br />

all cases only particles with faceted shapes were<br />

observed.<br />

More detailed information about the preparation<br />

procedure and results of characterization can<br />

be found elsewhere. 183,184 Preliminary results of<br />

field emission characterization of UDD reveal a<br />

considerable difference in emission behavior of<br />

single isolated diamond nanoparticle and multiparticle<br />

nanodiamond thin film. 184<br />

Nonclassical Fluorescence from<br />

Diamond Nanoparticle<br />

A new area for physical experimentation with<br />

possible applications in advanced information processing<br />

is the generation of light sources that are<br />

able to emit individual photons on demand.<br />

Beveratos et al. investigated the quantum properties<br />

of the light emitted by diamond nanocrystals<br />

containing a single nitrogen-vacancy color center.<br />

185 The typical size of diamond particles used in<br />

these experiments was 40 nm. According to the<br />

authors, there are several very important optical<br />

properties of diamond nanocrystals that are impor-<br />

282


FIGURE 39. Current-voltage characteristics of Si tips with UDD coatings of the same<br />

modification with different size of diamond crystallites: NdP and NdP 1 . (Reprinted from<br />

Ref. 180, Copyright 1999, with permission from the American Institute of Physics.)<br />

tant for their use as individual photon light sources.<br />

First, the subwavelength size of these nanocrystals<br />

renders refraction irrelevant. A nanocrystal can be<br />

regarded as a point light source. Second, the very<br />

small volume of diamond excited by the pump<br />

light yields very small background light. This is<br />

very important for single photon sources.<br />

By exciting nanodiamond crystals using a<br />

YAG laser (l=532 nm), the authors were able to<br />

observe fluorescence with almost backgroundfree<br />

photon antibunching from single nitrogenvacancy<br />

centers in diamond nanocrystals at room<br />

temperature. The excited state lifetime in the bulk<br />

is t b =11.6 ns. The measured lifetime in the<br />

nanocrystals was 25 ns. The authors argue that<br />

this lifetime modification is a quantum electrodynamic<br />

effect.<br />

3. Atomistic Simulations on Diamond<br />

Clusters<br />

a. Simulations of Diamond Clusters:<br />

Structural Properties<br />

Below we discuss two questions related to the<br />

number of atoms, the fraction of surface atoms in<br />

a diamond particle of a particular size, and possible<br />

shapes of a single nanodiamond particles. Although<br />

in reality diamond nanoparticles contain functional<br />

groups at the surface (Section III.B.2), the present<br />

analysis is restricted to hydrocarbon systems.<br />

How Many Atoms Are in a Single ND<br />

Particle?<br />

Plotted in Figure 41a is the number of atoms<br />

in a spherical diamond particle as a function of<br />

particle size estimated from the formula:<br />

Ntot = 4<br />

3V<br />

R 3<br />

π , (5)<br />

0<br />

where V 0 = 5.667 Å 3 is atomic volume in bulk<br />

diamond. For example, a particle with a diameter<br />

of 4.3 nm the equation yields about 7200 atoms,<br />

while a particle with diameter 10 nm is predicted<br />

to contain about 92,000 atoms.<br />

Another important characteristic that influences<br />

particle properties is the number of surface<br />

atoms. Assuming<br />

n<br />

S<br />

= ( n ) 23 /<br />

(6),<br />

Bulk<br />

283


FIGURE 40. Isolated nanodiamond particle on molybdenum tip. 184<br />

where n s and n bulk are surface and bulk atomic<br />

densities, respectively, the following analytical<br />

expression is often used to estimate the number of<br />

surface atoms in a spherical particle:<br />

N<br />

S<br />

3<br />

/ Ntot<br />

=<br />

13 /<br />

(7)<br />

4Rn<br />

Bulk<br />

It is possible to cut out a sphere of a given radius<br />

and count the number of atoms with a coordination<br />

number less than four at the surface of the<br />

sphere. The difference between the fraction of the<br />

surface atoms evaluated this way and from Eq. 6<br />

can be as big as 2 to 2.5 times. This difference<br />

arises from the presence of various surface facets<br />

in the second method for counting the number of<br />

surface atoms that are not accounted for in Eq. 6,<br />

which assumes a density of surface atoms equal<br />

to that for a (001) facet. The number of surface<br />

atoms for a spherical cluster as a function of<br />

cluster diameter taking into account surface<br />

faceting is plotted in Figure 41.<br />

What is the Shape of a Nanodiamond<br />

Cluster?<br />

Until recently, most of the experimental work<br />

dealing with nanodiamond produced by means of<br />

detonation described the shape of clusters as being<br />

spherical (Section III.A.2.b). Indeed, HRTEM<br />

pictures of nanodiamond agglomerates resemble<br />

spherical forms (Figure 42a) as well, as, for example,<br />

nanodiamond clusters embedded in metal-<br />

284


FIGURE 41. Total number of atoms and the number of surface atoms in a spherical nanodiamond<br />

particle as a function of particle size. Bottom image provides more details for smaller particles.<br />

lic matrices after chlorination of metal carbide<br />

(Figure 23). However, recent HRTEM images of<br />

a single nanodiamond cluster on the surface of a<br />

Mo tip clearly indicate the presence of facets at<br />

the particle surface, with the cluster resembling a<br />

polyhedral shape (Figure 15). This shape is similar<br />

to that of microscopically sized diamond particles<br />

formed in the gas phase during a CVD<br />

process as first was reported by Matsumoto and<br />

Matsui 29 (Figure 42b). The typical habit of these<br />

clusters was regularly shaped cubo-octahedron<br />

and twinned crystals, that is, twinned cubo-octahedron,<br />

an icosahedron and a decahedral-Wulffpolyhedron.<br />

The particles have exposed (100) and<br />

(111) facets and often possess chamfered edges<br />

along certain directions (such as ) (Figure<br />

42b). Multiply twinned nanodiamond particles<br />

have had been observed in meteorites (Figure 16).<br />

The shape of a diamond cluster is inherently<br />

connected to its stability, which in turn depends<br />

on the state of the surface atoms. Important factors<br />

that impact cluster stability include the presence<br />

of active groups on the surface of UDD<br />

particles after different types of purification treatments,<br />

hydrogenation of the surface, and possible<br />

surface reconstructions on a bare surface that eliminates<br />

(or at least reduces) the number of dangling<br />

bonds. As discussed in Section II based on ab<br />

285


FIGURE 42. TEM images of diamond clusters revealing different<br />

shapes: spherical (a) shapes of nanodiamond particles (After Ref.<br />

90, courtesy of G. Galli) and well-faceted diamond particles of<br />

micron and submicron sizes (b). (After Ref. 29.)<br />

initio simulations, bare diamond surfaces, containing<br />

(111) planes undergo ‘buckification’, 90,99–101<br />

that might be responsible for the observed spherical<br />

shapes of diamond particles. Particles containing<br />

(100) bare surfaces reconstructed in a<br />

manner similar to bulk diamond surfaces. 99–101 In<br />

addition, relaxation and geometrical parameters<br />

of hydrogenated nanodiamond surfaces were quite<br />

similar to those of bulk diamond. 99–101 Based on<br />

these facts, we performed simple evaluations of<br />

binding energies of hydrogenated clusters with<br />

different morphologies.<br />

Results of binding energies of fully hydrogenated<br />

spherical, cubo-octahedron and truncated in<br />

one direction octahedron clusters are plotted in<br />

Figure 43. Binding energies have been calculated<br />

relative to systems with the same number of carbon<br />

atoms in bulk diamond and hydrogen atoms<br />

as H 2 molecules. The results of DFT/LDA calculations<br />

by Kern and Hafner 186 on energetics of<br />

reconstructed (001)(2 × 1):H and (111)(1 × 1):H<br />

diamond surfaces were used for the calculations<br />

(summarized in Table 11). A dimer reconstruction<br />

of all (100) diamond surfaces has been assumed.<br />

The calculations suggest that a truncated octahedron<br />

cluster is more stable than a cubo-octahedron<br />

or a spherical cluster with the same number<br />

of atoms (Figure 43a). As apparent from the<br />

figure, the difference in binding energy between<br />

286


FIGURE 43. Binding energies of hydrogenated diamond clusters as a function of particle size.<br />

Binding energies calculated relative to systems with the same number of carbon atoms in bulk<br />

diamond and hydrogen atoms as H 2 molecules. LDA data 186 (Table 11) on energetics of (001) and<br />

(111) diamond surfaces are used for the calculations.<br />

287


spherical and cubo-octahedron clusters is small,<br />

on the order of ~0.01 to 0.02 eV/atom (for a<br />

comparison, the difference in cohesive energies<br />

of bulk diamond and graphite is about 0.02 eV/<br />

atom). These results can be explained using simple<br />

arguments about the relative fractions of surface<br />

atoms with favorable and unfavorable energies<br />

for clusters with particular shapes. As follows<br />

from Table 11, one-, two-, and three-coordinated<br />

carbon atoms at a diamond surface have different<br />

energies. Thus, a larger number of atoms with<br />

unfavorable on-site energies (such as, for example,<br />

twofold coordinated atoms at a (001) facet) would<br />

result in lower cluster stability. Indeed, as illustrated<br />

in Figure 44, the fraction of two-coordinated<br />

atoms (accordingly dimerized and hydrogenated<br />

in a final structure) is larger for cubo-octahedron<br />

clusters, resulting in the greatest cluster stability<br />

compared with the spherical and octahedron structures.<br />

It should be noted that spherical clusters resemble<br />

truncated octahedrons (Figure 45a) with<br />

additional adatoms (for bigger clusters — atomic<br />

islands) at the center of their facets and are intermediate<br />

shapes between octahedron and cubo-octahedron<br />

regarding the number of surface atoms with<br />

nonfavorable energies. The fraction of atoms with<br />

three dangling bonds at the surface of a spherical<br />

cluster is relatively small (Figure 45a). An analysis<br />

of their location at a surface of a spherical cluster<br />

show that they do not have undercoordinated nearest<br />

neighbors with which to form a (2 × 1) surface<br />

reconstruction. Thus, singly coordinated atoms would<br />

be rather etched or terminated by three hydrogen<br />

atoms. In general, from the analysis above, it can be<br />

concluded that the most stable shape of a fully hydrogenated<br />

diamond cluster is that with the maximum<br />

number of ‘energetically favorable’ surface<br />

atoms, for example, hydrogenated atoms with three<br />

carbon atoms as neighbors. The octahedron or slightly<br />

truncated octahedron cluster are primary candidates<br />

satisfying this condition.<br />

Due to the very small difference in stability<br />

between different carbon structures, it can be also<br />

expected that excitation (for instance, by an electron<br />

microscope) may be sufficient to induce transitions<br />

in particle shape. This may be the primary<br />

reason that different shapes of diamond clusters<br />

such as spherical (Figure 42a) or faceted (Figure<br />

15) are observed. As discussed in Section II, electron<br />

beam irradiation can even induce phase transitions,<br />

such as carbon onions to nanodiamond<br />

structures and vice versa. Similarly, Ijima and<br />

Ichihasi 187 showed that at the nanoscale, the shape<br />

of metallic particles is not necessarily constant<br />

because the energy of a nanoparticle shows many<br />

local minimum energy configurations, corresponding<br />

to different structures. A gold particle ~2 nm<br />

in size that was exposed to strong electron beam<br />

irradiation with a beam intensity between 15 and<br />

80 amp/cm 2 fluctuated between the cubo-octahedral,<br />

icosahedral, and single-twinned structures. 187<br />

Although not quite realistic, energetic characteristics<br />

of all-carbon nonreconstructed diamond<br />

clusters were evaluated using a bond order potential,<br />

188 which indicated stable relaxed diamond<br />

structures contrary to the first principle simulations.<br />

The results of these calculations indicated<br />

that cohesion energy, surface excess energies, and<br />

stiffness of spherical diamond clusters with<br />

nanoscale dimensions deviate considerably from<br />

their macroscopic values. For clusters of 2.2 nm<br />

(1100 atoms) and 4.8 nm (10 4 atoms), the calculated<br />

cohesive energy, was about of 95% and 99%<br />

of the bulk value, respectively. From an analysis<br />

of the asymptotic value of surface energy it was<br />

concluded that the spherical surface possess primarily<br />

a (100) character. Regarding stiffness, it<br />

was found that for clusters 2.2 nm and 5.8 nm<br />

(2 × 10 4 atoms) in diameter the stiffness was about<br />

92% and 99% of the bulk value, respectively.<br />

A current conclusion is that hydrogenated alldiamond<br />

clusters are the most energetically preferable<br />

forms, especially octahedrons and partially<br />

truncated octahedrons. Starting with CH 4 ,<br />

adamante (C 10 H 8 ), etc., all hydrogenated diamond<br />

clusters are mechanically stable.<br />

In contrast, if the (111) surface of a<br />

nanodiamond particle is not hydrogen terminated,<br />

the resulting structure would be an encapsulated<br />

fullerene at small particle sizes and ‘bucky diamond’<br />

with a diamond core and fullerene-like<br />

outer shell 91,99–101 (Figure 45c). Buckification has<br />

been observed for spherical particles at least up to<br />

3 nm in diameter. 91 The authors 91 performed ab<br />

initio calculations using the GGA density functional<br />

for the smallest size clusters, and a<br />

semiempirical tight-binding Hamiltonian for systems<br />

up to 3 nm (2425 atoms) in diameter. Both<br />

288


FIGURE 44. Fractions of one-, two-, and three-coordinated surface atoms<br />

(relative to the total number of surface atoms) for cubo-octahedron (a), spherical<br />

(b), and truncated in one direction octahedron clusters (c) as a function of<br />

number of atoms in a cluster.<br />

289


FIGURE 45. Structural models of nanodiamond clusters. Spherical (a) or polyhedral (b) (cubo-octahedron) shapes<br />

with all four-coordinated carbon atoms can be preserved if the cluster surface is terminated by hydrogen. Twocoordinated<br />

atoms on the surface of a spherical cluster are dark colored (a). This illustrates that a spherical cluster<br />

represents truncated octahedron with several adatoms in the center of its (001) or (111) facets. Bare surface of<br />

nonhydrogenated diamond clusters experience surface reconstruction resulting in a cluster with a diamond core and<br />

a fullerene-like shell(c) (courtesy of Guilia Galli 90 ).<br />

methods gave similar structural models for the<br />

smallest cluster sizes, indicating validity of the<br />

results for structures of larger clusters obtained<br />

with the tight-binding method. According to Ref.<br />

91, starting from ideally terminated diamond particles,<br />

the reconstruction occurs spontaneously at<br />

low temperature. The barrier between the ideal<br />

surface structure and the reconstructed surface is<br />

size dependent and increases as the size of the<br />

cluster is increased, achieving a value of the order<br />

of several tens of eV for the largest clusters as<br />

found in bulk diamond. 189 It is interesting to define<br />

the critical size, when diamond clusters will<br />

have an all-diamond structure (probably with reconstructed<br />

(111) surfaces forming Pandey chains<br />

similar to bulk diamond surfaces).<br />

In support of their fullerene-like structural<br />

model, the authors provide an X-ray absorption<br />

spectra of nanodiamond agglomerates. 90 The preedge<br />

signal of the nanodiamond spectra differs<br />

significantly from those of diamond and graphite<br />

and exhibits a characteristic 2-peak feature resembling<br />

that of a buckyball and C 70 . These<br />

nanodiamond peaks had been interpreted as the<br />

signature of a mixture of pentagons and hexagons<br />

on the reconstructed surface of the diamond core.<br />

b. Simulations of Diamond Clusters:<br />

Electronic Properties<br />

Relationships between various electronic properties<br />

and cluster size were explored with a selfconsistent<br />

environment dependent tight-binding<br />

(SCEDTB) model with the C-H terms fit to first<br />

principles electronic structure results for ethane,<br />

290


methane, benzene, and hydrogenated and<br />

diamond surfaces. 191 Plotted in Figure 46<br />

is the density of states for four diamond clusters<br />

and for bulk diamond. The bulk density of states<br />

is shifted by the averaged Coulomb potential due<br />

to the surface dipole layer (cf. Figure 47) experienced<br />

by carbon atoms in the cluster. All four<br />

clusters demonstrate a size dependence of the<br />

band gap (Figure 48). Dimensional effects are<br />

most apparent with respect to the states in the<br />

valence band; the highest-occupied molecular<br />

orbitals for the 34 and 161 atom clusters lie approximately<br />

2.5 eV and 1.25 eV, respectively,<br />

below the bulk valence band edge. At the same<br />

time even for the smallest cluster the energy of<br />

the lowest-unoccupied molecular orbitals coincide<br />

with the bulk conduction band edge. This<br />

result can be intuitively expected because the states<br />

with higher energies and smaller wavelengths are<br />

less sensitive to the dimensions of the system.<br />

Dimensional band gap widening is less than 0.4<br />

eV for the largest cluster examined (913 carbon<br />

atom; ~2 nm diameter), which leads to the conclusion<br />

that any band gap size effect is insignificant<br />

for cluster sizes larger than about 2 to 2.5<br />

nm. Minor deviations from the bulk spectrum for<br />

the largest cluster are due mainly to the presence<br />

of hydrogen states and not to the finite dimensions<br />

of the cluster. This result apparently disagrees<br />

with indirect X-ray measurements, which<br />

indicate that in a 3.6-nm cluster the conduction<br />

band edge position is 1.2 eV above the conduction<br />

band edge for bulk diamond. 170 At the same<br />

time, however, the results are consistent with recent<br />

ab initio calculations 90 (Figure 48b). Depending<br />

on the simulation approach, the authors<br />

conclude, the band gap for diamond clusters approaches<br />

that for bulk diamond at cluster sizes of<br />

about 2 nm according to Quantum Monte Carlo<br />

calculations, and about 1 to 1.2 nm according to<br />

density functional calculations using the generalized<br />

gradient approximation. 90 According to the<br />

FIGURE 46. Electronic density of states for the four clusters with a truncated octahedron shape (histogram) and for<br />

bulk diamond (solid line). The histogram at the bottom of the pane (a) is a spectrum generated from density functional<br />

theory. (After Ref. 191.)<br />

291


FIGURE 47. Coulomb potential distributions for the nanodiamond clusters\plotted along the and <br />

directions passing through the centers of mass of the clusters. Octahedron clusters are truncated along the <br />

axis and contain 34 (a), 161 (b), 435 (c) and 915 carbon atoms (d), correspondingly. (After Ref. 172.)<br />

FIGURE 48. Band gap vs. cluster size for fully hydrogenated diamond clusters calculated with a SCTB model 190 (a),<br />

and DFT /GGA (generalized gradient approximation) and DFT/TDLDA (time dependent LDA) methods 90 (b). In<br />

Figure 48b the HOMO-LUMO gaps are indicated by triangles. The dashed line indicates the bulk (indirect) gap<br />

calculated with DFT/GGA. Open circles correspond to calculations with TDLDA. 90 The fact that for larger clusters<br />

the band gap is lower than that for the bulk structure is due to a stress effect induced by hydrogen on the surface. 90<br />

The bulk band gap calculated with the SCTB approach (a) is 5.5. eV.<br />

292


esults on X-ray adsorbtion and emission experiments,<br />

no appreciable changes in band gap is<br />

observed for nanodiamond clusters with sizes<br />

exceeding 2 nm. 90<br />

Illustrated by Figure 49 are electronic properties<br />

of carbon nanoclusters with highly distorted<br />

bonds (consisting completely of 5 to 7 rings) as<br />

well as clusters with sp 2 /sp 3 bonding. According<br />

to the figure, the band gap becomes filled with the<br />

presence of sp 2 bonding.<br />

Coulomb potential distributions for the clusters<br />

obtained with the self-consistent environment<br />

dependent tight binding calculations 191 are plotted<br />

in Figure 47. The main feature of the potential<br />

distribution is a sharp rise at the cluster surface<br />

produced by hydrogen termination. Cluster size<br />

effects are apparently only significant for the<br />

smallest cluster. For the larger clusters the calculations<br />

predict that the potential inside the cluster<br />

does not change appreciably with increasing cluster<br />

size. In conjunction with the spectra plots the<br />

potential rise at the boundary gives a -1.4 5 eV<br />

electron affinity of the hydrogenated surface.<br />

This value is close to the experimentally<br />

FIGURE 49. Density of states of two nanocarbon clusters with highly disordered<br />

sp3 bonds (a) as well as with mixed sp 2 /sp 3 bonding (b). Clusters contain 435<br />

carbon atoms. (After Ref. 190.)<br />

293


measured electron affinity, 192 while density functional<br />

theory usually overestimates experimental<br />

values by ~0.6 to 0.8 eV. 193<br />

One of the major focuses in nanoelectronic<br />

applications of nanodiamond is the possibility of<br />

n- and p-type doping of the clusters. Recent first<br />

principles calculations addressed the structural<br />

stability of nanodiamond clusters doped with B,<br />

N, and P in substitutional positions. 194,195 Small<br />

hydrogenated diamond clusters (a few tens of<br />

C atoms) containing B, N, or P atoms in the center<br />

maintained the diamond lattice during relaxation.<br />

For future research, it would be interesting to<br />

evaluate substitutional energies of different<br />

dopants in the subsurface positions vs. position in<br />

the inner parts of nanoparticles, as well as their<br />

diffusional properties. Another interesting issue<br />

would be the difference in dopant behavior in<br />

nanoclusters as compared to that in bulk diamond<br />

when well with diamond surfaces at the<br />

macroscale.<br />

A major conclusion of the studies of electronic<br />

properties of nanodiamond clusters is that<br />

quantum confinement effects disappear at cluster<br />

sizes larger than approximately 2 nm, in contract<br />

to silicon and germanium clusters where quantum<br />

confinement effects persist up to 6 to 7 nm clusters.<br />

90 A practical realization of the size dependence<br />

of quantum confinement effects in ND clusters<br />

is likely difficult because of the very small<br />

cluster size (below ~1000 atoms for 2 nm cluster).<br />

In addition, according to the size distribution function<br />

of the UDD particles, only a small fraction of<br />

the clusters is produced with a 2 nm diameter.<br />

c. Diamond Nanorods<br />

Would Diamond Nanorods be Stronger<br />

than Fullerene Nanotubes?<br />

Diamond was always considered the strongest<br />

material in the macroscopic world until the<br />

extreme mechanical properties of carbon<br />

nanotubes were discovered (Table 13). After that<br />

discovery, numerous papers claimed that carbon<br />

nanotubes were the strongest material. It is difficult<br />

to make a fair comparison between two representatives<br />

from the macro- and nano-worlds<br />

unless some assumptions related to particular structures<br />

are made. Below we compare the mechanical<br />

properties of these two materials at equal conditions,<br />

assuming that it is possible to make diamond<br />

nanorods, and briefly discuss the routes for synthesis<br />

of one-dimensional diamond.<br />

When extrapolating the mechanical properties<br />

of carbon nanotubes to the macroscale, an<br />

ambitious assumption has to be made regarding<br />

their wall thickness; it is usually assumed to be<br />

equal to the interplanar spacing of graphite. At the<br />

nanoscale, an unambiguious characteristic of<br />

strength would be maximum force before failure<br />

for a given structure, rather than maximum stress.<br />

At least, it would be a useful characteristic for<br />

comparison of fiber-like nanostructures from different<br />

materials with similar diameters. For this<br />

force–based definition of a nanostructure strength,<br />

Figure 50 illustrates the dependence of fracture<br />

load vs. diameter for single wall nanotubes and<br />

diamond nanorods for three orientations, corresponding<br />

to the low-index axis (,,<br />

). 196 The calculations for nanotubes were<br />

done using ideal strength values for graphene<br />

sheets from ab-initio pseudopotential total energy<br />

calculations within the local density approximation<br />

228 (values are summarized in Tables 12 and<br />

13). Similar strength values for different diamond<br />

orientations were obtained by the same authors 197<br />

and by similar techniques, 198 so all the input parameters<br />

for the evaluations are self-consistent.<br />

Fracture forces are evaluated for MWNTs with a<br />

variable outer diameter and a fixed inner diameter<br />

of 4 nm, which corresponds to typical experimental<br />

values. 39<br />

The results of Figure 50 demonstrate that at<br />

small diameters, carbon nanotubes are stronger than<br />

diamond rods because of the superior strength of<br />

the single bonds in graphene over those in diamond<br />

(Table 13). However, as the load-bearing area and,<br />

correspondingly, the number of bonding sites increases<br />

with diameter, the diamond rods become<br />

stronger than SWNTs. The corresponding critical<br />

diameters for the three orientations of diamond<br />

nanorods are summarized in Table 13. While the<br />

strength of SWNT increases linearly with diameter,<br />

the strength of nanorods grows as the square<br />

of diameter. Figure 50 also illustrates that the<br />

strength of a MWNT is comparable to that of a<br />

294


diamond rod oriented in the direction and<br />

exceeds those for and -oriented<br />

nanorods. Although the load-bearing properties of<br />

MWNTs are quite impressive, in practical applications,<br />

special precautions must be taken to exploit<br />

this property. Experiments by Yu et al. 249 have<br />

shown that when clamps of a particular kind are<br />

applied to the outermost shell of a tensile loaded<br />

MWNT, failure occurs only in the outermost shell.<br />

This result suggests little load transfer between the<br />

outer shell and inner shells, and therefore the inner<br />

shells do not contribute to the load bearing crosssectional<br />

area of the system. Because of the integrity<br />

of diamond rods, this problem is absent from<br />

the structures. It was also shown 196 that by no<br />

means do carbon nanotubes posses the highest<br />

strength-to-weight ratio.<br />

To futhur explore whether diamond nanorods<br />

represent an important and viable target structure<br />

for synthesis, molecular modeling has been used<br />

to explore structures and to characterize binding<br />

energies for several example diamond nanorod<br />

configurations illustrated in Figure 51a. 196 Plotted<br />

in Figure 51b are energies as a function of the<br />

hydrogen-to-carbon ratio for the four diamond<br />

nanorods illustrated in Figure 51a, calculated with<br />

the bond-order analytic potential. 218 Also plotted<br />

are energies for (17,0) nanotubes of finite length<br />

whose ends have been hydrogen terminated. The<br />

energies reported are relative to systems with the<br />

same number of carbon and hydrogen atoms in<br />

graphite and as H 2 molecules, respectively. For<br />

large carbon-to-hydrogen ratios, the diamond<br />

nanorods and the single-walled fullerenes are<br />

roughly comparable in binding energy, while for<br />

small ratios the sp 3 -bonded structures are energetically<br />

favored, in agreement with a previous<br />

analysis of carbon-hydrogen clusters (Section II).<br />

295


To explore the size dependence of the<br />

Young’s modulus of diamond nanorods, the<br />

second derivative of strain energy in the system<br />

was been calculated 196 as a function of strain<br />

for a nanorod of 1.4 nm in diameter (Table 13).<br />

This value is an unambiguious characteristic<br />

and can easily be converted to Young’s modulus<br />

(see Section III.B.2 below). While an average<br />

value of the second derivative of strain<br />

energy is similar to that of bulk diamond, the<br />

strain energy changed in a different manner for<br />

subsurface carbon atoms and those in the rod<br />

center. The ‘strengthening’ of the atoms in the<br />

rod center (Table 13) had been observed Ref.<br />

196.<br />

Atomistic simulations 196 indicate that diamond<br />

nanorods represent an important and viable target<br />

structure for synthesis. Given the number of methods<br />

that have been used to produce covalently<br />

bonded whiskers, the prospects for synthesizing<br />

diamond nanorods appears to be very promising.<br />

For example, impressive results have been<br />

achieved in the growth of β-SiC nanorods and<br />

whiskers with radii as small as 10 nm. Synthesis<br />

has been accomplished by a number of methods,<br />

including hot filament chemical vapor deposition,<br />

219 laser ablation, 220 reduction-carburization, 226<br />

and sol-gel reactions followed by carbothermal<br />

reduction of xerogels. 267 However, as described<br />

in Section III.A.1, aligned diamond whiskers have<br />

296


FIGURE 50. Fracture force as a function of diameter for different one-dimensional carbon<br />

structural components: diamond nanorods with different crystallographic orientations, SWNTs<br />

and MWNTs. The types of nanotubes correspond to those with a zigzag arrangement of<br />

atoms. In the case of MWNTs, the outer diameter is a variable and the inner diameter is<br />

assumed 4 nm for all MWNTs. (After Ref. 196.)<br />

FIGURE 51. Illustrations of representative diamond nanorods (a). Binding energies given by a bond-order potential<br />

as a function of hydrogen-to-carbon ratio (b). Energies were calculated using a dimer reconstruction for all surfaces<br />

with structures corresponding to the (100) diamond surface. Open triangles: diamond nanorods. Solid circles: (17,0)<br />

nanotubes of different lengths whose ends are hydrogen terminated. (After Ref. 196.)<br />

297


een formed so far only by a ‘top down’ approach<br />

using air plasma etching of polycrystalline diamond<br />

films, particularly of as-grown diamond<br />

films and films with molybdenum deposited as an<br />

etch-resistant mask. 33 In addition to the ‘top-down’<br />

methods of diamond nanorod synthesis, direct<br />

conversion of carbon nanotubes to diamond<br />

nanorods might be also considered. It is known<br />

that carbon onions can be converted to structures<br />

containing nanodiamond cores under MeV e-beam<br />

irradiation 93 or ion beam irradiation. 123 So far,<br />

electron beam induced formation of amorphous<br />

carbon nanorods 10 to 20 nm in diameter from<br />

CVD-deposited carbon nanotubes has been realized<br />

in situ under HRSEM. 268 Similarly, it has<br />

been observed that carbon nanotubes can be transformed<br />

into amorphous carbon rods under irradiation<br />

with an Ar ion beam. 269<br />

Interestingly, diamond rods with diameters as<br />

small as 10 µm and several hundred micrometers<br />

long were fabricated by a variety of techniques<br />

back in the 1960s. 270–272 The monocrystalline rods<br />

were grown epitaxially on diamond seed crystals<br />

from a gaseous phase under low pressure conditions<br />

in the presence of a Ni, Fe, or Mn catalyst. 270<br />

Diamond whiskers were also grown in an electron<br />

microscope on the sharp edges of diamond or<br />

other dielectric crystals under electron beam irradiation<br />

from carbon-containing residual gases at<br />

low pressure. 271 Diamond whisker growth in a<br />

metal-carbon system at high pressures and temperatures<br />

conditions was also reported. 272<br />

B. <strong>Carbon</strong> Nanotubes<br />

1. Synthesis and Properties<br />

Synthesis and properties of nanotubes are<br />

topics comprehensively addressed in a variety of<br />

books. 39,199–203 Below we provide a brief summary<br />

of these topics.<br />

There are three most commonly used methods<br />

to produce carbon nanotubes: arc discharge,<br />

laser ablation, and chemical vapor deposition.<br />

There are also several nontraditional methods that<br />

have been developed (discussed, for example, in<br />

Ref. 39, 224). CVD methods have been used for<br />

at least 4 decades for the production of filamentous<br />

carbon; 203 this method is based on the decomposition<br />

of carbon-containing gases on metal<br />

catalysts at reaction temperatures below 1000 o C.<br />

Arc discharge and laser ablation methods are based<br />

on the condensation of carbon atoms generated<br />

from the evaporation of solid carbon sources. The<br />

evaporation temperature involved in these processes<br />

is close to the melting temperature of graphite,<br />

3000 to 4000 o C. The multiwall nanotubes<br />

reported by Iijima 13 in 1991 were produced by an<br />

arc-discharge method. Before the discovery of<br />

nanotubes, both arc discharge and laser ablation<br />

techniques had been used to produce fullerenes,<br />

but the optimal conditions to produce fullerenes<br />

is different that the optimal conditions required to<br />

produce nanotubes. In 1985 Smalley and colleagues<br />

12 reported that they had produced<br />

fullerenes by a laser vaporization method. The<br />

chronology of the major events in the synthesis of<br />

nanotubes and fullerenes is summarized in Table<br />

14.<br />

In an arc-discharge, carbon atoms are evaporated<br />

by a helium plasma generated by high a<br />

current between the anode and cathode. This<br />

method can produce both high-quality MWNTs<br />

and SWNTs (Figure 52). Growth of MWNTs by<br />

this method does not require, in principle, a catalyst.<br />

39,204 MWNTs can be obtained by controlling<br />

the growth conditions such as the pressure of the<br />

inert gas (optimal for MWNT ~500 Torr, for C 60 –<br />

below 100Torr 39 ) and the arc current. The synthesized<br />

MWNTs typically have a length on the<br />

order of 10 µm and diameters in the range of 5 to<br />

10 nm. They typically form tight bundles, but<br />

nanotubes themselves are very straight. The purification<br />

of MWNTs can be achieved by heating<br />

as-grown material in an oxygen environment.<br />

While a standard arc-evaporation method produces<br />

only multilayered nanotubes, the addition<br />

of metals such as cobalt to the graphite electrodes<br />

results in extremely fine nanotubes with singlelayer<br />

walls. 47,48 The optimization of SWNT growth<br />

was achieved using a carbon anode containing<br />

yttrium and nickel as a catalyst. 204<br />

An alternative method of growth of highquality<br />

SWNTs in large quantities was suggested<br />

in 1996. 49 Like the original method of preparing<br />

C 60 , this involved the laser vaporization of a graphite<br />

target containing a small amount of catalyst<br />

298


and resulted in a high yield of single-walled<br />

nanotubes with unusually uniform diameters.<br />

These highly uniform nanotubes had a greater<br />

tendency to form aligned ropes than those prepared<br />

using arc evaporation. The ropes consist of<br />

tens of individual nanotubes close-packed into<br />

hexagonal crystals. The initial experiments indicated<br />

that the rope samples contained a very high<br />

proportion of nanotubes with a specific armchair<br />

structure, but later other types of SWNT were<br />

grown. SWNTs grown both by laser ablation and<br />

arc discharge typically contain by products of<br />

fullerenes, graphitic polyhedrons with encapsulated<br />

metal particles, and amorphous carbon particles<br />

or overcoatings on SWNT walls. A widely<br />

used purification method developed by Smalley<br />

and co-workers 207 involves refluxing the as-grown<br />

SWNTs in a nitric acid solution for an extended<br />

period of time.<br />

The CVD growth process involves heating a<br />

catalyst material to high temperatures in a furnace<br />

and flowing a hydrocarbon gas through the<br />

reactor (Figure 52). The key parameters of the<br />

CVD nanotube growth are the hydrocarbons,<br />

catalysts, and growth temperature. The catalyst<br />

species are typically transition-metal<br />

nanoparticles formed on a support substrate such<br />

as alumina. For MWNT growth, mostly ethylene<br />

or acetylene is used as a hydrocarbon feedstock,<br />

and the growth temperature is typically in the<br />

range of 550 to 750 o C. 204 There are the same<br />

typical catalytic metals (iron, cobalt, nickel) for<br />

the CVD process, as for laser ablation and arc<br />

discharge methods. The drawback of the CVD<br />

growth of MWNT is the high defect density, a<br />

problem that still needs to be addressed. 204 Recent<br />

progress in the fabrication of MWNT using<br />

the catalytic CVD method shows that narrow<br />

diameter double and triple walled nanotubes can<br />

be grown with high yield. 210 Growth of bulk<br />

amounts of SWNTs with a high degree of structural<br />

perfection was enabled by a CVD method 204<br />

using an appropriate catalyst, a temperature range<br />

of 800 to 1000 o C, and preferably methane as the<br />

carbon feedstock. Optimization of the catalyst to<br />

grow highly aligned SWNTs nanotubes is described<br />

in Ref. 204. The maximum length of<br />

individual SWNTs grown by this method is 150<br />

micrometers. 204 The SWNT produced can be<br />

separated from the support material by acidic<br />

treatment. The high interest in CVD nanotube<br />

growth is also due to the fact that aligned and<br />

299


FIGURE 52. Schematic representation of the experimental setup for the three most common<br />

methods of nanotube growth (adapted from Ref. 204). Type of nanotube grown (MWNT or<br />

SWNT), necessity to use a catalyst, as well as the possibility to grow fullerenes by the same<br />

method are also specified. Available data on nanotube yield are also provided.<br />

ordered nanotube structures can be grown on<br />

specific growth sites of prepatterned substrates<br />

with control that is not possible with the arc<br />

discharge or laser ablation techniques. 204 Thus,<br />

using the CVD method in combination with<br />

microfabrication techniques, individual SWNT<br />

can be integrated into nanotube-based electronic<br />

and chemical-mechanical devices that are functional<br />

on the molecular scale. The relatively low<br />

temperatures of the process and the ability to<br />

pattern the catalyst material directly on device<br />

substrates make catalytic CVD the method of<br />

choice for nanotube device development.<br />

To illustrate how the diameter of the nanotubes<br />

depends on the method of growth and specific<br />

growth conditions, we refer the reader to Table 15<br />

(adapted from Ref. 205). On the basis of the<br />

analysis of experimental data in the literature,<br />

summarized in Table 15, the authors 205 concluded<br />

that the majority of the catalytic mechanisms underlying<br />

the formation of various nanocarbon<br />

deposits on catalytic substrates include some com-<br />

300


mon steps. A thermodynamic analysis of carbon<br />

nucleation on the metal surface demonstrates that<br />

a variation of the reaction parameters, such as the<br />

temperature and the nature of the metal catalyst<br />

and promoters, can lead to the formation of different<br />

carbon deposits, such as filamentous carbon,<br />

multiwall nanotubes, or single-wall nanotubes.<br />

As an opposite extreme to the challenge of<br />

growth of an individual nanotube at a predetermined<br />

site, the synthesis of large uniform and<br />

ordered micro- and eventually macrostructures of<br />

SWNTs has been pursued rigorously. Recently,<br />

the self-assembly of single crystals of SWNTs<br />

using thermolysis of nanopatterned precursors has<br />

been reported. 51 Micrometer-sized crystals of<br />

SWNTs were composed of arrays of thousands of<br />

SWNTs with identical diameters and chirality.<br />

The precursor material for SWNT growth consisted<br />

of a heterostructure composed of alternate<br />

layers of buckyballs and thermally evaporated<br />

nickel that resulted in very small nucleation sites<br />

and subsequent self-assembly of the SWNT crystals.<br />

Even more recently, 32 long nanotube strands,<br />

up to several centimeters in length, consisting of<br />

aligned SWNT were synthesized by the catalytic<br />

pyrolysis of n-hexane with an enhanced vertical<br />

floating technique. The long strands of nanotubes<br />

were assembled from arrays of nanotubes, which<br />

were intrinsically long.<br />

The industrial capacity for the production of<br />

carbon nanotubes is steadily increasing. Information<br />

on leading commercial suppliers of carbon<br />

301


nanofibers and nanotubes is provided in Ref. 224.<br />

As a particular example of services provided by<br />

one of the current SWNT suppliers, we refer the<br />

reader to <strong>Carbon</strong> Nanotechnologies, Inc, (CNI)<br />

Texas. In addition to the production of SWNTs<br />

(in the CNI case, by laser ablation method), the<br />

list of services provided by the vendors includes<br />

derivatization of end-of-tube and the sidewall of<br />

nanotubes, which is important for modifying physical<br />

properties, cross-linking, and enhancing solubility<br />

and dispersion; enabling technology for enduse<br />

applications such as cutting, alignment,<br />

dispersion, solubilization, and wrapping with polymer<br />

surfactants.<br />

In general, major challenges in nanotube<br />

growth remain the ability to gain control over the<br />

nanotube chirality and diameter, and the ability to<br />

directly grow semiconducting or metallic<br />

nanotubes from and to any desired sites. 204<br />

There is a long list of unique properties of<br />

nanotubes that make them one of the most important<br />

structural and functional materials for<br />

nanotechnological applications. One of the striking<br />

features of nanotubes is their combination of<br />

widely variable electronic properties and extreme<br />

mechanical properties. These are discussed in detail<br />

in a later section. Depending on their precise<br />

structure, nanotubes can be either metallic conductors<br />

or semiconductors. The semiconducting<br />

nanotubes are of two types: one has a band gap of<br />

a few hundredths of an electron volt, while others<br />

have about a 1-eV band gap depending on the<br />

radius. Ropes have been measured with a resistivity<br />

of 10 –4 ohm-cm at 300 K, 49 making them the<br />

most conductive fibers known. Individual<br />

nanotubes have been observed to conduct electrons<br />

ballistically, for example, without scattering,<br />

with coherence lengths of several microns. 211<br />

In addition, they can carry the highest current<br />

density of any known material, measured 212 as<br />

high as 10 9 A/cm 2 . Quantum conductance of<br />

SWNTs has been reported 213 as well as their intrinsic<br />

superconductivity. 214 Selected properties<br />

of SWNTs are summarized in Table 16.<br />

2. Mechanical Properties of <strong>Carbon</strong><br />

Nanotubes<br />

The mechanics of carbon nanotubes is a topic<br />

of very intensive research as confirmed by the<br />

number of excellent reviews published within the<br />

302


last year. 221–225 The focus of much of the research<br />

has been on properties of individual NTs, such as<br />

their Young’s modulus, bending stiffness, buckling<br />

criteria, tensile and compressive strength.<br />

Advanced mechanical properties of nanotubes are<br />

important in projected applications such as<br />

nanoelectromechanical systems, nanosensors,<br />

nanocircuits, drug delivery devices, and reinforcing<br />

structures in nanocomposites. 223–225 Recently,<br />

a super-hard bulk material consisting solely of<br />

nanotubes has been created, suggesting that macroscopic<br />

structures consisting exclusively of<br />

nanotubes might be possible. 51<br />

Below we summarize experimental and theoretical<br />

data reported to date on two important mechanical<br />

characteristics of nanotubes, their Young’s<br />

modulus, and tensile strength. There is a surprisingly<br />

wide variation within the measured and predicted<br />

values of these properties. For example, there<br />

is fivefold variation in measured properties of<br />

SWNTs and a 40-fold variation in the predicted<br />

fracture strengths of SWNT. 227,228 To a large extent,<br />

this discrepancy can be attributed to the fact that<br />

definitions characterizing macroscopic material behavior<br />

under load should be used with caution in<br />

describing nanometer-scale objects. To address this<br />

important issue, the area of nanomechanics is emerging<br />

as a new discipline that describes the behavior of<br />

materials and structures at the nanoscale. 221–223 New<br />

mechanical characteristics that are different from<br />

those used in conventional continuum mechanics<br />

will likely be defined to describe the deformation<br />

behavior of nanostructures in the near future. Depending<br />

on the demands of applications taking place<br />

solely at the nanoscale (in contrast to macroscopic<br />

applications in composites, for instance), mechanical<br />

characteristics might be introduced along with<br />

materials characteristics that depend on a specimen’s<br />

geometry. This issue has been partially addressed in<br />

the Section III.A.3.c on the comparison of the fracture<br />

strengths of carbon nanotubes and hypothetical<br />

diamond nanorods and is briefly touched on in the<br />

discussion of the fracture strengths of nanotubes.<br />

a. Young Modulus of Nanotubes<br />

The definition of Young’s modulus relies on<br />

the continuum hypothesis that assumes spatial<br />

uniformity of a material and is defined as a material<br />

property (a characteristic intrinsic to the<br />

material independent on specimen geometry). As<br />

the specimen size diminishes, the discrete structure<br />

of the material can no longer be homogenized<br />

into a continuum. As has been pointed out<br />

by Yakobson et al., 221 if one considers graphene<br />

as a two-dimensional material, a MWNT should<br />

be considered as an engineered structure composed<br />

of rolled graphene sheets rather than a<br />

material. However, the definition of material-like<br />

characteristics, such as Young’s modulus, nevertheless<br />

has been applied to nanotubes when considering<br />

their application as an engineering material.<br />

Below we consider in more detail the<br />

discrepancy between Young’s modulus calculated<br />

by two means, both of which originate in continuum<br />

mechanics, that do not contradict one<br />

another if applied to macroscale materials, but<br />

give a fivefold difference in predicted Young’s<br />

modulus of nanotubes. This result has been pointed<br />

out by Yakobson et al. by applying a shell model<br />

to carbon nanotubes 235 that served an otherwise<br />

very useful role in predicting of buckling behavior<br />

of nanotubes. This is probably the most striking<br />

inconsistency in the application of continuum<br />

mechanics approaches to the nanoscale.<br />

Yakobson’s Paradox<br />

Calculating a Young’s modulus is straightforward<br />

if the potential energy functions describing<br />

interatomic interactions are known. The calculations<br />

involve the second derivative of the<br />

strain energy with respect to the applied strain:<br />

Y<br />

b =<br />

⎛<br />

2<br />

1 ∂ E⎞<br />

V<br />

⎜ 2 ⎟<br />

⎝ ∂ε<br />

⎠<br />

ε = 0<br />

(8)<br />

where V=Sh is the volume of the object, h in the<br />

object’s thickness, and S is the surface area of the<br />

nanotube at zero strain. The value of ∂ 2 E/∂ε 2 is a<br />

unique characteristic independent of the structure<br />

or material under consideration. The average value<br />

for nanotubes obtained from recent first principle<br />

simulations is 56 eV/atom and that of graphite is<br />

57.3 eV/atom (Table 17). 229 For simulations using<br />

303


many-body interatomic potentials, the value of<br />

∂ 2 E/∂ε 2 is comparable. 221 Assuming h=0.34 nm<br />

(the interplanar distance in graphite), the calculated<br />

Young’s modulus of nanotubes with Eq. 6<br />

using first principles data 229 are slightly lower<br />

than that of graphite (1030 GPa). This value is<br />

comparable to the lowest reported experimental<br />

values of SWNT extracted from data on thermal<br />

oscillations and bending tests, which are about<br />

1.2 to 1.3 TPa, 233,234 where the same assumption<br />

on the tube thickness had also been made in the<br />

corresponding expressions for the thermal vibrations<br />

amplitude and for the nanotube deflection<br />

under bending used to extract the Young’s modulus.<br />

The contradiction with the values reported<br />

above appears when the Young’s modulus of<br />

SWNT and wall thickness are estimated from the<br />

in-plane stiffness and flexural rigidity calculated<br />

in the shell model, 235 where a SWNT is approximated<br />

by a continuum isotropic shell. The deformation<br />

energy of a nanotube is a function of inplane<br />

strain ε and the changes of curvature k in the<br />

axial and circumferential directions of a shell.<br />

Using the shell model, a wide variety of shell<br />

buckling behavior, particularly critical strains or<br />

curvatures vs. applied load such as compression,<br />

bending, torsion can be predicted. Molecular dynamics<br />

simulations of transformations of shapes<br />

of SWNTs subject to large nonlinear deformations<br />

demonstrated remarkable agreement with<br />

tube buckling behavior predicted from the shell<br />

model. 235 The only material parameters required<br />

in the shell model are in-plane stiffness C,<br />

Poisson’s ratio ν, and the flexural rigidity D that<br />

can be obtained from first principle simulations.<br />

The simulation of the axial tension of a nanotube<br />

provides values for Poisson’s ratio ν and the inplane<br />

stiffness C<br />

⎛<br />

2<br />

1 ∂ E⎞<br />

C = S<br />

⎜ 2 ⎟<br />

⎝ ∂ε<br />

⎠<br />

ε = 0<br />

(9a)<br />

The flexural rigidity, the dependence of the strain<br />

energy on its curvature, D can be calculated from<br />

304


a comparison of strain energy U of tubes of different<br />

diameters d:<br />

D=<br />

Ud<br />

2<br />

/ 2<br />

(9b)<br />

Recent first principles-based estimations 229 provide<br />

D=3.9 eVÅ 2 /atom and show no dependency<br />

on tube chirality. The in-plain stiffness for different<br />

types of nanotubes are given in Table 17.<br />

As we can see so far, the nanotube thickness is<br />

not involved in the estimations of C and D in the<br />

shell model. On the other hand, a continuum<br />

shell can be assigned the modulus Y sh and the<br />

thickness h unambiguously, to formally match<br />

the ν, C and D:<br />

C= Yshh, (10a)<br />

3<br />

Yshh<br />

D =<br />

2<br />

12( 1−ν<br />

)<br />

(10b)<br />

Using the first principles parameters reported<br />

above for C and D and a first principle’s Poisson<br />

ratio ν=0.15, 229 the shell Young modulus would<br />

be 3.86 TPa (~5 TPa with an analytic potential for<br />

describing interatomic interactions) and with a<br />

shell thickness h=0.9 Å. These values differ significantly<br />

from those reported above. Based on<br />

this discussion, these questions remain until we<br />

can apply a material definition to an atomically<br />

thick structure, as well as on limitation of the<br />

applicability of Eq. 10 to nanostructures.<br />

The in-plane stiffness C is an unambiguous<br />

parameter, and considering only axial tension,<br />

according to Eq. 8a and h can take any values<br />

satisfying . If we assume h=0.34 nm, the resulting<br />

Young’s modulus of a SWNT will be of the order<br />

of ~1 TPa using the first principles in-plane stiffness<br />

value. 229 It should be noted that for nanotubes<br />

with diameters less than 0.68 nm (doubled<br />

interplanar space in graphite), using h=0.34 nm<br />

makes no sense, although in many articles this<br />

value is used for estimations of Y for nanotubes<br />

of any diameter. Nanotubes with small radii in<br />

principle should be considered as rods with corresponding<br />

cross-sectional areas. 230<br />

For SWNT ropes, which are a three-dimensional<br />

material and therefore the classic definition<br />

of Y is valid, the Young’s modulus would unambiguously<br />

be recovered as:<br />

2<br />

ρb<br />

∂ E<br />

Y =<br />

2<br />

(11)<br />

m ∂ε<br />

where ρ b is the bulk density and m is the mass of<br />

the atom. First principles Y values for SWNT<br />

ropes are of the order of ~1 TPa (Table 18).<br />

Thus, there is an interesting situation in the<br />

application of the continuum shell model to tubular<br />

nanostructures. On the one hand, the description<br />

of nanotube characteristics within the shell<br />

model is intrinsically self-consistent and is very<br />

useful for predicting tube buckling behavior. On<br />

the other hand, there is remarkable discrepancy<br />

between the Young’s modulus of SWNT recovered<br />

from the shell model and that corresponding<br />

to SWNT ropes or graphite. To avoid this problem,<br />

Ru 236 proposed an intrinsic bending stiffness<br />

for carbon NTs in order to decouple the bending<br />

shell stiffness of NTs from their ill-defined effective<br />

thickness and to ensure a consistent use of the<br />

classic shell theory (i.e, shell thickness — only<br />

for shell theory).<br />

At present, theoretical issues on the Young’s<br />

modulus of nanotubes can be summarized as following.<br />

221,229,236<br />

1. The elastic properties of a cage graphite<br />

structure can be characterized unambiguously<br />

only by its in-plane and flexural rigidity<br />

parameters (Eq. 9), both of which can be<br />

calculated from first principles.<br />

2. Using these rigidity parameters (and Poisson<br />

ratio), SWNT buckling behavior can be<br />

predicted very well from the macroscopic<br />

shell model, as confirmed by atomistic simulations.<br />

3. Using rigidity parameters, the shell Young’s<br />

modulus and shell thickness can be defined<br />

unambiguously (Eq. 10). Currently, the fact<br />

is that these characteristic are inconsistent<br />

with Young’s modulus and thickness of 3-D<br />

materials such as SWNT ropes and MWNTs<br />

and therefore are attributed only to the shell<br />

model for understanding its high resilience. 236<br />

4. If cage structures are distributed statistically<br />

uniformly over a large cross-sectional area<br />

305


306


(as SWNT ropes or MWNT with relatively<br />

big radius), the bulk Young’s modulus can<br />

be recovered unambiguously (Eq. 4).<br />

5. An appropriate estimation of the bulk<br />

Young’s modulus of a SWNT cannot be<br />

made using Equations 9 and 8a due to the<br />

uncertainty of nanotube bulk density or an<br />

arbitrary geometrical thickness, respectively.<br />

Nevertheless, by convention, h=0.34 nm is<br />

usually presumed so that calculated results<br />

on in-plane rigidity of SWNTs are reported<br />

in terms of Young’s modulus.<br />

6. Experiments addressing the Young’s modulus<br />

of SWNT (Table 18) use analytical expressions<br />

relating measured values and<br />

Young’s modulus. An assumption on SWNT<br />

wall thickness is also required in these estimations,<br />

and it is assumed to be 0.34 nm.<br />

Within the above summary, questions still<br />

exist such as, for example, when can a MWNT be<br />

considered a ‘bulk’ material so that the Eq. 4 can<br />

be applied. For example, Govindjee and<br />

Sackman 237 considered an elastic multisheet model<br />

to show the explicit dependence of material properties<br />

on the system size when a continuum crosssection<br />

assumption is made for a multishell system<br />

subjected to bending. The continuum<br />

assumption is shown to hold when more than 201<br />

shells are present in the macromechanical system<br />

considered.<br />

There is an additional consideration in the<br />

nanomechanics of SWNT related to the possible<br />

changing of the nature of bonding when π− and<br />

σ- bonds experience a different influence of deformation<br />

during, for example, a bending test, so<br />

that when, for example, a nanotube is bent, its<br />

effective tension-related properties are different<br />

from those in a pure tension test.<br />

Probably useful atomistic simulations can be<br />

done with sheets of diamond of different thickness.<br />

Unlike nanotubes, the thickness can be continuously<br />

increased and tendencies in in-plane<br />

stiffness and flexural rigidity can be analyzed.<br />

A different approach for analyzing the linear<br />

elastic modulus of a SWNT is described in, 239<br />

where nanoscale continuum theory is established<br />

to directly incorporate interatomic potentials into<br />

a continuum analysis without any parameter fitting.<br />

The theory links interatomic potentials and<br />

atomic structure of a material to a constitutive<br />

model on the continuum level. Reported Young’s<br />

moduli are within the range 500 to 700 GPa,<br />

depending on the type of the interatomic potential<br />

used in the calculations. These values are somewhat<br />

lower than those obtained from first principles<br />

or pure atomistic simulations with analytic<br />

potentials.<br />

Experimental and Theoretical Young’s<br />

Modulus for CNTs<br />

Currently, measured Young’s moduli of CNTs<br />

are typically of the order of 1.0 to 1.3 TPa (Table<br />

18), and those predicted from first-principle simulations<br />

are of the order of 1 TPa (Table 19).<br />

However, the maximum observed and predicted<br />

values are up to 5 Tpa, which is probably an<br />

artifact in the experimental measurements and the<br />

Young’s modulus for the shell model in the theoretical<br />

analysis as discussed above.<br />

A number of experimental methods have been<br />

used to date to deduce Y (Table 18). These methods<br />

include transmission electron microscopy<br />

(TEM) observations of thermally excited vibrations,<br />

240 thermal stresses monitored by micro-<br />

Raman spectroscopy, 241 atomic force microscopy<br />

(AFM)-based measurements of lateral deflection<br />

of CNTs, 242 and a direct pulling technique by<br />

applying an axial force to a long NT rope 23,24 or to<br />

a single MWNT 25 (Figure 53).<br />

According to Table 18, the TEM-based methods<br />

suffer from relatively large margins of error<br />

margin of the order of 100%. For AFM-based<br />

techniques, the error margin can be of the order of<br />

50%. Significant errors mostly arise from difficulties<br />

in anchoring the ends of the nanotube. 246<br />

In addition, to extract a Young’s modulus most<br />

experimental techniques make some assumptions<br />

regarding the mechanical behavior of the system<br />

and assume that continuum elasticity theory is<br />

valid so that well-established expressions from<br />

elasticity theory for loaded tubes can be used at<br />

the nanoscale. However, the resulting values can<br />

be very sensitive to geometrical assumptions. For<br />

example, measured values of Y of Mo-based nanowhiskers<br />

varied by about a factor of two depending<br />

on the assumed geometry of the cross-sec-<br />

307


308


FIGURE 53. Illustration of a multiwall carbon nanotube (~11 nm diameter) prior to<br />

tensile testing (a). Arrows indicate direction of loading (white is the actual pulling<br />

direction); (b) fracture surfaces of both ends of the tube after breakage. (Reprinted from<br />

Ref. 245, with permission from Elsevier Science.)<br />

tional area of the nanocrystal. 246 An excellent review<br />

of instrumentation for measuring Young’s<br />

moduli is provided in Ref. 223.<br />

In contrast to the experimental results, there<br />

are no appreciable differences between the<br />

Young’s modulus for nanotubes and for a single<br />

graphene sheet predicted from first principles<br />

simulations (Table 19). Also, the results from<br />

Table 19 indicate that the effect of curvature and<br />

chirality on the elastic properties of NTs is small.<br />

More detailed analysis of variations in the strength<br />

of C-C bonding as the graphene sheet is rolled<br />

into a tube had been performed in Ref. 227. Based<br />

on a first principles cluster method, the authors<br />

adopted the Mulliken population analysis to estimate<br />

the overlap integrals between carbon atoms<br />

that measure the strength of the covalent bond.<br />

According to these results, the binding energy,<br />

elastic modulus, and tensile strength of carbon<br />

nanotubes must be less than that of graphite. Indeed,<br />

results of first principles calculations 232 and 229<br />

indicate an increase in the in-plane rigidity of<br />

nanotubes (approaching that of graphene) with<br />

increasing radius (Table 17). The Poisson’s ratio<br />

also retains graphitic values except for a possible<br />

slight reduction for small radii. It shows, however,<br />

chirality dependence (Tables 17 and 19).<br />

Limitations of Continuum Mechanics at<br />

the Nanoscale<br />

Atomistic simulations of buckling of selected<br />

nanotubes 235 demonstrated that bifurcation-buckling<br />

equations from shell theory can be applied, in<br />

principle, at the nanoscale. However, the deformation<br />

behavior of every nanotube cannot be satisfactorily<br />

described using a shell model. A comprehensive<br />

analysis of the limitations on the<br />

applicability of continuum shell models to the<br />

entire class of nanotubes has been performed by<br />

Harik. 230,248<br />

A scaling analysis has been used to identify<br />

the key nondimensional geometrical parameters<br />

that control the buckling behavior of NTs. Based<br />

309


on the relationship between buckling behavior<br />

and NT structure, four broad classes of CNTs had<br />

been identified (Figure 54a): thin and thick shells,<br />

high aspect ratio (long) NTs, and NT beams. A<br />

representative applicability map according to the<br />

NT classes had been also constructed (Figure 54b).<br />

The descriptive name of each class indicates the<br />

structural properties of NTs as well as the potential<br />

continuum models that can be used to predict<br />

their global mechanical behavior. Thus, NT shells<br />

behave either as thin shells or thick shells (hollow<br />

cylinders) depending on the thickness-to-radius<br />

ratio. The long NTs (class II) have structural behavior<br />

similar to the behavior of columns. The<br />

NT beams (class III) deform like macroscopic<br />

beams.<br />

The results of the analysis have important<br />

practical applications. For NTs with the same<br />

FIGURE 54. Schematic representation of four classes of nanotube geometries (a) used in the applicability<br />

map for the continuum beam model (b). Ranges of values for non-dimensional geometric parameters<br />

of nanotubes define NT classes and indicate the limits in the applicability of the thin-shell model<br />

for NTs. Parameters L NT , R NT , and a 1 represent the three length scales involved in the nano-mechanical<br />

problem. a 1 is the width of the hexagonal carbon rings, which is about 0.24 nm. L NT, R NT, d NT are nanotube<br />

length, radius and diameter, respectively. (Reprinted from Ref. 230 with permission from V. Harik.)<br />

310


values of the nondimensional parameters, they<br />

must have identical critical strain and buckling<br />

modes, even if the individual structural features<br />

are different. This can help to reduce the number<br />

of MD simulations needed to describe a whole<br />

class of NTs. In experimental studies, specific<br />

equivalent-continuum models should be used for<br />

a system under investigation according to the<br />

applicability map for data reduction or NT probe<br />

design.<br />

b. Nanotube Tensile Strength<br />

CNT tensile strengths predicted using first<br />

principles calculations and obtained experimentally<br />

are summarized in Tables 19 and 20, respectively.<br />

Measurements were done on individual<br />

MWNTs and SWNT ropes. Yu et al. obtained the<br />

tensile strength of individual MWNTs by scanning<br />

electron microscopy. 249 The MWNTs broke<br />

in the outermost layer, and the tensile strength of<br />

the MWNT layer ranged from 11 to 63 GPa.<br />

Analysis of the stress-strain curves for the individual<br />

MWNTs indicated that the corresponding<br />

Young’s modulus of the outer layers varied from<br />

270 to 950 GPa. More recently, the first observation<br />

of the actual breaking of a MWNT in tension<br />

in TEM has been reported 245 (Figure 53). Figure<br />

53b reveals that the nanotube has apparently<br />

‘healed’ itself after failure by forming a closed<br />

311


end-cap. Based on the observation of the force<br />

required to break the nanotube a tensile strength<br />

of 150 GPa (experimental uncertainties ~30%)<br />

was estimated. The corresponding deformation<br />

was ~5%. This is the highest reported fracture<br />

strength of any material and is relatively close to<br />

the ideal fracture strength of graphene. From corresponding<br />

bending studies on such nanotubes,<br />

the Young’s modulus was estimated to be 0.9 TPa<br />

(±20%). Ab initio estimations of ideal fracture<br />

strengths along the directions corresponding to<br />

the zigzag and armchair atoms arrangements at<br />

the edge are 209 GPa and 226 GPa, respectively. 228<br />

This ideal tensile strength had been obtained for<br />

an ideal situation where there is no symmetry<br />

breaking (no defects, no stress inhomoginities<br />

due to temperature, for example), and hence it<br />

establishes an upper limit on the strength. In similar<br />

simulation conditions, ideal fracture strengths<br />

for diamond range from ~95 GPa in the <br />

direction 198,228 to 225 GPa in the direction.<br />

198 It would be valuable to study the failure of<br />

deformed graphene in conditions of small perturbations<br />

using the first principles approaches.<br />

The only reported first principles-based estimation<br />

of a (6,6) nanotube tensile strength in<br />

elongated nanotubes is surprisingly very low (~6<br />

GPa). 227 1 ⎛ ∂Etot<br />

⎞<br />

It was calculated as σ TH = ⎜ ⎟ ,<br />

s ⎝ ∂ε<br />

⎠<br />

ε=<br />

ε<br />

where s is the surface area of cross-section, and<br />

εCR is maximum strain in the system (~0.4). The<br />

corresponding Young’s modulus is also lower<br />

(0.76 TPa) than those reported in other studies<br />

(Table 19). Thus alternative first principles calculations<br />

are required. Recently, the ductile behavior<br />

of carbon nanotubes was investigated by largescale<br />

quantum calculations. 256 While the formation<br />

energy of strain-induced topological defects determines<br />

the thermodynamic limits of the elastic<br />

response and of the mechanical resistance to applied<br />

tension, it was found that the activation<br />

barriers for the formation of such defects are much<br />

larger than estimated previously. Unfortunately,<br />

ultimate tensile stress values had not been evaluated<br />

in Ref. 256. A comprehensive analysis of<br />

nanotube ductile behavior is provided in the review,<br />

221 but recent results 256 should be taken in<br />

the account.<br />

CR<br />

Molecular dynamics simulations of nanotubes<br />

under tension had been performed using a bond<br />

order analytic potential in Ref. 261 However,<br />

while using the bond order potential for estimating<br />

elastic properties for moderate deformation<br />

(up to 15%) is reasonable, the bond-order potential<br />

used in the simulations 261 is not appropriate<br />

for simulating system behavior at deformations<br />

above approximately 15% due to acut-off via a<br />

switching function of the interatomic potential.<br />

When bond lengths exceed 1.7 Å (the beginning<br />

of the cut-off range), a steep nonphysical increase<br />

in the interatomic forces results in very high values<br />

of ultimate tensile strength. When developing<br />

an interatomic potential suitable for fracture simulations,<br />

it is important that any cut-off function be<br />

used at distances beyond the inflection point in<br />

the potential function. The inflection point relates<br />

to the peak interatomic force and therefore influences<br />

fracture stresses in a system. In our previous<br />

work on diamond cleavage, we avoided this<br />

problem by shifting the cut-off function beyond<br />

the inflection point and keeping a list of nearest<br />

neighbors during the run. 42 Recently, 43 a modified<br />

Morse function was developed to specifically<br />

address fracture of CNTs by adjusting the position<br />

of the inflection point to provide an ultimate<br />

fracture strain of nanotubes that reproduces the<br />

value obtained from the Yu et al. experiment<br />

(~10%). 249 The corresponding fracture stress under<br />

conditions of symmetry breaking (temperature,<br />

presence of vacancies) was estimated to be<br />

~90 GPa. However, a reliable value can probably<br />

only be obtained with quantum mechanics–based<br />

dynamic simulations.<br />

Interesting results on the tensile strengths of<br />

SWNTs under hydrostatic pressure have been<br />

reported in Ref. 44. Simulations used the bondorder<br />

potential with DFT simulations of selected<br />

configurations. Hydrostatic pressure on the<br />

nanotube walls was simulated by encapsulating<br />

hydrogen molecules to capped nanotubes. The<br />

reported failure strength under hydrostatic pressure<br />

was 63 GPa for a (5,5) nanotube, and that for<br />

a nanotube containing twinned 7-5-5-7 pairs was<br />

only 3 GPa lower.<br />

Finally, regarding the practical aspects of<br />

nanotube strength, the first kinetic approach to<br />

CNT real-time strength has been discussed in<br />

312


Ref. 45 predicting the yield strain range. It was<br />

found that the value depends on the NT chirality,<br />

in a way very different from the thermodynamic<br />

assessments.<br />

Estimating a three-dimensional ideal strength<br />

for nanotubes suffers from the same problem of<br />

uncertainty in defining the cross-sectional area of<br />

a nanotube as in the estimation of Young’s modulus.<br />

According to Roundy, 266 depending on the<br />

application a more useful metric might be force<br />

divided by the mass per unit length (i.e., the stress<br />

divided by mass density) (Figure 55). This would<br />

assume a strength-to-weight ratio characteristic<br />

of the structural material and can provide a single<br />

nanotube with an ideal strength that would be<br />

independent of the spatial arrangement of the tubes<br />

(e.g., MWNT vs. NT ropes). It would also be<br />

convenient to perform a comparison of the characteristic<br />

for nano-specimens from different materials.<br />

For example, the ratio between the strengthto-weight<br />

characteristic for graphite and diamond<br />

(using data from Tables 12 and 21) would be 1.56<br />

and 3.7 for and diamond orientations,<br />

correspondingly. The higher the value, the<br />

more weak/heavy is the nano-specimen.<br />

3. Assemblies of <strong>Carbon</strong> Nanotubes and<br />

Nanodiamond<br />

Certain types of nanocarbons such as<br />

fullerenes, nanotubes, and diamond clusters can<br />

serve as basic building blocks for constructing<br />

more complicated structures that might exhibit<br />

new properties and find novel applications. In this<br />

section the feasibility of assembling carbon<br />

nanotubes and nanodiamond clusters is discussed.<br />

There is a geometrical similarity between the<br />

{111} planes of diamond and individual graphene<br />

sheets in which pairs of diamond planes resemble<br />

“puckered” graphite. This relationship, together<br />

with the lengthening of carbon-carbon bonds from<br />

1.42 Å in graphite to 1.54 Å in diamond, results<br />

in the near epitaxial relations<br />

FIGURE 55. Illustration as a convenient strength characteristic of a nanostructure can be defined. It<br />

would be the maximum force before failure divided by the mass per unit length. This definition does<br />

not involve the problematic definition of cross-sectional area and is independent of the dimensionality<br />

of the nanostructure.<br />

313


Graphite (0001) || diamond (111)<br />

Graphite [ 1120 ] || diamond [ 11 0]<br />

between graphite and diamond planes. These relations<br />

have been utilized in several models for<br />

understanding important structures and processes<br />

in carbon materials, including diamond nucleation<br />

on graphite 275 and graphitization of diamond<br />

surfaces and clusters. 82,189,276,277 There are also similar<br />

but less obvious relations between fullerene<br />

nanotubes and diamond that lead to a number of<br />

mechanically and chemically stable structures with<br />

potential structural materials and nanoelectronic<br />

device applications.<br />

Several recent experiments have provided<br />

compelling evidence for stable hybrid carbon<br />

nanotube-diamond structures. In experiments by<br />

Kuznetsov et al., 82,277 the formation of nanometric<br />

closed curved graphitic structures with tubular or<br />

conical forms attached to the surface of a diamond<br />

particle were observed in HRTEM images<br />

of diamond particles after high-temperature annealing.<br />

The TEM images show concentric graphitic<br />

shells corresponding to the top view of<br />

nested carbon nanotubes (Figure 56a) as well as a<br />

side view of nanotubes on the edges of particles<br />

(Figure 56b). The authors suggest that multilayered<br />

graphitic caps that form during the initial<br />

annealing of micron-sized diamond particles at<br />

1800 to 2000 K transform into closed carbon<br />

nanotubes attached to the diamond surface via<br />

reconstruction of the edges of diamond (111)<br />

planes orthogonal to the surface into graphite<br />

(0001) planes. 277<br />

FIGURE 56. TEM images of carbon nanotubes/nanofolds attached to a diamond surface. Top<br />

view (a) and side view (b). (Reprinted from Ref. 277, with permission from Elsevier Science.)<br />

314


Guidelines for creating a range of robust<br />

nanotube-diamond structures have been developed<br />

recently 278 and tested with the bond-order<br />

potential. 218 The analysis was restricted to (n,0)<br />

and (n,n) nanotubes perpendicularly attached to<br />

diamond (001) and (111) facets. These rules were<br />

derived based on geometrical considerations in<br />

combination with energies and stresses estimated<br />

from atomic modeling using an analytic potential<br />

function. 278 Geometrical considerations include<br />

analysis of the total and local mismatches between<br />

a nanotube and the diamond surface. A<br />

total mismatch is a mismatch between the perimeter<br />

of a nanotube and the perimeter of a polygon<br />

on a diamond surface formed by atoms participating<br />

in bonding (Figure 57). For example, the total<br />

mismatch for a (6,0) nanotube is 6.7%, while that<br />

for (24,0) nanotube would be only 2.6%. To create<br />

a strong chemical interface between a (n,0)<br />

nanotube and the corresponding n-sided polygon<br />

formed by dangling bonds on a diamond surface,<br />

the shape of the polygon should be as close to<br />

circular as possible to match that of a nanotube<br />

edge. Thus, in addition to total mismatch, a local<br />

mismatch can be defined as the distance between<br />

a single vertex of a polygon on a diamond surface<br />

and the point of the projection of a corresponding<br />

atom from a nanotube edge. Atomic level simulations<br />

have suggested that local lateral mismatches<br />

as large as about 1 Å do not necessarily inhibit<br />

FIGURE 57. Schemes of the possible connection between the (111) diamond surface and zigzag nanotubes. Dots<br />

connected by solid lines correspond to the atomic sites available for bonding at nanotube edges. Crosses<br />

correspond to atomic sites at the (111) diamond surface. Dashed lines connect sites on the diamond surface<br />

participating in the bonding with the specific nanotube. Stars at the contours of nanotubes (e,f) denote the dangling<br />

bonds at the interface diamond/nanotube after the nanotube attachment. (Reprinted from Ref. 278.)<br />

315


strong bond formation. For (6,0) and (24,0)<br />

nanotubes this local mismatch would be 0.17 Å<br />

and 0.68 Å, respectively. Nanotubes with larger<br />

radii posses less total mismatch but higher local<br />

mismatch, thus the interface energy of the relaxed<br />

structures is higher for nanotubes with larger radii<br />

(Table 22).<br />

Based on criteria related to the total and local<br />

mismatches and relative symmetry of a nanotube<br />

and site available for bonding on a diamond surface,<br />

six distinct groups of (n,0) nanotubes with<br />

different degrees of bond formation with a diamond<br />

(111) facet can be identified depending on<br />

the nanotube parameter n (Figure 56). 278 The strongest<br />

interfaces are formed by nanotubes from the<br />

first two groups that correspond to nanotubes with<br />

sixfold (6 × M,0) and threefold (6 × M+3,0) symmetry,<br />

where M is an integer. The energetic characteristics<br />

of nanotubes of different groups are<br />

summarized in Table 22.<br />

The symmetry of (n,n) nanotubes do not<br />

match well with that of a defect-free diamond<br />

(111) facet, and therefore interfaces of this type<br />

will not in general show strong bonding. However,<br />

there is a strong similarity between the<br />

five-fold symmetric (111) facets of adiamond<br />

pentaparticle and the ends of (5 × M, 5 × M)<br />

nanotubes that can result in strong bonding. 278<br />

The total mismatch between a pentaparticle of a<br />

nanoscopic size and (5,5) nanotube (Figure 58a)<br />

is about –4.1%. Simulations predict that the local<br />

mismatch can be accommodated for (5 × M,<br />

5 × M) nanotubes at least up to M=3. As discussed<br />

in Section III.A, pentaparticles are often<br />

observed among nanodiamond clusters.<br />

Because of the fourfold symmetry of (100)<br />

diamond facets, a general scheme like that outlined<br />

above for the attachment of nanotubes to<br />

(111) facets could not be developed. Analysis of<br />

bonding geometries together with molecular modeling<br />

studies, however, has been able to identify<br />

specific cases of strong bonding. 278 One such example<br />

is a (12,0) nanotube attached to the (100)<br />

facet of a diamond cluster, which is illustrated in<br />

Figure 58b. The total mismatch between the surface<br />

sites and the nanotube is only 4.9%. Interface<br />

energies for selected zigzag nanotubes attached to<br />

an (001) diamond surface are provided in Table<br />

22.<br />

Depending on a nanotubes morphology, some<br />

types of open nanotubes can be chemically connected<br />

with different diamond surfaces, atom-toatom.<br />

Structures without dangling bonds at the<br />

interface can be formed; this is important for<br />

nanoelectronic applications, because dangling<br />

bonds at the interface can trap electrons and suppress<br />

conductivity through the interface. By combining<br />

metallic or semiconducting nanotubes with<br />

diamond clusters or substrates, different types of<br />

heterojunctions can be designed for carbon-based<br />

nanoelectronics applications. For example, a hybrid<br />

structure consisting of a short nanotube sandwiched<br />

between two diamond clusters mimics the<br />

double barrier structure of a resonant tunneling<br />

diode.<br />

The nanotube/nanodiamond composite, illustrated<br />

in Figure 58, may, in principle, serve as tips<br />

for a field emitting array. 172,191 According to selfconsistent<br />

tight binding simulations, 191 the work<br />

functions of a single closed nanotube and a single<br />

316


FIGURE 58. Illustration of the relaxed hybrid interface structure between a diamond pentaparticle and a (5,5)<br />

nanotube (a) and diamond truncated octahedron cluster and the (12,0) nanotube (b). The nanotube is attached to<br />

the (100) facet of the cluster. (Reprinted from Ref. 278.)<br />

317


nanotube with an edge terminated by hydrogen are<br />

about 4.5 to 5 eV. The barrier at the back contact<br />

of the diamond cluster/metallic nanotube varies<br />

from 4.5 to 2-3 eV, depending if the nanodiamond<br />

cluster consists of pure diamond phase (Figure 58)<br />

or contains the tetrahedrally coordinated (ta-C)<br />

carbon phase. 191 Taking into account the negative<br />

electron affinity at the hydrogenated surface of a<br />

diamond cluster, the resulting emission barrier for<br />

the hybrid structure is comparable or lower than<br />

that for a single nanotube (Figure 59). Calculations<br />

also indicate the presence of the wave functions<br />

with energies near the Fermi level that are continuously<br />

extended along both the nanotube and diamond<br />

cluster. 191 This enhances the possibility of<br />

current flow from a nanotube to a diamond cluster.<br />

Probably, the major advantage of the above<br />

design of a nanotube capped with a nanodiamond<br />

particle is the potential reduction of nanotube<br />

erosion resulting in increased device lifetime.<br />

Indeed, the current density for emission from a<br />

single nanotube is high in compared with a<br />

nanotube capped with a nanodiamond particle<br />

that posses more emission sites. Based on estimates<br />

using the bond order potential, an hydrogenated<br />

diamond surface is more mechanically stable<br />

than a capped or hydrogenated nanotube edge.<br />

Recently, 279 partitioned real-space density functional<br />

calculations of field evaporation of carbon<br />

clusters from SWNT were performed. The calculations<br />

demonstrate that the activation-energy barrier<br />

for field evaporation of carbon clusters and<br />

hydrocarbon clusters from SWNT’s decreases as<br />

the electric field increases. It was also found that<br />

evaporation from an open-ended carbon nanotubes<br />

occurs more easily than from capped nanotubes.<br />

Another interesting result is that adsorbed hydrogen<br />

weakens the C-C bonds and significantly reduces<br />

the activation-energy barrier for field evaporation.<br />

Regarding nanotechnology applications, a<br />

nanotube forming strong chemical bonds with a<br />

diamond cantilever can be a good candidate for a<br />

proximal probe tip.<br />

The nanodiamond/nanotube hybrid structures<br />

discussed above can, in principle, be synthesized<br />

by manipulating diamond clusters with a proximal<br />

probe tip with a nanotube mounted at the end<br />

of the tip. While this method of assembling<br />

nanoscale building blocks would be a powerful<br />

tool to test a concept, it is not suitable for the mass<br />

production. One of the more practical ways of<br />

fabricating the suggested hybrid structures might<br />

be electrophoresis/dielectrophoresis techniques<br />

FIGURE 59. Homogeneous emission through a nanodiamond cluster. Band structure for graphite/<br />

diamond/vacuum when (a) no field is applied and (b) under applied field. The difference between the<br />

Fermi level and CB edge depends on geometry and may vary between 5.5 eV and 3.0 eV. CB edge<br />

position in tetrahedrally coordinated a-C depends on the amorphization degree. Band structures in<br />

applied fields for a-C with zero (c) and positive (d) EA. (Reprinted from Ref.191.)<br />

318


involving directional movement of charged/polarized<br />

diamond particles in the suspensions under<br />

an applied electric field. Electrophoretic deposition<br />

of nanosized diamond particles from<br />

isopropyl alcohol suspensions on highly oriented<br />

pyrolytic graphite (HOPG) substrates has<br />

been thoroughly investigated in Ref. 71. The<br />

nanoparticles acquire a negative surface charge in<br />

aqueous and organic suspensions. However, in<br />

isopropyl alcohol suspensions, H+ ions generated<br />

in situ by the reaction between iodine and acetone<br />

adsorbed on the diamond particles, resulting<br />

in a positive surface charge and deposition<br />

of particles on a cathode. SEM and AFM pictures<br />

demonstrated that nanodiamond deposited<br />

as individual particles has a spherical shape. It<br />

was also revealed that the defect sites, such as<br />

steps on HOPG surface, are favorable for the<br />

deposition of particles. These techniques have<br />

been also used successfully for the deposition of<br />

nanodiamond powder on arrays of silicon tips<br />

for cold cathode fabrication. 180,181 Similar to the<br />

above results, deposition of nanodiamond powder<br />

on arrays of carbon nanotubes can be considered.<br />

IV. APPLICATIONS OF CARBON<br />

NANOSTRUCTURES<br />

In the worldwide study “Nanostructure Science<br />

and Technology” 1999, 281 four broadly defined<br />

and, in principle overlapping, application<br />

areas for nanostructure science and technology<br />

were suggested as a classification scheme. These<br />

areas are dispersions and coatings, high surface<br />

area materials, functional nanodevices, and consolidated<br />

materials. Figure 60 illustrates how this<br />

classification can be applied specifically for carbon-based<br />

nanostructures, which due to their abundant<br />

forms fill all four categories of the application<br />

fields.<br />

FIGURE 60. Schematic representation of four major areas of the applicability of the carbon nanostructures drawn<br />

according to the general classification of nanostructures in Ref. 281.<br />

319


Nanocarbon materials possess remarkable<br />

properties, and the potential applications look<br />

unlimited. While nanostructural materials from<br />

the graphite family have a well-established reputation<br />

in a variety of applications, 4 broad market<br />

acceptance of more esoteric nanocarbon-based<br />

technologies, however, will be enabled only if the<br />

fabrication costs are reduced and bulk production<br />

realized. For example, in 1998 a standard price of<br />

first gram- and subgram-sized nanotube samples<br />

was about $2000 per gram. Currently, prices vary<br />

from $500 a gram down to tens of dollars a gram<br />

for unpurified multiwall nanotubes. According to<br />

Fortune Magazine, 282 <strong>Carbon</strong> Nanotechnologies,<br />

Inc., TX, projects that as nanotube prices drop,<br />

more and more markets will open up with a total<br />

market value totaling $100 billion a year. According<br />

to Ref. 282, projecting the price of nanotubes<br />

down to $15,000 a pound, or about $33 a gram, a<br />

yearly supply of one ton would be able to support<br />

the manufacture of several billion dollars’ worth<br />

of flat-panel displays for PCs and television sets.<br />

In Table 23 we provide a tentative price list for<br />

nanodiamond particles, carbon nanotubes, and<br />

fullerenes as of 2002. Because the quality, purity,<br />

and exact form of the final product (powders, suspensions,<br />

etc.) vary, it is difficult to make meaningful<br />

price comparisons between different foundries,<br />

so the information provided does not pretend<br />

to advertise or diminish products of any of the<br />

companies listed. As can be seen from Table 23,<br />

prices for the highest purity nanodiamond particles,<br />

fullerenes, and nanotubes are all different by<br />

an order of magnitude in increasing order. An<br />

analysis of the economic impact of nanocarbon use<br />

shows that it increases the price of traditional products.<br />

For example, the utilization of ultradispersed<br />

diamond for polymer (rubber) composites in the<br />

amount of 1 gram per 1 kg of polymer 1 requires an<br />

additional investment of $2 per 1 kg of polymer<br />

composite. This sounds reasonable.<br />

Nanodiamond production by a single company<br />

currently can be of the order of tons per<br />

year, 1,3 so it is a quite mature technology. Technologies<br />

for purification and application in different<br />

areas are quite developed also (at least in<br />

countries of the Former Soviet Union (FSU)). 1,3<br />

Novel applications continue to emerge as the technology<br />

is developed on an industrial scale. For<br />

example, aggregate-free nanodiamond-metal solutions<br />

for electroplating have been developed for<br />

industrial production in Germany. 285 Among<br />

nanotube foundries, bulk quantity production is<br />

listed on the websites of Hyperion (MWNT),<br />

Rosseter Holdings Ltd, INP Toulouse France<br />

(MWNTs), Nanoledge (arc-grown SWNTs). For<br />

example, Guangzhou reports production of<br />

MWNT as 10 kg/day (95% purity) and SWNT as<br />

of 100 g/day (50% purity).<br />

In addition to the necessity of having effective<br />

and inexpensive methods of fabricating high<br />

quality nanostructures in bulk quantities (e.g.,<br />

for applications in nanocomposites), it is also<br />

important to develop controllable methods of<br />

nanostructure integration that can be scaled-up<br />

for volume production of functional devices. For<br />

example, for nanotube-based electronics, the<br />

strategies under development for achieving scaleup<br />

are self-assembly of nanospecies or self-alignment<br />

through controlled CVD growth on surfaces<br />

patterned with catalytic particles combined<br />

with microfabrication techniques. 204<br />

Below we discuss the applications of UDD in<br />

more detail as a field less familiar to many audiences.<br />

The applications of ultrananocrystalline diamond<br />

films and carbide-derived diamond are<br />

briefly outlined because these topics are thoroughly<br />

described in a number of recent publications.<br />

2,14,63 Among the most popular nanotube applications<br />

that have been described in recent<br />

reviews, we have chosen to discuss only the rapidly<br />

evolving areas of cold cathodes and functional<br />

devices. We conclude the section with a<br />

brief description of medical applications of<br />

fullerenes and the ever-increasing role that computational<br />

methods play in the development of<br />

new applications of carbon materials.<br />

A. Diamond-Based Nanostructured<br />

Materials for Macroscopic Applications<br />

Below we discuss in more detail current and<br />

potential applications of three different classes of<br />

nanodiamond, which, in our opinion, are the major<br />

classes of diamond-based materials on the<br />

nanotechnology market. Those materials are<br />

ultradispersed diamond particles (UDD), which has<br />

320


321


een on the market in FSU countries for decades,<br />

and two recently synthesized nanodiamond structures,<br />

namely, ultrananocrystalline diamond films<br />

and carbide-derived nanodiamond. While UNCD<br />

films possess very unique properties, carbide-derived<br />

diamond is distinct by the simplicity of its<br />

production. These three materials are synthesized<br />

by completely different techniques and have rather<br />

different properties, providing each of them specific<br />

application niches (Figure 61).<br />

1. Applications of Ultradispersed<br />

Diamond<br />

An extended review on the technological applications<br />

of UDD had been written recently by<br />

Dolmatov. 1 The material is used in the form of a<br />

powder, suspension, or paste. Below we provide<br />

the digest of UDD applications according to Ref<br />

1.<br />

a. Metal-Diamond Galvanic Coatings<br />

Electrochemical and chemical deposition of<br />

UDD together with metals using standard galvanic<br />

equipment has been demonstrated to be<br />

beneficial in a variety of applications in machine<br />

building, shipbuilding, the aircraft industry, tools<br />

for electronics, electrical engineering, medicine,<br />

and the watch and jewelry industry. The advantages<br />

of adding UDD to galvanic coatings include<br />

an increase in wear-resistance and microhardness<br />

FIGURE 61. Schematic representation of major areas of current and prospective applications of UDD,<br />

ultrananocrystalline diamond films and carbide-derived diamond films.<br />

322


(Table 24); an increase in corrosion resistance<br />

and a decrease in porosity; a dramatic decrease of<br />

friction coefficient; considerable improvement of<br />

adhesion and cohesion; and high throwing power<br />

of electrolyte. According to Ref. 1, 283, the service<br />

life of products is increased 2 to 10 times,<br />

even when the coating thicknesses is decreased<br />

by a factor of 2 to 3. The strengthening effect is<br />

observed in coatings of many metals, including<br />

silver, gold, and platinum, which are employed in<br />

numerous electronics applications. Particularly,<br />

UDD is most widely used in strengthening chromium<br />

coatings deposited using an electrolytic<br />

process. In this process, UDD-containing additives<br />

are added to the chrome-plating electrolyte<br />

without any modification to the standard production<br />

line. Such coatings increase by a few times<br />

the operating life of molds, high-precision bearing<br />

surfaces, and other similar components. UDD<br />

is also used in the production of cutting and razor<br />

tools.<br />

The UDD content in a metal coating averages<br />

0.3 to 0.5 wt. %. The amount of UDD consumed<br />

for a metal layer thickness of 1 mm is 0.2 g<br />

(1 carat) / m 2 . Table 24 provides more detailed<br />

information for different metals where UDD has<br />

been used successfully to enhance the performance<br />

of galvanic coatings. According to Dolmatov, 284<br />

metal-UDD galvanic coatings are a major market<br />

niche for the current UDD consumption from Diamond<br />

Center, Inc. followed by applications such as<br />

polishing materials and nanocomposites.<br />

In the near future PlasmaChem GmbH, Inc.<br />

will commercialize a new high purity aggregatefree<br />

ND product for application in the electroplating/electroless<br />

processes based on industrial processes/compositions<br />

developed in the framework<br />

of a European project. 285<br />

b. UDD for Polishing Pastes and<br />

Suspensions<br />

UDD pastes are being used for finishing precision<br />

polished materials for electronics, radio<br />

engineering, optics, medical, machine building,<br />

and jewelry industries 1 . The compositions with<br />

UDD allow one to obtain a surface of any geometrical<br />

form with a relief height roughness of<br />

2 to 8 nm. Recently, 4 Å roughness has been<br />

323


achieved for Al 2 O 3 , and SiC surfaces using UDD<br />

suspensions (according to the Alit, Inc., Ukraine).<br />

UDD is employed in polishing compositions used<br />

for the final treatment of silicon wafers in the<br />

microelectronics industry. UDD has been also<br />

used in the electronics industry as a component of<br />

heat-removal pastes and compounds for chip packaging,<br />

replacing the highly toxic beryllium oxide<br />

that has been used traditionally.<br />

The amount of UDD consumed for this application<br />

is 1 to 10 g/m 2 .<br />

c. UDD for Polymer Compositions<br />

Polymer composites with enhanced mechanical<br />

properties are required by the aircraft,<br />

motor and tractor, ship building, medical, chemical,<br />

and petrochemical industries, as well as in<br />

the manufacture of seals, stop valves for various<br />

purposes, and protective and antifriction<br />

coatings. The addition of UDD to polymers<br />

provides an increase in their mechanical<br />

strength, wear-resistance, and heat-ageing resistance<br />

(Table 25). Highly effective coatings<br />

based on the incorporation of UDD in fluoroelastomers<br />

and polysiloxanes were developed;<br />

the elastic strength of rubbers based on<br />

polyisoprene, butadiene-styrene, butadiene-nitrile,<br />

and natural rubbers were considerably improved.<br />

1,286 For example, for fluoroelastomers<br />

filled with UDD particles the stretch modulus<br />

at 100% elongation and the conditional rupture<br />

strength increased more than tenfold (from 8.5<br />

to 92 MPa and from 15.7 to 173 MPa, respectively).<br />

In this case, the elongation increased by<br />

a factor of 1.6 (from 280 to 480%). One of the<br />

mechanisms that have been attributed to the<br />

influence of UDD particles on the strength properties<br />

of polymer composites is an increase in<br />

cross-linking. 1 Additives of UDD into the rubbers<br />

decrease attrition wear by an average of 3<br />

to 5 times, increase rupture strength by 30%,<br />

and breaking temperature by 15%. Epoxy adhesives<br />

that incorporate UDD have high adhesion<br />

and cohesion properties.<br />

The specific consumption of UDD or diamond<br />

blend (mix containing UDD and significant<br />

percent of other carbon based products of detonation<br />

explosion) is 1 to 5 kg per 1000 kg of rubber<br />

(polymer) and 1 to 5 kg per 1000 m 2 of polymer<br />

coating or film.<br />

324


d. UDD for Lubricating Oils, Greases, and<br />

Lubricant Coolants<br />

Modified lubricant compositions are used in<br />

machinery, metal treatment, engine building, ship<br />

building, and the aircraft and transportation industries.<br />

The addition of UDD and diamond blend to oils<br />

allow one to obtain sedimentation-stable and environmentally<br />

safe systems with particle sizes of less than<br />

0.5 µm. 1 The use of nanodiamonds in oils increases<br />

the service life of motors and transmissions. Friction<br />

torque is reduced by 20 to 40%, and the wear of<br />

rubbed surfaces is decreased by 30 to 40%.<br />

The specific consumption of UDD or DB in<br />

these applications is 0.01 to 0.2 kg per 1000 kg of<br />

oil.<br />

e. UDD for Systems of Magnetic Recording<br />

UDD is also used as an antifriction additive<br />

and a physical modifier for ferro-lacquer coatings<br />

of magnetic tapes and disks, and also as an additive<br />

to electrochemical deposition of composite magneto-recording<br />

tapes to improve the properties of<br />

magnetic recording devices. 287 The addition of UDD<br />

decreases ferro-magnetic grain size, thus allowing<br />

an increase in recording density while reducing<br />

abrasive wear and friction coefficient<br />

f. UDD for Application in Intermetallics<br />

Based on Copper, Zinc, and Tin<br />

UDD can be used in specific application fields<br />

such as when surfaces experience high frictional<br />

forces produced by very harsh working conditions<br />

that displace plastic and liquid lubricating<br />

materials. UDD is an ideal composite material for<br />

intermetallics based on copper with zinc or tin<br />

(the UDD content is no more than 15 volume %).<br />

The addition of UDD decreases the frictional<br />

forces two to six times. 288<br />

g. UDD for Biology and Medicine<br />

According to Refs. 1, 289 UDD is considered<br />

a potential medical agent (not just drug delivery<br />

agent) in oncology, gastroenterology, vascular<br />

disease, and an efficient remedy for the aftereffects<br />

of burns and skin diseases. The beneficial<br />

property of UDD for medical applications is its<br />

anomalously high adsorbtion capacity, high specific<br />

surface areas, abundance of free electrons on<br />

the surface (a multiple radical donor), nanoscale<br />

size, significant amount of oxygen-containing<br />

functional groups on the surface, chemically inert<br />

cores and hydrophoblicity of the surface. Due to<br />

their high adsorption capacities, UDD exhibits<br />

extremely high absorbing/bonding activities with<br />

respect to pathogenic viruses, microbes, and bacteria.<br />

The absorbing/bonding may be selective to<br />

particular drugs, which can enhance the drug’s<br />

activity. According to Refs. 1, 289, UDD exhibits<br />

no carcinogenic or mutagenic properties and is<br />

not toxic.<br />

The use of UDD in the form of aqueous and<br />

oil suspensions showed promising results for curing<br />

cancer by removing toxins from organism,<br />

normalizing peristaltic of the bowels, and improving<br />

blood characteristics. 289 The use of UDD<br />

in chemical and radiotherapy appears to be promising<br />

in the cure of malignant tumors by preventing<br />

the mutagenic effect of drags. According to<br />

Ref. 289, a theraupeutic course requires ~0.02 to<br />

0.5 g of UDD. While preliminary results, particularly<br />

in oncology, are encouraging, 289 this area of<br />

reseach has been explored very little because of<br />

insufficient funding for the research in the countries<br />

of the FSU where the application was first<br />

developed.<br />

A research group from the Ukraine also recently<br />

reported a study on the use of UDD as an<br />

adsorbent for the purification of biological media.<br />

290<br />

h. UDD Dinter as an Adsorbent of a New<br />

Type<br />

<strong>Carbon</strong>-containing adsorbents are widely used<br />

in various industries such as medicine and pharmacology.<br />

The most abundant of these adsorbents<br />

are activated coals and graphitized thermal carbon<br />

black. Synthetic diamond, particularly submicron<br />

diamond composites as well as sintered<br />

UDD, represents a new class of carbon containing<br />

325


adsorbents, 291 characterized by chemical inertness<br />

and high strength. Another beneficial effect is the<br />

possibility of using the adsorbent repeatedly by<br />

modifying and recovering the diamond surface.<br />

In addition to the application areas listed<br />

above, UDD is used as a component in the production<br />

of diamond ceramics and molds made of<br />

diamond-containing materials. UDD had also been<br />

employed for seeding substrates used in the CVD<br />

growth of diamond films. Dry UDD is known to<br />

absorb and retain water in amounts that are four<br />

times the weight of the UDD. This allows its use<br />

as an inert solid water absorber in materials whose<br />

quality is determined by the residual water content,<br />

for example, in magnetic carriers.<br />

2. Applications of Pure Phase<br />

Ultrananocrystalline Diamond Films<br />

As discussed in Section III.A, ultrananocrystalline<br />

diamond is grown using a new plasma deposition<br />

process that utilizes a high content of noble gas. This<br />

process was developed at Argonne National Laboratory<br />

and produces films with ultrasmall (2 to 5 nm)<br />

grains with atomically abrupt grain boundaries. UNCD<br />

films are superior in many ways to traditional microcrystalline<br />

diamond films: they are smooth, dense,<br />

pinhole free, and phase-pure, and can be conformally<br />

coated on a wide variety of materials and high-aspectratio<br />

structures. The set of unique properties include<br />

mechanical (high hardness ~100 GPa, and Young’s<br />

modulus ~960 GPa), tribological (extremely low friction<br />

~0.01), transport (tunable electrical conductivity,<br />

high thermal conductivity), electrochemical (wide<br />

working potential window), and electron emission<br />

(low, stable threshold voltage) characteristics. The<br />

UNCD has been considered for a variety of applications,<br />

including MEMS and moving mechanical assembly<br />

devices, surface acoustic wave (SAW) devices,<br />

biosensors and electrochemical sensors, coatings<br />

for field emission arrays, photonic and RF switching,<br />

and neural prostheses.<br />

Studies of UNCD-coated flat substrates and<br />

microtip arrays for cold cathodes applications 2,292<br />

have yielded consistently low threshold fields<br />

(1 to 2 V/µm), high total emission currents (up to<br />

10 mA), and stable emission during long-duration<br />

testing (up to 14 days).<br />

Ultrananocrystalline diamond has been also<br />

grown with the incorporation of nitrogen up to 8<br />

× 10 20 atoms/cm 3 with the addition of nitrogen to<br />

plasmas during the CVD growth of diamond<br />

films. 292 This is the highest carrier concentrations<br />

seen for any n-type diamond material to date<br />

resulting in several orders of magnitude increase<br />

in UNCD conductivity that promise applications<br />

in heterojunction electronic devices.<br />

UNCD electrodes exhibit a wide working<br />

potential window, a low background current, and<br />

high degree of electrochemical activity for redox<br />

systems. These results, in combination with the<br />

biocompatibility properties of UNCD, could lead<br />

to the application of UNCD electrodes for nerve<br />

stimulation. 2<br />

Future MEMS applications that involve significant<br />

rolling or sliding contact (MEMS moving<br />

mechanical assemblies (MEMS MMAs)) will require<br />

the use of new materials with significantly<br />

improved mechanical and tribological properties<br />

and the ability to perform in harsh environments.<br />

Because the feature resolution in polycrystalline<br />

MEMS is limited by grain size, the use of MEMS<br />

made by conventional CVD diamond methods is<br />

limited. In addition, the conventional CVD diamond<br />

films typically have large grain sizes (~1<br />

mm), high internal stress, poor intergranular adhesion,<br />

and rough surfaces (rms ~1 µm). Alternatively,<br />

diamond-like coatings, generally grown by<br />

physical vapor deposition, cannot cover high aspect<br />

ratio MEMS features conformally and require<br />

high-temperature post-deposition processing to<br />

relieve stress, which compromises their mechanical<br />

properties. Ultrananocrystalline diamond coatings<br />

possess morphological and mechanical properties<br />

that are ideally suited for MEMS applications<br />

in general, and MMA use in particular. 2,293 The<br />

roughness of the film is about 20 to 40 nm and the<br />

friction coefficient can be as low as 0.01. The<br />

surfaces are very smooth (rms~30 to 40 nm) and<br />

the hardness is as high as ~100 GPa. When compared<br />

with Si-based MEMS, the brittle fracture<br />

strength is 23 times that of Si, and the projected<br />

wear life of MEMS MMAs from diamond is 10,000<br />

times greater than that of Si MMAs. The group<br />

from Argonne National Laboratory demonstrated<br />

three-dimensional MEMS structures fabricated<br />

from UNCD material, including cantilevers and<br />

326


multilevel devices, acting as precursors to<br />

microbearings and gears (Figure 62).<br />

Applications of UDNC as biocompatible MEMS<br />

devices, biosensors and biological electrodes are being<br />

also explored. A special program is funded by the DOE<br />

to develop UDNC-based artificial retinas to restore<br />

sight to people blinded by the retinitis picmentosa<br />

condition. 294 Functionalization of UNCD to attach DNA<br />

molecules has been demonstrated also. 294<br />

Nanocrystalline diamond coatings on suitable<br />

substrates are promising materials for medical<br />

implants, cardiovascular surgery, and for the coating<br />

of certain components of artificial heart valves<br />

due to their extremely high chemical inertness,<br />

smoothness of the surface, and good adhesion of<br />

the coatings to the substrate. 295 Nanocrystalline<br />

diamond coatings (reported grain size ~nm) deposited<br />

by RF-PCVD of methane with nitrogen<br />

on mm-sized steel implants that had been inserted<br />

to tissue and bones for up to 52 weeks 295 demonstrated<br />

excellent biocompatibility and biostability<br />

(Figure 63).<br />

We would like to conclude this subsection by<br />

outlining the medial application of diamond from a<br />

completely different perspective. Within the concepts<br />

of the molecular nanotechnology, 308 carbonbased<br />

materials could play an important role in building<br />

“nanorobots”, structures that have been<br />

envisioned for applications in nanomedicine 309 (Figure<br />

64). Nanomedicine may be defined as the monitoring,<br />

repair, construction, and control of human<br />

biological systems at the molecular level using engineered<br />

nanodevices and nanostructures. 309<br />

Fullerene nanotube–based rotors, shuffles, and<br />

gears 310 have been computationally designed as well<br />

as those based on so-called diamonoid materials. 311<br />

3. Applications of Carbide-Derived<br />

Diamond-Structured <strong>Carbon</strong><br />

Recently, a method for the synthesis of diamond-structured<br />

carbon in bulk quantities by extracting<br />

silicon from silicon carbide or metal car-<br />

FIGURE 62. A UNCD MEMS cantilever strain gauge with a 100-nm feature resolution<br />

produced by UNCD blanket film deposition and etching through a hard oxide mask.<br />

The UNCD film is 3.3 µm thick, and it is separated from the substrate by 2 µm.<br />

(Reprinted from Ref. 293, with permission from Elsevier Science.)<br />

327


FIGURE 63. Illustration of a very good biotolerance of the implants coated with the nanocrystalline<br />

diamond layers. In subcutaneous tissue, muscles and bones, thin connective tissue capsules<br />

built from fibrocytes and collagen fibers were formed. Optical microscope analysis of the wall<br />

of the capsule after 52 weeks did not reveal any phagocytic reaction or products of corrosion.<br />

(Reprinted from Ref. 295, with permission from Elsevier Science.)<br />

FIGURE 64. Science fiction: using diamond in medical nanorobots. The picture is adapted<br />

from M. D. Lemonick, “...And Will They Go Inside Us? Given the Promise of Nanotechnology,<br />

It’s a Safe Bet,” Time Magazine 154 (8 November 1999):93. Artist: Joe Letrola.<br />

328


ide was developed. 14 In principle, chlorination of<br />

carbides for the production of carbon-based materials<br />

and, particularly, nanoporous carbon, is a relatively<br />

mature technology that has been commercialized (see,<br />

for example, http://www.skeleton-technologies.com).<br />

However, the synthesis of nanocrystalline diamond<br />

by this technique 14 is a recent achievement. Coatings<br />

of diamond-structured carbon produced by this route<br />

show hardness values in excess of 50 GPa and Young’s<br />

moduli up to 800 GPa.<br />

The carbide-derived carbon and diamond coatings<br />

show excellent tribological behavior in both room<br />

air and dry nitrogen and are at the stage of commercialization<br />

for tribilogical applications, particularly as<br />

nanodiamond coatings for SiC dynamic seals for<br />

water pumps. 296 The coatings are self-lubricating, with<br />

remarkably low friction coefficients that can be tailored<br />

by altering the reaction parameters and show no<br />

measurable wear. 296 Favorable tribological properties<br />

of carbide-derived nanodiamond make it a favored<br />

candidate for applications in the manufacturing of<br />

different types of prosthesis. 297<br />

Conformal coatings produced by selective etching<br />

can be useful in micro-electro-mechanical systems<br />

(MEMS) applications, where very thin and<br />

uniform coatings are required. In addition, permeability<br />

of the films produced by chlorination of SiC<br />

and an extremely narrow pore size distribution in<br />

carbide-derived carbon provide effective molecular<br />

sieves, high-surface area electrodes and other<br />

applications, where vapor-deposited diamond films<br />

cannot be applied. The large-scale solid-state synthesis<br />

of technical diamond at ambient pressure<br />

and moderate temperatures with no plasma activation<br />

can provide diamond materials at a low cost<br />

for a variety of high-volume applications such as<br />

brake pads, where diamond could not be used before<br />

because of its cost.<br />

B. <strong>Carbon</strong> Nanotubes in Advanced<br />

Electron Sources: Are <strong>Carbon</strong><br />

Nanotubes Exceptional Electron<br />

Sources?<br />

Advanced electron sources often are regarded<br />

as the most important short-term application of<br />

carbon nanotubes. In this section, we summarize<br />

the technical parameters of nanocarbon electron<br />

emitters and provide critical assessments of their<br />

status and prospects. Electron emission properties<br />

of carbon nanotubes are considered.<br />

329


330


1. Practical Issues of Field Emission<br />

Electron Sources<br />

Field emitters are cold electron sources whose<br />

operation is based on geometry effects. Despite<br />

their apparent simplicity and the expectations about<br />

device applications, field emitters are rarely used<br />

in real-world devices. The reason is that so far<br />

field emitters do not satisfy the practical technical<br />

requirements of high current density and emission<br />

stability, 314,315 as well as a low operating<br />

voltage.<br />

The two major designs of electron sources are<br />

a single point electron source (e.g., a single<br />

nanotube) and a large area electron source, which<br />

could be either an array of tips (e.g., nanotubes)<br />

or a continuous film of emitting material.<br />

In this section we provide critical assessments<br />

of the status of carbon emitters by comparison<br />

with their most recently reported emission parameters<br />

to those of field emitters made from other<br />

materials (Figures 65 and 66).<br />

2. <strong>Carbon</strong> Nanotubes Emitters<br />

Typical carbon nanotubes field emitters are<br />

shown in Figure 67. With their high aspect ratio<br />

and unique conducting properties, there has been<br />

a great deal of recent speculation regarding field<br />

emission from carbon nanotubes, with some researchers<br />

claiming properties that are far superior<br />

to more conventional field emitting structures. In<br />

this section we attempt to critically review carbon<br />

nanotubes field emission cathodes from the viewpoint<br />

of the field emission community, including<br />

candid comparisons to other well-established field<br />

emitters. In our comparison, we categorize different<br />

types of field emitters as “low-voltage emitters”,<br />

“high-current emitters”, and “gated emitters”.<br />

a. Low-Voltage Emitters<br />

Low-voltage (10 to 20 V) field emission from<br />

carbon nanotubes was reported recently in Ref.<br />

319, where a very small emitter-to-anode distance<br />

(6 µm) was used. We note that similar<br />

effects can be achieved with other types of field<br />

emitters. The first field emission diode that operated<br />

at less than 10 V applied was reported in<br />

1989 by Makhov. 314 It consisted of sharp wedges<br />

of single crystal silicon operated at a cathode-toanode<br />

distance of less than 1 µm. Since then,<br />

several similar devices have been demonstrated,<br />

including recent results from diamond tips (see<br />

Table 26). In all cases, the low-voltage operation<br />

FIGURE 65. Conventional field emission technologies: (a) Single Mo tip field emitter (etched wire);<br />

(b) ungated Si field emitter arrays. (Reprinted from Ref. 316, Copyright 1994, with permission from<br />

American Institute of Physics.)<br />

331


FIGURE 66. Gated Mo field emitters. (Reprinted from Ref. 317, Copyright<br />

1998, with permission from American Institute of Physics.)<br />

was achieved by using very small vacuum gaps of<br />

less than 10 µm. Unfortunately, this type of device<br />

has limited usefulness; they are all essentially<br />

diodes with little or no advantages compared<br />

with solid-state diodes. Table 26 compares<br />

recently reported low-voltage operation of carbon<br />

nanotubes field emission diodes with previously<br />

reported field emission devices.<br />

b. High-Current Emitters<br />

The potential for high-emission currents has<br />

always been an attractive feature of field emitters,<br />

and there are many reports of “high current densities”<br />

from novel emission materials. However,<br />

these reports can be misleading and must be interpreted<br />

carefully in terms of practical applications.<br />

For example, a small total current of 10 µA can<br />

result both in very high and very low current<br />

density depending on the choice of emission area.<br />

The emission area of practical merit is the total<br />

cathode area α total (independent of the number of<br />

emission sites). All of the reported data regarding<br />

carbon nanotubes field emission has been obtained<br />

using a proximity probe of very small area<br />

(1 to 100 µm 2 ), with the total current measured in<br />

these experiments typically only 1 to 100 µA.<br />

Nevertheless, claims of high-current densities and<br />

device applications for microwave tubes have been<br />

made. The practical question, however, is “what<br />

will the current be when the anode diameter is<br />

increased to 1 cm?” It is likely that the current<br />

will be much less than 10,000 amperes. As was<br />

shown by Göhl et al., 323 an increase in probe<br />

diameter results in a drastic decrease in probe<br />

current density.<br />

The parameter of merit for device applications<br />

is the integral current density, that is, the<br />

total current emitted divided by the entire cathode<br />

area. That is, a very high current density is<br />

obtained only from a very small cathode area.<br />

(The record current density of about 2500 A/<br />

cm 2324 was obtained from an array of emitters<br />

332


FIGURE 67. Fragments of typical CNT field emitters: (a) supported single nanotube; 329 (b)<br />

array of CNT grown by CVD (at different magnifications); 326 (c) attached CNTs. 332 (Reprinted<br />

from Ref. 329 with permission from Springer-Verlag GmbH & Co. KG (a) and Ref.<br />

326 (b) and Ref. 332 (c) with permission from the American Institute of Physics.)<br />

333


with an integral area of 25 × 25 µm 2 and a total<br />

current of a few mA.) A realistic challenge is to<br />

demonstrate a total current of 1 A (or more) from<br />

a macroscopic area of 1 cm 2 . Overall judgements<br />

about the potential usefulness of new cathode<br />

materials cannot be made without data from a<br />

minimum set of practical parameters: maximum<br />

total current at failure I max , integral current density<br />

J max , operating voltages V th (threshold voltage),<br />

and V max (voltage for maximum current or<br />

failure) and transconductance g m ~ I max / V max . 315<br />

Note that the typical maximum current of “highcurrent”<br />

carbon nanotubes emitters is within 0.1<br />

to 2 mA. For comparison, 4 mA of current was<br />

obtained from a single ZrC tip. 327 Table 27 compares<br />

maximum current and transconductance of<br />

different field emitters.<br />

c. Gated Emitters<br />

Gated field emitters shown in Figure 66 are<br />

basic components of field emission displays (FED)<br />

and microtriodes for vacuum microelectronics<br />

devices. 314 Typical gate voltages of metal field<br />

emitters prepared by standard technique is in the<br />

range 50 to 100 V. To date, only a few reports on<br />

gated carbon nanotubes emitters have been published.<br />

Hsu and Shaw 328 grew multiwalled carbon<br />

nanotubes on conventional silicon gated field<br />

emitters with 2.5 µm gate aperture. The silicon<br />

tips in these particular emitters were relatively<br />

blunt. A gated field emitter cell with carbon<br />

nanotubes grown on Si tips is shown in Figure 68.<br />

However, it was found from SEM examination<br />

that only 10% of Si tips had carbon nanotubes. 328<br />

Figure 69 shows emission current-voltage characteristics<br />

from Si field emission arrays with and<br />

without carbon nanotubes. As can be seen, emitters<br />

without carbon nanotubes had a much higher<br />

threshold voltage. Field emission arrays with carbon<br />

nanotubes showed threshold voltages near 20<br />

V. Table 28 compares emission characteristics of<br />

carbon nanotubes 328 and Mo 317 gated field emitters.<br />

Indeed, carbon nanotubes allow for the reduction<br />

of the operational voltage for emitters<br />

with relatively large gate aperture. However, the<br />

current per cell and total current of a carbon<br />

nanotubes field emission array is considerably<br />

lower than metal field emitters.<br />

d. Maximum Current from Individual<br />

Nanotubes<br />

Experiments with individual carbon nanotubes<br />

s mounted on metal tips have shown maximum<br />

emission currents of the order of 100 µA. 329<br />

334


FIGURE 68. Gated carbon nanotubes-on-silicon post field emitter cell.<br />

(Reprinted from Ref. 328, Copyright 2002, with permission from the<br />

American Institute of Physics.)<br />

335


336<br />

FIGURE 69. Emission current-voltage characteristics (anode current vs gate<br />

voltage) from arrays with and without CNTs. (Reprinted from Ref. 328, Copyright<br />

2002, with permission from the American Institute of Physics.)


Nilsson et al. 330 studied the emission-degradation<br />

behavior of carbon nanotubes thin-film electron<br />

emitters. The authors investigated MWNT arrays<br />

grown by CVD on a silicon wafer. They found<br />

that current-dependent emission degradation<br />

started at about 300 nA per emitter. This result is<br />

consistent with Ref. 328 (see Table 28). In the<br />

authors’ opinion, the primary degradation mechanism<br />

of a carbon nanotubes film on Si was the<br />

Joule heating at the carbon nanotubes /silicon<br />

interface. 330 The authors 330 discuss the difference<br />

between their result and the result reported by<br />

Bonard et al., 329 where a maximum current of 100<br />

µA was obtained from an individual MWNT<br />

mounted on the tip of an etched gold wire. This<br />

difference can be explained in part by the different<br />

contact resistance between carbon nanotubes<br />

and gold and carbon nanotubes and Si. 329<br />

In general, the average current per emission<br />

site in a carbon nanotubes emitter is smaller than<br />

in metal field emission arrays. The maximum<br />

current of 100 µA obtained from individual<br />

nanotubes is also typically the maximum current<br />

for metal (e.g., Mo) emitters, 314 though emitters<br />

from metal carbides, 327 or metal tips with diamond<br />

coatings, 331 result in larger maximum currents,<br />

up to few mA per tip. 327<br />

has to keep in mind that such an application must<br />

compete with existing fluorescent tubes. In general,<br />

to achieve a brightness of 10 000 cd/m 2 , the<br />

emission current density should be about 0.5 to 1<br />

µa/cm 2 at voltages of 3 to 5 kV. For luminescent<br />

tubes of practical size these conditions will result<br />

in high power consumption. Indeed, the power<br />

consumption of the reported carbon nanotubes<br />

field emission lamp is more than 10 times higher<br />

than for a conventional fluorescent tube. 334 The<br />

authors 334 believe that it would be possible to<br />

decrease the power consumption by phosphor<br />

optimization. It should be noted, however, that<br />

the phosphor efficiency by electron excitation is<br />

always lower than the efficiency by UV excitation<br />

for fundamental physical reasons, and therefore<br />

it may not be possible to fabricate field emission<br />

lamps with efficiencies comparable with<br />

fluorescent tubes.<br />

Among other proposals for using carbon<br />

nanotubes as field emission electron sources, we<br />

mention X-ray tubes, 335 electron guns for electron<br />

microscopy, 336 and vacuum microtriodes. 337 However,<br />

unless considerable improvement in integral<br />

current density and emitter lifetime is shown,<br />

major breakthroughs in the performance or application<br />

of these devices will be difficult to achieve.<br />

3. Field Emission Devices<br />

a. Displays and Lamps<br />

Field emission flat panel displays are often<br />

cited as one of the most promising applications of<br />

carbon nanotube field emitters, and several groups<br />

have demonstrated the operation of carbon<br />

nanotubes displays. Probably, the most advanced<br />

FED prototypes were made at Samsung. 332,333 We<br />

note somewhat pessimistically, however, that<br />

many different kinds of field emission displays<br />

have been demonstrated during the last 12 years,<br />

with but so far none have proven to be a successful<br />

commercial alternative to other technologies<br />

such as LEDs, electroluminescent, and plasma<br />

displays.<br />

One of the recently proposed applications of<br />

carbon nanotubes was in the luminescent tubes<br />

for general lighting purposes. 334 However, one<br />

4. Emission Mechanism and Specific<br />

Features of <strong>Carbon</strong> Nanotubes<br />

In general, carbon nanotubes emitters are<br />

based on the same operation principle as more<br />

traditional emitters, such as metal tips. As a general<br />

observation, the open-ended carbon nanotubes<br />

appear to emit better than those with caps, and<br />

SWNTs emit better than MWNTs.<br />

The key factor determining emission properties<br />

of carbon nanotubes is the geometrical field<br />

enhancement, which for individual emitters is<br />

roughly proportional to their aspect ratio. The<br />

difference between carbon nanotubes and conventional<br />

field emitters and the potential advantage<br />

of carbon nanotubes is their cylindrical shape,<br />

which may enable higher field enhancement.<br />

However, this advantage turns into a drawback<br />

when we consider the maximum current from<br />

individual nanotube emission sites. It is limited to<br />

337


hundreds of nA, 328,330 for example, very different<br />

from tens of µA for metal tips. 314,317<br />

Another difference between carbon nanotubes<br />

and conventional metal field emitters is the fact<br />

that carbon nanotubes can be flexed, bent, and<br />

reoriented by an electric field. 338 For moderate<br />

electric fields, the flexing and reorienting is reversible,<br />

but under high-field conditions sufficient<br />

to extract large field emission currents, the<br />

nanotube remains irreversibly deformed.<br />

Nanotubes that field emitted at a high currents<br />

for long times were foreshortened 338 , suggesting<br />

a lifetime-limiting mechanism for vacuum<br />

microelectronic devices.<br />

Conclusion on Field Emission from <strong>Carbon</strong><br />

Nanotubes<br />

Despite statements claiming carbon nanotubes<br />

to be “excellent”, high-current, and low-field<br />

emitters, realistic assessment suggests that carbon<br />

nanotubes may ultimately offer few advantages<br />

over alternative field emission technologies. In<br />

fact, a more precise comparison of conventional<br />

metal field emitters and carbon nanotubes emitters<br />

(see Tables 26 to 28) does not reveal any<br />

remarkable advantages of the latter.<br />

C. <strong>Carbon</strong> Nanotubes as<br />

Nanoelectronics Components<br />

<strong>Carbon</strong> nanotubes are considered as a very<br />

important subset of nanoelectronic materials. 339<br />

Semiconducting properties of carbon nanotubes<br />

made them interesting for electronics applications.<br />

One of the requirements for a material for active<br />

electronic devices is the ability to control its electron<br />

transport properties. In CNTs such control<br />

can be achieved by geometry (e.g., diameter, shape,<br />

helicity) or surface adsorbates.<br />

The operational principle of most electronic<br />

devices is based on either conductivity modulation<br />

of a channel or on highly nonlinear I-V<br />

characteristics. Both field effect transistors<br />

(FETs) and intratube p-n junctions made from<br />

individual semiconducting nanotubes have been<br />

demonstrated.<br />

Doped <strong>Carbon</strong> Nanotubes Structures<br />

As-grown semiconducting carbon nanotubes<br />

have p-type conductivity. 340–343 <strong>Carbon</strong> nanotubes<br />

can be doped chemically by controlling surface<br />

adsorbates. 341–343 Doped structures with p-n junctions<br />

were formed on individual single-walled<br />

carbon nanotubes by doping with alkaline metals<br />

and organic molecules. Doped structures with both<br />

negative differential conductance 341 and unipolar<br />

conductance (rectification) 342 were demonstrated.<br />

In a recent work by Kong et al. 340 a p-n-p junction<br />

was chemically defined on an individual carbon<br />

nanotubes. The n-doped region was produced by<br />

local deposition of potassium on a p-type carbon<br />

nanotubes. From transport measurements, the<br />

authors 340 concluded that nanometer-scale width<br />

tunnel barriers at the p-n junctions dominate the<br />

electrical characteristics of the system. Figure 70.<br />

shows a schematic p-n-p carbon nanotubes device<br />

and the corresponding band diagram. At low temperatures,<br />

the structure can be regarded as a quantum<br />

dot confined between the two p-n junctions.<br />

Single electron transistor behavior of this carbon<br />

nanotubes p-n-p structure was reported. 340 In principle,<br />

by local doping of carbon nanotubes, the<br />

realization of such functional devices as Esaki<br />

tunnel diodes or Shokley p-n-p-n diodes should<br />

be possible.<br />

Field Effect Transistors<br />

The possibility for modulating the conductance<br />

through a carbon nanotubes by a gate electrode<br />

was first demonstrated by Dekker et al. in<br />

1998. 344 Since then, several groups have demonstrated<br />

field effect transistor (FET ) device structures<br />

in which a gate electrode modulates the<br />

conductivity of a conducting channel by a factor<br />

of 10 5 . 345 Large arrays of carbon nanotube FETs<br />

have been fabricated. 346 Simple prototypes of electronic<br />

circuit elements were demonstrated, such<br />

as a voltage inverter or NOT gate circuit using<br />

one n-channel and one p-channel FET.<br />

In carbon nanotubes FETs, the carbon<br />

nanotubes acts as a channel, connecting two metal<br />

electrodes. The channel conductance is modulated<br />

by third gate electrode. Several different<br />

338


FIGURE 70. (a) Schematic CNT p-n-p structure; (b) AFM image showing a 200-<br />

nm wide window (dark), corresponding to potassium doped n-type region; (c)<br />

corresponding band diagram. (Reprinted from Ref. 340, Copyright 2002, with<br />

permission from American Institute of Physics.)<br />

types of carbon nanotubes FET were reported:<br />

“bottom gate”, 344 “top gate”, 347 and vertical. 348<br />

The top gate structure shown in Figure 71 347 appears<br />

the most promising one, because it allows<br />

for a thinner gate insulator, protects the carbon<br />

nanotubes channel from exposure to air, and can<br />

be made suitable for high-frequency operation. 347<br />

Wind et al. 347 recently reported top gate<br />

p-channel carbon nanotubes FET with channel<br />

length of 260 nm and having a maximum<br />

transconductance of 3.25 µS. The authors compared<br />

the parameters of the carbon nanotubes FET<br />

to the state-of-the art Si transistors (see table in<br />

Figure 71c). In the authors’ opinion, the dc performance<br />

of a carbon nanotubes FET is much better<br />

than a Si FET. It should be noted, however, that in<br />

their comparision the authors used electrical<br />

charcteristics normalized per channel width, which<br />

was assumed to be 1.4 nm. Another assumption<br />

was that only one nanotube was present in the<br />

channel. The third assumption is that the<br />

transconductance will increase proportional to the<br />

number of parallel nanotubes in the channel. All of<br />

these assumptions need to be carefully investigated.<br />

For circuit operation of a FET, the total<br />

transconductance is one of the key parameters. The<br />

total transconductance reported in Ref. 347 is<br />

3.25 µS, which is much smaller than the<br />

339


340<br />

FIGURE 71. (a) Schematic cross-section of the top gate CNT FET showing the gate<br />

and source and drain electrodes. (b) Output characteristic of a top gate p-type<br />

CNFET with a Ti gate and a gate oxide thickness of 15 nm. The gate voltage values<br />

range from 20.1 to 21.1 V above the threshold voltage, which is 20.5 V. Inset:<br />

Transfer characteristic of the CNFET for V ds = 20.6 V. (c) Comparison of key device<br />

parameters for CNT and Si FET. (Reprinted from Ref. 328, Copyright 2002, with<br />

permission from American Institute of Physics.)


transconductance of practical Si FETs. Javey et<br />

al. 349 reported fabrication nanotube FET arrays with<br />

local bottom gates.<br />

While as-made FETs are all p-type, a local<br />

electrical manipulation method was used to convert<br />

specified nanotube FETs into n-type. The authors<br />

349 found that a p-type FET can be converted<br />

into n-type by applying a high local gate voltage<br />

combined with a large source-drain bias for a certain<br />

duration. This conversion is shown in the current<br />

vs bottom gate voltage (I-V g ) curves in Figure<br />

72a for a transistor before and after applying a<br />

local gate voltage of -40 V and a source-drain bias<br />

of 20 V for 5 min. The authors note, however, that<br />

some of the p-FETs become rather insulating over<br />

a large gate range after this electrical manipulation<br />

step. The yield of p- to n-conversion by our local<br />

manipulation method was about 50%.<br />

This local doping approach leads to multiple<br />

n-FETs coexisting with p-FETs on a chip. Complementary<br />

logic gates with up to six complementary<br />

transistors and three stage ring oscillators were<br />

realized by connecting the n- and p-transistors.<br />

Figure 72b shows the output characteristics of an<br />

FIGURE 72. Local manipulation for n-type tube FETs for complementary devices. (a)<br />

Source-drain current vs. gate voltage (I-Vg) curves of a SWNT-FET in air and after local<br />

electrical manipulation in vacuum, showing the p- to n-type conversion. Bias = 10 mV. (b)<br />

Transfer characteristics of an inverter made from a p-type nanotube FET and an n-type FET<br />

obtained by electrical manipulation. The operating voltage applied is V DD = -5 V. (Reprinted<br />

from Ref. 349, Copyright 2002, with permission from the American Chemical Society.)<br />

341


inverter (NOT gate). It consists of an as-made p-<br />

type tube FET and an n-type FET, and obtained<br />

by electrical manipulation. This complementary<br />

NOT logic gate exhibits a high voltage gain of 8.<br />

Complementary NOR, OR, NAND, and AND<br />

logic gates are built with up to six (3 p-type/3 n-<br />

type) SWNT-FETs. 349<br />

All published characteristics of carbon<br />

nanotubes electronic logic gates are static I-V<br />

characteristics. Data on the dynamic time-dependent<br />

response of these components are very important<br />

for fair assessments of their usefulness for<br />

practical nanoelectronic devices. At this point,<br />

there are only very few reports on the operational<br />

speed of carbon nanotubes devices. Ring oscillators<br />

with an oscillation frequency of 5 Hz made<br />

using three unipolar p-type carbon nanotubes FETs<br />

and resistors were reported in Ref. 350. Javey et<br />

al. 349 reported a complementary carbon nanotubes<br />

FET ring oscillator with three SWNT based inverters.<br />

The oscillating frequency of the device<br />

was measured to be about 220 Hz (Figure 73).<br />

Thus, the speed of carbon nanotubes devices, reported<br />

at this point, is very slow. The authors 349<br />

explain this as due to “extrinsic” factors such as<br />

external resistance and capacitance of the circuit.<br />

The demonstration of high-speed operation of<br />

nanotube electronic devices is probably the most<br />

important issue for assessments of the feasibility<br />

of nanotube electronics.<br />

The theoretical understanding of the operation<br />

of a carbon nanotubes FET remains incomplete.<br />

351 While initially it was believed that the<br />

gate voltage changes the conductivity of carbon<br />

nanotubes channel, later there was increasing<br />

evidence that the Schottky barriers at the carbon<br />

nanotubes/metal contacts may play a key role.<br />

Heinze et al., 351 based on their theoretical results,<br />

concluded that carbon nanotubes FETs operate as<br />

unconventional Schottky barrier transistors in<br />

which switching occurs primarily by modulation<br />

of the contact resistance rather than the channel<br />

resistance. Clearly, more work is needed for a<br />

complete understanding of carbon nanotubes FET<br />

operation.<br />

Besides the specific technical issues discussed<br />

above, there is a more fundamental question concerning<br />

possible nanoelectronic applications of<br />

carbon nanotubes: What is the best direction to<br />

pursue for alternate information processing technologies,<br />

for example, carbon nanotubes, molecular<br />

electronics, etc.? Today’s attempts in most<br />

cases replicate silicon technology with new<br />

switches, or integrate nonsilicon components, such<br />

FIGURE 73. Ouput characteristics of a ring oscillator made of three complementary nanotube<br />

inverters. The output frequency is 220 Hz. V DD = -4 V. (Reprinted from Ref. 349, Copyright 2002,<br />

with permission from American Chemical Society.)<br />

342


as carbon nanotubes with silicon, for example,<br />

CMOS channel replacement technologies. However,<br />

as long as electron transport and energy<br />

barriers govern device operation, use of carbon<br />

nanotubes to replace the channel of a silicon<br />

MOSFET would not measurably extend silicon<br />

CMOS technology.<br />

Three important parameters in the realization<br />

of digital systems are the device switching speed,<br />

switching energy dissipation, and integration density.<br />

The question that should be examined is<br />

“What are the ultimate limits to the speed, size,<br />

density and dissipated energy of a carbon<br />

nanotubes switch (e.g., a FET switch)?<br />

In addition, several issues need to be addressed<br />

to assess the practical feasibility of nanotube based<br />

integrated electronics. Possibilities for integration<br />

of individual carbon nanotubes components<br />

in a complex circuit (billions of components per<br />

cm 2 ) are unclear at this point.<br />

D. Medical Applications of Fullerene-<br />

Based Materials<br />

The field of fullerene and carbon nanotube<br />

biology, as well as applications of fullerene derivatives<br />

in biology and medicine, has become increasingly<br />

popular. 298,299 In the mid-1990s it was demonstrated<br />

that fullerene compounds have biological<br />

activity and their potential as therapeutic products<br />

for the treatment of several diseases had been reported.<br />

Recently, a private biopharmaceutical company,<br />

C Sixty Inc. was created with a primary<br />

focus on the discovery and the development of a<br />

new class of therapeutics based on the fullerene<br />

molecule. C Sixty’s lead products are based on the<br />

modification of the fullerene molecule and are<br />

aimed at the treatment of cancer, AIDS, and<br />

neurodegenerative diseases. 301<br />

The flexible chemical reactivity of C 60 has<br />

already resulted in numerous fullerene compounds<br />

that are now available for study. In addition, at 7.2<br />

Å in diameter, C 60 is similar in size to steroid<br />

hormones or peptide alpha-helices, and thus<br />

fullerene compounds are ideal molecules to serve<br />

as ligands for enzymes and receptors. 298 While<br />

fullerene C 60 itself shows no solubility in water,<br />

many fullerene compounds can be very water<br />

soluble. Such derivatives of C 60 contain polar side<br />

chains, and the water solubility increases with the<br />

number of polar groups. A number of useful<br />

fullerene-based therapeutics have been reported,<br />

including antiviral agents and anti-cancer drugs, as<br />

well as biosensors for diagnostic applications; 300,301<br />

as a protective agent against iron-induced oxidative<br />

stress; 306 or as an in vitro antibacterial agent. 307<br />

Fullerenes had been demonstrated to be useful<br />

in DNA-templated assemblies of inorganic-organic<br />

building blocks. 312 The fullerene-DNA complexation<br />

significantly altered the structure of DNA<br />

(DNA become very condensed). This complexe<br />

might be potentially useful for gene delivery.<br />

The exploration of nanotubes in biomedical<br />

applications is also underway. Cells have been<br />

shown to grow on nanotubes, so they have no<br />

toxic effect. 313 The cells also do not adhere to the<br />

nanotubes, potentially giving rise to applications<br />

such as coatings for prosthetics, and antifouling<br />

coatings for ships. The ability to chemically modify<br />

the sidewalls of nanotubes can be considered for<br />

biomedical applications such as vascular stents,<br />

and neuron growth and regeneration. 313<br />

Thus, all major forms of carbon at the<br />

nanoscale appear to be valuable resources for<br />

biomedical applications.<br />

E. Atomic Modeling of <strong>Carbon</strong><br />

<strong>Nanostructures</strong> as a Tool for Developing<br />

New Materials and Technologies<br />

The results of a number of modeling and simulation<br />

studies on carbon nanostructures have been<br />

discussed in appropriate places elsewhere in this<br />

article. For completeness, we include in this section<br />

a brief discussion of some situations in which<br />

theory and modeling has led to experiments in<br />

developing new technologies and applications involving<br />

carbon nanostructures. In some cases the<br />

experiment has confirmed the theoretical predictions,<br />

while in other cases the corresponding experimental<br />

measurements have yet to be made.<br />

1. Fullerene Structures<br />

Probably the most celebrated triumph of theory<br />

in the area of carbon structures is the prediction<br />

that the electronic properties of carbon nanotubes,<br />

343


specifically the absence or presence of a band<br />

gap, depends on the tubule’s helical structure and<br />

radius. 199,280,353,354 Using simple tight binding<br />

theory based on near-neighbor π bonding, it can<br />

be shown that when the relation (2n 1 +n 2 )=3q is<br />

satisfied, a nanotube is predicted to have a zero<br />

band gap, with all other structures being semiconductors.<br />

In this expression, n 1 and n 2 are the integer<br />

values of the in-plane graphite lattice vectors<br />

that define the wrapping of a nanotube, and q is an<br />

integer. Further analysis shows that with the exception<br />

of the n 1 =n 2 structures, the zero band gaps<br />

predicted by simple tight binding theory are due<br />

to a degeneracy of the highest occupied molecular<br />

orbital (HOMO) and the lowest unoccupied molecular<br />

orbital (LUMO) at the Gamma point, that<br />

is, these structures are semimetals. More detailed<br />

calculations show that s-p mixing removes this<br />

degeneracy and introduces a band gap, albeit with<br />

a value that is much smaller than the structures<br />

predicted to be semiconductors by simple tight<br />

binding theory, and that the magnitude of the<br />

small band gap decreases with increasing nanotube<br />

radius. 354 In the case of the n 1 =n 2 structures, theory<br />

has shown that the zero band gap is due to a<br />

crossing of the HOMO and the LUMO, resulting<br />

in a true metal.<br />

Based on the predicted nanotube wrappingband<br />

gap prediction, several groups have suggested<br />

that joining two nanotubes with similar<br />

radii but different helical structures could create a<br />

metal-semiconductor junction. 355,356 Calculations<br />

have shown that such a junction is energetically<br />

feasible without introducing undesirable radical<br />

electronic states. Similarly, theory proposed that<br />

the electronic properties of nanotubes could be<br />

tuned further by chemisorbing species to their<br />

walls. 95,357 An example structure of this type is<br />

illustrated in Figure 74, where ethylene molecules<br />

are chemisorbed along part of a (6,0) nanotube at<br />

positions that are predicted to open a band gap.<br />

Based on tight binding calculations, the structure<br />

shown is predicted to have a Schottky barrier of<br />

0.44 eV, and metal-induced gaps states that decay<br />

from the metallic to the semiconducting region of<br />

the nanotube. 95,357 The decay behavior of these<br />

states is quantitatively similar to those at more<br />

conventional interfaces such as Al-Si boundaries.<br />

White and Todorov have predicted that<br />

nanotubes display ballistic electron conduction,<br />

with localization lengths of the order of hundreds<br />

of nanometers (or more) .358 This is exceptional<br />

behavior, which strongly suggests applications of<br />

nanotubes as efficient nanoscale wires. Additional<br />

FIGURE 74. Illustration of a region of a nanotube containing a metal-semiconductor junction due<br />

to ethylene chemisorption.<br />

344


details regarding transport in nanotubes can be<br />

found in Ref. 359.<br />

Theory has also predicted that the value of a<br />

band gap in a semiconducting nanotube can be<br />

changed by an applied stress, and that band gaps<br />

can even be added or removed for large<br />

stresses. 357,360,361 This has led to suggestions of<br />

nanoscale strain and vibration gauges using<br />

nanotubes as sensors in which changes in conductivity<br />

due to strain are utilized. 357 Structures with<br />

this specialized function have not yet been constructed.<br />

Recently, Srivastava and co-workers have<br />

modeled the structure and electron transport<br />

through both symmetric and asymmetric nanotube<br />

“Y junctions”. 362,363 These structures, which when<br />

appropriately connected resemble neuro-networks,<br />

were found to be energetically stable and electronically<br />

robust with respect to defect states. The<br />

transport calculations suggest that switching and<br />

rectification properties of these structures depend<br />

strongly on the symmetry, and to a lesser degree<br />

on chirality. For symmetric junctions, for example,<br />

the calculations show that perfect rectification<br />

across a junction is possible depending on the<br />

helical structure of the nanotubes, but that for<br />

asymmetric junctions rectification is not possible.<br />

Recent simulations and experiments have indicated<br />

that nanotube junctions can be created by<br />

electron beam heating of cross nanotubes. 364 Taken<br />

together, the predictions relating band gaps to<br />

structure and chemisorption, ballistic electron<br />

transport models, and the apparently rich behavior<br />

and structures possible with Y junctions makes<br />

a strong case for using these structures for<br />

nanoscale electronic device applications, with the<br />

caveats mentioned above. 365<br />

As discussed above, molecular modeling in<br />

conjunction with continuum treatments of<br />

nanotubes showed that severely bent nanotubes<br />

will kink (much like a garden hose), and that kink<br />

formation is largely reversible as nanotubes are<br />

straightened. 366,235 The kink structures observed<br />

in these early simulations were remarkably similar<br />

to those subsequently seen with electron and<br />

atomic force microscopies. 366,367 Both indicate a<br />

flattened region behind the kink, and ridges along<br />

the top and the bottom of the kink. Reversible<br />

kink formation has also been characterized theoretically<br />

for nanotubes used as molecular<br />

indentors, 368,369 a property that supports experimental<br />

applications of nanotubes in nanometerscale<br />

metrology. 370 Similar to kink formation,<br />

theory predicted that nanotubes with large radii<br />

will collapse into ribbon structures due to van der<br />

Waals attraction between opposing walls of the<br />

nanotube. 371,372 These structures have been observed<br />

experimentally. 372,302<br />

An interesting example of theory and experiment<br />

working together is the prediction that kinks<br />

in nanotubes may act as sites of enhanced reactivity.<br />

303 Using molecular modeling and a tight binding<br />

model, theory predicted that the ridges along<br />

the top and bottom of kinks on nanotubes produce<br />

atomic geometries that resemble sp 3 bonding,<br />

which result in the formation of radical states in<br />

the band gap of a (17,0) nanotube. Furthermore,<br />

these radical states, theory predicted, are sites for<br />

strong chemisorption of hydrogen atoms, and<br />

therefore kinks should display enhanced reactivity.<br />

Experimental studies carried out in parallel<br />

with the calculations showed that kink sites are<br />

more susceptible to attack by dilute nitric acid<br />

than are nonkinked regions of nanotubes. From a<br />

technology viewpoint, these results suggest that<br />

combining mechanical bending with chemical<br />

interactions is a viable route for controlling the<br />

cutting of nanotubes to precise lengths for various<br />

electronic device applications.<br />

Several modeling studies have been carried<br />

out to explore the properties of nanotube-polymer<br />

composites, particularly how load transfer between<br />

the nanofiber and matrix can be maintained.<br />

Molecular modeling has suggested that polymers<br />

can optimize nonbonded interactions with a<br />

nanotube by wrapping helically around the<br />

nanotubes. 304 Molecular modeling studies have<br />

also shown that chemical cross-linking between a<br />

nanotube and a polymer matrix can significantly<br />

enhance load transfer, 305 and that somewhat surprisingly<br />

such cross links may not significantly<br />

decrease the high modulus that make nanotubes<br />

attractive for composite applications in the first<br />

place. 95<br />

Molecular modeling studies using classic, tight<br />

binding, and first principles atomic forces have<br />

also been used to characterize the plastic deformation<br />

of uniaxially strained nanotubes. 66,256 These<br />

345


calculations predicted a Stone-Whales transformation<br />

that leads to what can be interpreted as<br />

dislocation formation and motion. Because the<br />

kink motion involves bond rotation, and bonds<br />

can have different orientations with respect to the<br />

nanotube axis for different helical wrappings, the<br />

calculations predicted that the strength of<br />

nanotubes depends strongly on their structure.<br />

Interestingly, the formation of a dislocation results<br />

in an interface between two nanotubes with<br />

different helical properties but similar radii, which,<br />

depending on the structures of the nanotube, can<br />

result in a metal-semiconductor junction like that<br />

discussed above. Hence, these calculations lead<br />

to predictions for specific conditions under which<br />

these junctions could potentially be made, as well<br />

as predictions of structures with optimized plastic<br />

properties for structural materials applications.<br />

Modeling has indicated that another potentially<br />

important application of nanotubes is the<br />

separation of organic molecular mixtures. 70 Simulations<br />

of diffusive flow of methane/isobutene,<br />

methane/n-butane, and methane/ethane binary<br />

mixtures through single nanotubes and nanotube<br />

bundles were carried out. The simulations showed<br />

effective separation of the butane structures from<br />

methane, but that because of the similarity in size<br />

separation of methane and ethane was as effective.<br />

The simulations also showed that the helical<br />

structure of nanotubes has little effect on the diffusion<br />

coefficients and hence the separation properties,<br />

but that the radius had a large effect with<br />

smaller radii structures showing better discrimination<br />

between different molecules.<br />

Molecular modeling and tight binding methods<br />

have also been used to characterize the structural<br />

and electronic properties of fullerene<br />

nanocones and nanostructures assembled from<br />

nanocones. 352 The calculations suggested that depending<br />

on their radius, these carbon nanocones<br />

can exhibit conventional cone shapes or can form<br />

concentric wave-like metastable structures. Furthermore,<br />

the calculations showed that single<br />

nanocones can be assembled into extended twodimensional<br />

structures that can be in a self-similar<br />

fashion with fivefold symmetry. Predicted electronic<br />

properties of nanocones indicated that the<br />

pentagon in the center of a cone is the most probable<br />

spot for electron field emission, suggesting<br />

the application of these structures as localized<br />

electron sources for templating at scales below<br />

more traditional lithographies.<br />

2. Nanodiamond Clusters<br />

The high symmetry of nanotubes helped to<br />

facilitate much of the successful theory and modeling<br />

done on these and related systems. In the<br />

case of nanodiamond clusters, there is a much<br />

greater variation in structure and perhaps fewer<br />

clear-cut technological applications, and so less<br />

has been done in terms of theory leading applications.<br />

Galli and co-workers used accurate first principles<br />

methods to characterize the dependence of<br />

band gap on the size of hydrogen-terminated nanodiamond<br />

clusters, as well as to predict surface<br />

reconstructions of clusters without chemisorbed<br />

hydrogen that are unique to these clusters. 90 The<br />

calculations indicate that for the hydrogen-terminated<br />

structures the HOMO-LUMO gap, which is<br />

given as 8.9 eV for methane, becomes comparable<br />

to the band gap for bulk diamond for clusters<br />

as small as 1 nm. This result is in sharp<br />

contrast to silicon and germanium clusters, which<br />

show quantum confinement effects for larger clusters.<br />

The lack of quantum confinement effects in<br />

diamond clusters is consistent with experimental<br />

emission and adsorption studies. The first principles<br />

calculations also predicted surface reconstructions<br />

without hydrogen termination that had<br />

significant π character. These structures resemble<br />

diamond clusters encapsulated in fullerenes and<br />

have been termed “bucky diamond” clusters.<br />

While not necessarily suggesting new technological<br />

applications of these structures, the results of<br />

these studies do point toward potential limitations<br />

and opportunities for using nano-diamond clusters<br />

for nanoelectronics applications.<br />

Modeling has also been used to characterize<br />

the bonding and stability of hybrid nanodiamondnanotube<br />

structures. 278 The modeling studies,<br />

which included both a many-body classic potential<br />

and tight-binding calculations, showed that<br />

several classes of hybrids exist with reasonably<br />

low-strain structures interfaces. There are several<br />

possible applications of these systems, including<br />

346


as arrays of field emitters (Figure 75) and as<br />

nanoscale diodes (Figure 76). While there are<br />

indications that such structures have been realized<br />

experimentally, deliberate experimental studies<br />

aimed at technological applications of these<br />

structures have not to date been carried out.<br />

In an interesting set of calculations, Park,<br />

Srivastava, and Cho used first principles calculations<br />

to examine a 31 P atom positioned at the<br />

center of a diamond nanocrystallite as a solidstate<br />

binary qubit. 195 The calculations indicated<br />

that the impurity is much more stable at a substitutional<br />

site than an interstitial site, and that the<br />

placement of the impurity is robust with respect<br />

to diffusion. The calculations also indicated that<br />

coupling between the nuclear spin and the weakly<br />

bound (valance) donor electrons is suitable for<br />

single qubit applications. This result, together with<br />

the stability of nanotube-diamond hybrid structures,<br />

suggests that nano-diamond clusters could<br />

be useful in computing applications.<br />

V. CONCLUSIONS AND FUTURE<br />

OUTLOOK<br />

It is an interesting period in the science of<br />

carbon, when the different communities such as<br />

those conducting research on graphite-based materials,<br />

fullerene nanotubes, and diamond, which<br />

were working rather independently and are now<br />

merging into one discipline as their interests converge<br />

at the nanoscale. There is a clear tendency<br />

to understand the properties of all-carbon entities<br />

at the nanoscale within a unified framework, including<br />

interrelationships between the various<br />

forms of nanocarbon, conditions under which one<br />

form transforms to another, and the possibility of<br />

FIGURE 75. Illustration of a “mushroom” array of nanodiamond-tubule hybrid structures.<br />

347


FIGURE 76. Illustration of a resonance tunneling diode created from a nanodiamondfullerene<br />

hybrid structure.<br />

combining nanocarbon entities into hierarchical<br />

structures.<br />

There is increased research activity in the thermodynamics<br />

and kinetics of carbon at the nanoscale.<br />

Ab initio level of sophisticated computer simulations<br />

outlined the sequence of the most stable forms<br />

of carbon at the nanoscale, which are rings,<br />

fullerenes, diamond clusters, and graphite particles<br />

as the system size is increased. Interesting forms of<br />

nonhydrogenated nanodiamond such as bucky-diamond<br />

have been discovered, and experimental<br />

confirmation has been provided. Presumably, there<br />

are enough data accumulated for the development<br />

of the three-dimensional carbon phase diagram at<br />

the nanoscale, where the third parameter is ths size<br />

of the carbon entity. Tremendous tasks remain to<br />

be done in order to develop an understanding of the<br />

general mechanisms of the nucleation and the<br />

growth of carbon nanospecies under a variety of<br />

conditions as well as the kinetics of transformation<br />

between different carbon forms.<br />

Special attention in the current review had<br />

been devoted to nanodiamond, which has very<br />

diverse structures at the nanoscale, ranging from<br />

individual clusters to high-purity films. As discussed<br />

in the review, these nanodiamond forms<br />

can be produced by very diverse techniques, ranging<br />

from the detonation explosive method to the<br />

low-pressure CVD method or by the method of<br />

chlorinating carbides. It is interesting that<br />

ultradispersed diamond has a long history of application<br />

in the countries of FSU, before<br />

nanotechnology became a popular topic, in such<br />

traditional areas as galvanic coatings, polymer<br />

composites, polishing, and additions to lubricants.<br />

It can be easily produced in ton quantities; however,<br />

it is very hard to produce UDD completely<br />

free of contamination due to the incombustible<br />

impurities (metals, nonmetals, and their oxides)<br />

that are located in interparticle areas within agglomerates<br />

of UDD particles. A high purity of<br />

UDD is not required, however, for the traditional<br />

applications mentioned above. Depending on the<br />

application, earlier developed technologies for<br />

UDD purification, surface modification as well as<br />

aggregate-free UDD suspensions are being developed<br />

for industrial-scale use. It should be emphasized<br />

that developing an understanding of the<br />

relationship between the complex structure of the<br />

surface of UDD and their physical properties are<br />

areas of recent very active research. 384 Especially<br />

important, although very challenging, is developing<br />

an understanding of the properties of individual<br />

nanodiamond particles. Regarding novel<br />

applications of UDD, we pointed out their biomedical<br />

applications, an area of research that has<br />

just started to emerge.<br />

The properties of carbon nanotubes as well<br />

as their prospective applications have been the<br />

focus of a variety of reviews and books. In the<br />

348


present work, we discussed carbon nanotube<br />

mechanical properties in the context that they<br />

are very instructive test systems with which<br />

new concepts for describing the mechanical<br />

behavior of nanostructures are required. This is<br />

being addressed by an emerging discipline —<br />

nanomechanics. We also discussed the prospects<br />

of the application of carbon nanotubes in<br />

field emission and nanoelectronic devices.<br />

While carbon nanotubes are very interesting<br />

research objects for materials science and solidstate<br />

physics, their potential electronic applications<br />

may be somewhat overstated. We believe<br />

that a more critical assessment of the potential<br />

of carbon nanotubes for electronics is needed.<br />

To emphasize the increasing role of atomic<br />

modeling of carbon nanostructures as a tool<br />

for developing new materials and technologies<br />

we included a brief discussion of some<br />

situations in which theory and modeling has<br />

led to experiments in developing new technologies<br />

and applications involving carbon<br />

nanostructures.<br />

In conclusion, the carbon family of materials<br />

at the nanoscale, with its wide diversity of forms,<br />

is a rapidly developing area from both the point of<br />

view of fundamental research as well as the current<br />

and perspective nanotechnological applications.<br />

ACKNOWLEDGMENTS<br />

The authors greatly acknowledge the help of<br />

Gary E. McGuire and John Hren for the critical<br />

reading of the manuscript; A. Barnard and G.<br />

Galli for providing their results before publication;<br />

as well as V. Kuznetsov, V. Dolmatov, A.<br />

Koscheev, A. Kalachev, T. Daulton, Y. Gogotsy,<br />

D. Gruen, D. Areshkin, D. Roundy, R. Ruoff, F.<br />

Ree, J. Viecelli, C. Pickard, V. Harik for very<br />

fruitful discussions. OAS acknowledges support<br />

from the Office of Naval Research through contract<br />

N00014-95-1-0270. DWB acknowledges<br />

support from the Office of Naval Research through<br />

contract N00014-95-1-0270 and through a subcontract<br />

from the University of North Carolina at<br />

Chapel Hill, and from NASA-Ames and NASA-<br />

Langly.<br />

REFERENCES<br />

1. V. Y. Dolmatov, Russian Chemical Reviews 70, 607–<br />

626 (2001).<br />

2. D. M. Gruen, Annual Rev. Mater. Sci. 29, 211–259<br />

(1999).<br />

3. A. L. Vereschagin, Detonation Nanodiamonds (Altai<br />

State Technical University, Barnaul, Russian Federation,<br />

2001), in Russian.<br />

4. M.Inagaki, New carbons, Elsevier, 2000.<br />

5. Carbyne and Carbynoid Structures, ed. By R. B.<br />

Heimann, S. E. Evsyukov and L. Kavan, Kluwer<br />

Academic Pub., 1999.<br />

6. Nanostructured <strong>Carbon</strong> for Advanced Applications,<br />

edited by G. Benedek, P. Milani and V. G. Ralchenko,<br />

NATO Science Series vol. 24, Kluwer Acad. Publ.,<br />

2001.<br />

7. Dresselhaus MS, Endo M, Relation of carbon<br />

nanotubes to other carbon materials, Top Appl Phys<br />

80: 11–28 2001.<br />

8. Shekhar Subramoney, Advanced Materials 10, 1157–<br />

1171, 1998.<br />

9. Dresselhaus MS, Dresselhaus G, Nanostructructured<br />

Materials 9 (1–8): 33–42 1997.<br />

10. M. Dresselhaus, p. 9–27, in Supercarbon, Synthesis,<br />

Properties and Applications, S. Yoshimura, R. P. H.<br />

Chang, Eds., Springer, 1998.<br />

11. Collin G, CFI-Ceramic Forum Intern. 77, 28–35, 2000.<br />

12. H. W. Kroto, J. R. Heath, S. C. O’Brien, R. F. Curl,<br />

R. E. Smalley, Nature 318, 162–163,1985.<br />

13. S. Iijima, Nature 354, 56–58, 1991.<br />

14. Gogotsi Y., S. Welz, D. A. Ersoy, and M. J. McNallan,<br />

Nature 411, 283–287 2001.<br />

15. R. Hazen, The Diamond Makers, Cambridge Univ.<br />

Press, 1999.<br />

16. P. DeCarli J. Jamieson, Science 133,1821, 1961; P. S.<br />

de Carli, US Patent 3,238,019, 1966.<br />

17. Volkov K. V., Danilenko V. V., Elin V. I., Fizika<br />

Gorenia i Vzriva (Russian) 26, 123–125, 1990.<br />

18. N. R. Greiner, D. S. Phillips, J. D. Johnson, and F.<br />

Volk, Nature 333, 440, 1988.<br />

19. Dresselhaus MS, Dresselhaus G, Saito R., Nanotechnology<br />

in <strong>Carbon</strong> Materials, in Nanotechnology, ed. G. Timp,<br />

AIP Press, p. 285–331, 1999.<br />

20. Osawa E., Yoshida M., Fujita M., MRS Bull. 19(11) p.<br />

33, 1994.<br />

21. Cataldo F., <strong>Carbon</strong> 40, 157–162, 2002.<br />

22. Heinmann R. B., Evsyukov S. E., and Koga, Y., <strong>Carbon</strong><br />

35 (1997), 1654–1658.<br />

23. Supercarbon, Synthesis, properties, and Applications,<br />

Eds. S. Yoshimura, R. P. H. Chang, Springer, 1998.<br />

24. Session titled ‘The third form of carbone: fullerenes<br />

— properties and prospects, in ASTM/TSM materials<br />

week ’92, Chicago, Il, Nov 2–5, 1992, ASM international,<br />

Materials Park, OH, 1992.<br />

25. Kudryavtsev Y. P, Evsyukov, S. E, et al., Carbyne—<br />

the third allotropic form of carbon, Russ. Chem. Bull.<br />

42, 399–413, 1993.<br />

349


26. Lagow R. J., Kampa J. J., et al. Synthesis of linear<br />

acetylenic carbon: the “sp” carbon allotrope, Science<br />

267, 362–367, 1995.<br />

27. Amorphous <strong>Carbon</strong>: State of the Art, Ed. S. Silva et<br />

al., World Scientific, 1997.<br />

28. Haddon R. C., Phil. Trans. Roy Soc. London, Ser. A<br />

343, 53, 1993.<br />

29. Matsumoto, S., Matsui, Y., J. Mater. Sci. 18, 1785–<br />

1791, 1983.<br />

30. Stein S. E., Nature 346, 517–521, 1990.<br />

31. Gogotsi Y. et al., Science 290, 317–320, 2000.<br />

32. Zhu HW, Xu CL, Wu DH, Wei BQ, Vajtai R, Ajayan<br />

PM, Science 296, 884–886 2002.<br />

33. Baik, E. -S. ; Baik, Y. -J. ; Jeon, Dongryul. Journal<br />

of Materials Research 15, 923–926, 2000; Baik, E. -<br />

S. ; Baik, Y. -J. ; Lee, S. W. ; Jeon, D., Thin Solid<br />

Films (2000), 377–378 295–298. .<br />

34. Bacon R., J. Appl. Phys. 31, 283,1960.<br />

35. Jones D. E. H., New Scientist 36, 1966.<br />

36. Osawa E., Kagaku, 25, 854, 1970 (in Japanese).<br />

37. GBochvar D. A., Galperin E. G., Dokl. Acad. Sci.<br />

SSSR 209, 610, 1973.<br />

38. Iijima S., J. Microscopy 119, 99, 1980.<br />

39. Harris P., <strong>Carbon</strong> Nanotubes and Related Structures,<br />

Cambridge University Press, 1999.<br />

40. Paquette L. A. et al., J. Am. Chem. Soc. 105, 5446, 1983.<br />

41. Rohlfling E. A., et al., J. Chem. Phys. 81, 3322, 1984.<br />

42. W. Krätschmer, L. D. Lamb, K. Fostiropoulos, D. R.<br />

Huffman, Nature 347, 354–358, 1990.<br />

43. B. J. Derjaguin and D. V. Fedoseev, Sci. Am. 233,<br />

102, 1975.<br />

44. Curl R. F., Smalley R. E., Scientific American, October,<br />

p. 54, 1991.<br />

45. C. N. R. Rao, R. Seshadri, A. Govindaraj, R. Sen,<br />

Mater. Sci. Eng. R 15, 209–262, 1995.<br />

46. D. Ugarte, Nature 359, 707–709, 1992.<br />

47. Iijima, T. Ichihashi, Nature 363, 603–605,1993.<br />

48. D S Bethune, C H Kiang, M S DeVries, G Gorman,<br />

R Savoy and R Beyers, Nature 365,605, 1993.<br />

49. Thess A, R. Lee, P. Nikolaev et al., Science 273, 483,<br />

1996.<br />

50. Ren Z. F. et al., Science 282, 1105, 1998.<br />

51. R. R. Schilttler, J. W. Seo, J. K. Gimzewski, C. Durkan,<br />

M. S. M. Saifullah, and M. E. Welland, Science 292,<br />

1136–1138, 2001.<br />

52. Becker L., Bada J. et al. Science 265, 642–645, 1994.<br />

53. Barnard A. S., The Diamond Formula, Butterworth-<br />

Heinmann, 2000.<br />

54. Covan, J. R.,. Dunnington, B. W., Holzman, US Patent<br />

3,401,019, 1968.<br />

55. Staver A. M., Gubareva N. M. Lyamkin A. I., Petrov<br />

E. A Fizika Gorenia i Vzriva (Rus) 1984, v. 20, pp<br />

440–442; Staver A. M., et al. patent of USSR/No.<br />

1165007 (1982).<br />

56. Baum F., L. P. Orlenko, K. P. Stanyukovich, V. P.<br />

Chelyshev, B. I. Shekhter, “Fizika vzryva”/`Physics<br />

of Explosion`/, “Nauka” Publishers, Moscow, pp. 97–<br />

125, 145–152, 1975.<br />

57. Abadurov, et al. United States Patent 4,483,836, 1984.<br />

58. Lewis R. S., Ming T., Wacker J. E. et al., Nature 326,<br />

160–162, 1987.<br />

59. Vereschagin, et al. United States Patent 5,861,349;<br />

Azarova OA, et al. Russian Patent 2056158 (1996).<br />

60. Kuznetsov V. L., Chuvilin A. L., Butenko Y. V. et al.,<br />

Chem. Phys. Lett. 209, 72, 1994.<br />

61. Banhart and P. M. Ajayan, Nature 382, 433, 1996.<br />

62. Gruen D. M., Liu S., Krauss A. et al., J. Appl. Phys.<br />

75, p1758, 1994; Appl. Phys. Let. 64, p. 1502, 1994.<br />

63. Y. G. Gogotsi, J. D. Jeon, M. J. McNallan, J. Mater.<br />

Chem. 7, 1841–1848, 1997.<br />

64. “Ximia i zizn” (“Chemistry and Life”), scientific popular<br />

journal (in Russian), 1, p 14–16, 1999 (about work<br />

of Koscheev A.).<br />

65. H. Prinzbach, A. Weiler et al., Nature 407, 60–63,<br />

2000.<br />

66. M. B. Nardelli. B. I. Yakobson and J. Bernholc, Phys.<br />

Rev. Lett. 81, 4656 (1998).<br />

67. Robertson,-D. -H. ; Brenner,-D. -W. ; Mintmire,-J. -<br />

W., Phys. Rev. B 45, 12592–5, 1992.<br />

68. T W Ebbesen and P M Ajayan, Nature 358, p220,<br />

1992.<br />

69. Peng l. M., Zhang Z. L., Xue Z. Q. et al., Phys. Rev.<br />

Let. 85, 3249–3252, 2000.<br />

70. Z. G. Mao and S. B. Sinnott, J. Phys. Chem. B 105<br />

6916 (2001).<br />

71. Affoune AM, Prasad BLV, Sato H, Enoki T, Kaburagi<br />

Y, Hishiyama, Chem. Phys. Lett. 348, 17–20, 2001.<br />

72. Nuth J. A. III, Nature 329, 589, 1987.<br />

73. Badziag P., Verwoerd W. S et al., Nature 343, 244–<br />

245, 1990.<br />

74. Frenklach M, Kematik RHuang D, et al., J. Appl.<br />

Phys 66, 395–399, 1989.<br />

75. Fedoseev,-D. -V. ; Lavrent’ev,-A. -V. ; Deryagin,-B.<br />

-V. Colloid Jour. 42, 592–5, 1980.<br />

76. E. F. Chaikovskii, G. Rosenberg, Dokl. Acad. Nauk<br />

SSSR 279, 1372, 1984.<br />

77. Amlof J., Luthi H. P., Supercomputer Research in<br />

Chemistry and Chemical Enguneering, ACS Symposium,<br />

ASC, 1987.<br />

78. M. Y. Gamarnik, Phys. Rev. B 54, 2150–2156, 1992.<br />

79. Winter N. W., Ree F. H., J. Comp. Aided Design 5,<br />

279, 1998.<br />

80. Ree FH, Winter NW, Glosli JN, et al. Physica B 265,<br />

223–229, 1999.<br />

81. Barnard A. S, Russo S. P., Snook I. K., J. Chem. Phys.<br />

submitted.<br />

82. Kuznetsov V. L., Zilberberg I. L., et al., J. Appl. Phys.<br />

86, 863 1999.<br />

83. Fugaciu F., Hermanna H, Seifert G., Phys. Rev. B, 60,<br />

pp. 10711–10714, 1999.<br />

84. Hermanna H., Fugaciu F., Seifert G., Appl. Phys. Lett.<br />

79, pp. 63–65, 2001.<br />

85. Park N, Lee K, Han SW, Yu JJ, Ihm J, Phys. Rev. B<br />

65, 121405, 2002.<br />

86. Viecelli JA, Ree FH, J. Appl. Phys. 88, 683–690,<br />

2000.<br />

350


87. Viecelli JA, Bastea S, Glosli JN, Ree FH, J. Chem.<br />

Phys. 115, 2730–2736,2001.<br />

88. Vereshchagin AL, Combust. Exp. Shock Waves 38,<br />

358–359, 2002.<br />

89. Ball K. D., Berry R. S., Kunz R. E. et al., Science 271,<br />

963, 1996.<br />

90. Raty J. Y., Galli G., Buuren T., in press.<br />

91. Bundy FP, Bassett WA, Weathers MS, et al., <strong>Carbon</strong><br />

34,141–153 1996.<br />

92. V. L. Kuznetsov, A. L. Chuvilin, Yu. V. Butenko, et<br />

al. in: Science and Technology of Fullerene Materials,<br />

edited by P. Bernuer, et al. (Mater. Res. Soc. Proc.<br />

359, Pittsburgh, PA, 1995), pp. 105–110, 1995.<br />

93. Zaiser M, Lyutovich Y, Banhart F Phys. Rev. B 62,<br />

3058–3064, 2000; Banhart F Phys. Sol. State 44, 399–<br />

404, 2002.<br />

94. Yu. V. Butenko, V. L. Kuznetsov, A. L. Chuvilin, V.<br />

N. Kolomiichuk S. V. Stankus, R. A. Khairulin, B.<br />

Segall, J. Appl. Phys. 88, 4380–4388, 2000.<br />

95. D. W. Brenner, O. A. Shenderova, D. A. Areshkin, J.<br />

D. Schall, Computer Modeling in Engineering and<br />

Sciences 3,643 (2002).<br />

96. Jiang Q, Li JC, Wilde G., J. Phys. C 26, 5623–5627, 2000.<br />

97. N. M. Hwang, J. H. Hahn, D. Yoon, J. Cryst. Growth<br />

162, 55–68, 1996.<br />

98. N. M. Hwang, J. H. Hahn, D. Yoon, J. Cryst. Growth,<br />

160, 87, 1996.<br />

99. Barnard A. S, Russo S. P., Snook I. K., Diam. Rel.<br />

Mat., in press.<br />

100. Barnard A. S, Russo S. P., Snook I. K., Surf. Rev.<br />

Lett., in press.<br />

101. Barnard A. S, Russo S. P., Snook I. K., Phil. Mag.<br />

Let., 83, 1, p. 39–45..<br />

102. Adams G., Sankey O., et al., Science 256, 1792–1795,<br />

1992.<br />

103. Tomanek D., Schluter M. A., Phys. Rev. Let. 67,<br />

2331–2335, 1991.<br />

104. J. L. Martins et al., J. Cluster Sci., p. 513, 2001.<br />

105. Jones R. O, J. Chem. Phys. 110, 5189, 1999.<br />

106. Churilov GN, Novikov PV, Tarabanko VE, Lopatin<br />

VA, Vnukova NG, Bulina NV, <strong>Carbon</strong> 40, 891–896<br />

2002.<br />

107. Lozovik YE, Popov AM, Phys. Sol. State 44,186–<br />

194, 2002.<br />

108. S. B. Sinnott R. Andrews, D. Qian, et al. Chem. Phys.<br />

Letters 315, 25–30, 1999.<br />

109. Lin EE, Phys. Sol. St. 42, 1946–1951, 2000.<br />

110. Kiang C. H., J. Chem. Phys. 113, 4763–4666, 2000.<br />

111. Frenklach M, HowardW., Huang D, et al., Appl. Phys.<br />

Lett. 59, 546–548, 1991.<br />

112. Mitura S, J. Cryst. Growth 80, 417–424, 1987.<br />

113. P. R. Buerki, S. Leutwyler. J. Appl. Phys. 69, p. 3739,<br />

1991.<br />

114. P. R. Buerki, S. Leutwyler. J. Nanostruct. Mater. 4, p.<br />

577, 1994.<br />

115. Z. Chang, X. Jiang, S. Qin, X. Zhang, Z. Xue, S. Pang,<br />

2d Inter. Conf. on Applic. of Diamond Films and<br />

Related Materials, Tokyo, Japan, 1993, p. 463.<br />

116. Hyo-S. Ahn, Hyun-M. Park, Doh-Y. Kim and Nong<br />

M. Hwang, J. Crystal Growth 234, 399–403, 2002.<br />

117. Daulton T. L., Eisenhour D. D., Bernatowiez T., et al.<br />

Geochim. Cosm. Acta, 60, p. 4853–4873, 1996.<br />

118. T. L. Daulton, Meteorite!, 5, p. 26–29, 1999.<br />

119. Clayton DD, Meyer BS, Sanderson CI, Astrophys J.<br />

447, 894, 1995.<br />

120. Ozima M. and Mochizuki K, Meteorite 28, 416, 1993.<br />

121. Tielens A, Seab C., Hollenbach, D etal., Astroph. J.<br />

319, L109, 1987.<br />

122. Daulton TL, Kirk MA, Lewis RS, Rehn LE, Nicl. Inst.<br />

Meth. Phys. Res. B 175, 12–20, 2001.<br />

123. P. Wesolowski, Y. Lyutovich, F. Banhart, H. D.<br />

Carstanjen and H. Kronmuller. Appl. Phys. Lett. 71, p.<br />

1948, 1998.<br />

124. Daulton T., Ozima M, Science 271, 1260–1263, 1996.<br />

125. Allamandola L. J. et al. Science 260, 64–66, 1993.<br />

126. Huss G. R., Lewis R. S., Geochim. Cosmochim Acta<br />

59, 115–160, 1995.<br />

127. Koscheev A. P., Gromov M. D. et al., Nature 412,<br />

616–819, 2001.<br />

128. Fedoseev V. D., Bukhovets V. L., Varshavskaya et<br />

al., <strong>Carbon</strong> 21, 237–241, 1983.<br />

129. Alam M, Debroy T, Roy R, et al., <strong>Carbon</strong> 27, 289–<br />

294 1989.<br />

130. Wei B, Zhang J, Liang J, Wu D., <strong>Carbon</strong> 36, 997–<br />

1001, 1998.<br />

131. Meguro T, Hida A, Suzuki M, Koguchi Y, Takai H,<br />

Yamamoto Y, Maeda K, Aoyagi Y, Appl. Phys. Lett.<br />

79, 3866–3868, 2001.<br />

132. S. Yugo, T. Kania, T. Kimuro and T. Muto, Appl.<br />

Phys. Let. 58, 1036, 1991.<br />

133. W. J. Zhang et al. Phys. Rev. B 61, 5579, 2000.<br />

134. Y. Lifshitz, S. Kasi, J. Rabalais, Phys. Rev. Lett. 62,<br />

1290, 1989.<br />

135. F. Banhart., Rep. Progr. Phys. 62 1181, 1999.<br />

136. Tomita S, Fujii M, Hayashi S, Yamamoto K, Diam.<br />

Rel. Mat. 9, 856–860, 2000.<br />

137. H. Liu, D. Dandy, Diam Rel. Mat. 4 1173–1188,<br />

1995.<br />

138. J. Butler, H. Windischmann, MRS Bull. 23 (No. 9),<br />

22, 1998.<br />

139. Lifshitz Y, Kohler T, Frauenheim T, et al., Science<br />

297, 1531–1533, 2002.<br />

140. S. T. Lee, H. Y. Peng, X. T. Zhou et al., Science 287,<br />

104–106, 2000.<br />

141. Gruen DM MRS Bull. 26, 771–776, 2001.<br />

142. S. Bhattacharyya et al., Appl. Phys. Lett. 79, 1441,<br />

2001.<br />

143. S Jiao Sumant A, Kirk MA, Gruen DM et al, J. Appl.<br />

Phys. 90,118, 2001.<br />

144. S. Welz, Y. Gogotsi, M. J. McNallan, submitted to J.<br />

Appl. Phys.<br />

145. Gogotsy, private communication.<br />

146. Gordeev S. K., in Nanostructured <strong>Carbon</strong> for Advanced<br />

Applications, eds. G. Benedek, P. Milani and<br />

V. G. Ralchenko, NATO Science Series vol. 24,<br />

Kluwer Acad. Publ., 2001, p. 71–88.<br />

351


147. K. Yamada, and Y. Tanabeb <strong>Carbon</strong> 40, 261–269,<br />

2002.<br />

148. Y. Q. Zhu1, T. Sekine, T. Kobayashi, E. Takazawa,<br />

M. Terrones and H. Terrones Chem. Phys. Lett. 287,<br />

689–693, 1998.<br />

149. R. B. Heinmann, p. 409–425, in Ref. 5<br />

150. Y. Z. Ma, G. T. Zou, H. B. Yang and J. F. Meng, Appl.<br />

Phys. Lett. 65, 822–823, 1994.<br />

151. Sundqvist B, Adv. Phys. 48(1), 1–134, 1999.<br />

152. Limin C., Cunxiao Gao, Haiping Sunc, Guangtian,<br />

<strong>Carbon</strong> 39, 311–314, 2001.<br />

153. US Patent 5,449,491 (1995).<br />

154. M. Núñez-Regueiro, P. Monceau and J. -L. Hodeau,<br />

Nature 355, 237–239, 1992.<br />

155. H. Yusa, Diam Relat. Mat. 11, 87–91, 2002.<br />

156. Zhang WJ, Sun XS, Peng HY, Wang N, Lee CS, Bello<br />

I, Lee ST Phys. Rev. B 61, 5579–5586, 2000.<br />

157. Heera V, Skorupa W, Pecz B, Dobos L, Appl. Phys.<br />

Lett. 76, 2847–2849, 2000.<br />

158. M. S. Shaw, J. D. Johnson, J. Appl. Phys. 62, 2080–<br />

2087, 1987.<br />

159. A. P. Ershov, A. L. Kupershtokh, Fiz. Gorenia Vryva<br />

27, 111, 1991.<br />

160. G. V. Sakovich, V. D. Gubarevich, F. Z. Badaev et<br />

al., Dokl. Acad. Nauk SSSR 310, 402,1990.<br />

161. A. V. Kurdyumov, N. F. Ostrovskaya, V. B. Zelyavskii<br />

et al., Sverkht. Mater. 23,1998.<br />

162. G. P. Bogatyreva, Y. I. Sozin, N. A. Oleynik, Sverkht.<br />

Mater. 5, 1998.<br />

163. Kuznetsov VL, Chuvilin AL, Moroz EM, et al., <strong>Carbon</strong><br />

32, 873–882, 1994.<br />

164. Aleksenskii AE, Baidakova MV, Vul’ AY, Siklitskii<br />

VI, Phys. Sol. State 41, 668–671 1999.<br />

165. Aleksenskii AE, Baidakova MV, Vul AY, Davydov<br />

VY, Pevtsova YA, Phys. Sol. State 39, 1007–1015,<br />

1997.<br />

166. Aleksenskii AE, Baidakova MV, Vul’ AY, Dideikin<br />

AT, Siklitskii VI, Vul’ SP, Phys. Sol. State 42, 1575–<br />

1578, 2000.<br />

167. M. Yoshikava, Y. Mori, H. Obata, M. Maegawa, G.<br />

Katagiri, H. Ishida, A. Ishitani, Appl. Phys. Lett. 67,<br />

694, 1995.<br />

168. A. L. Vereschagin and G. V. Sakovich, Mendeleev<br />

Commun. 1, 39–41, 2001.<br />

169. E. D. Obraztsova, V. L. Kuznetsov, E. N. Loubnin, S.<br />

M. Pimenov, V. G. Pereverzev, in Nanoparticles in<br />

Solids and Solutions, by J. H. Fendler and I. Dekany<br />

(eds) (1996 Kluwer Academic Publisher) p. 485–496.<br />

170. Y. K. Chang, H. H. Hsieh, W. F. Pong, et al. Phys.<br />

Rev. Lett. 82, 5377, 1998.<br />

171. L. Ley, J. Ristein, R. Graupner, Phys. Rev. Lett. 84,<br />

5679, 2000.<br />

172. D. A. Areshkin, O. A. Shenderova, V. V. Zhirnov, A.<br />

F. Pal, J. J. Hren, D. W. Brenner, Mat. Res. Soc.<br />

Symp. Proc. Vol. 621, R5. 16. 1, 2001.<br />

173. K. Suenaga, T. Tence, C. Mory, C. Colliex, H. Kato,<br />

T. Okazaki, H. Shinohara, K. Hirahara, S. Bandow, S.<br />

Iijima, Science 290, 2280, 2000.<br />

174. R. P. Fehlhaber and L. A. Bursill, Phil. Mag B 79,<br />

477, 1999.<br />

175. Q. Chen, S. Yun, Mater. Res. Bul. 35, 1915, 2000.<br />

176. P. W. Chen, Y. S. Ding, Q. Chen, F. L. Huang, S. R.<br />

Yun, Diam. Rel. Mat. 9 1722, 2000.<br />

177. E. Maillard-Schaller, O. M. Kuettel, L. Diederich, L.<br />

Schlapbach, V. V. Zhirnov, P. I. Belobrov, Diam.<br />

Related Mat. 8, 805, 1999.<br />

178. S. K. Gordeev, P. I. Belobrov, N. I. Kiselev, E. A.<br />

Petrakovskaya, T. C. Ekstrom, Mat. Res. Soc. Symp.<br />

Proc. Vol. 63 (2001) p. F14. 16. 1.<br />

179. P. I. Belobrov, S. K. Gordeev, E. A. Petrakovskaya,<br />

and O. V. Falaleev, Doklady Physics 46, 38, 2001.<br />

180. V. V. Zhirnov, O. M. Küttel, O. Gröning, A. N.<br />

Alimova, P. Y. Detkov, P. I. Belobrov, E. Maillard-<br />

Schaller, L. Schlapbach, J. Vac. Sci. Technol. 17, 666,<br />

1999.<br />

181. A. N. Alimova, N. N. Chubun, P. I. Belobrov, P. Y.<br />

Detkov, V. V. Zhirnov, J. Vac, Sci. Technol B 17, 715,<br />

1999.<br />

182. G. L. Bilbro, Diamond and Related Mat. 11, 1572, 2002.<br />

183. A. V. Kvit, T. Tyler, V. V. Zhirnov, J. J. Hren, to be<br />

published.<br />

184. T. Tyler, V. V. Zhirnov, A. V. Kvit, J. J. Hren, Abstr.<br />

62 nd Physical Electronics Conf., Atlanta,, GA June<br />

12–14 2002, to be published.<br />

185. A. Beveratos, R. Brouri, T. Gacoin, J-P. Poizat, P.<br />

Grangier, Phys. Rev. A 64, 1802, 2001.<br />

186. Kern G. and Hafner J., Phys. Rev. B 56, 4203–4210,<br />

1997.<br />

187. Iijima S. and Ichihashi, Nature 363, 603, 1993.<br />

188. Halicioglu T, Phys. Stat. Sol. B 199, 345–350, 1997.<br />

189. A. de Vita, G. Galli, A. Canning and R. Car, Nature<br />

379 523, 1996.<br />

190. Areshkin D., Shenderova O., Brenner D., unpublished.<br />

191. Areshkin D., Shenderova O., Brenner D., submitted.<br />

192. Ristein, J., Diam. Related Materials 9, 1129–1137,<br />

2000.<br />

193. Rutter, M. J. and Robertson, J., Phys. Rev. B 57,9241–<br />

9245, 1998.<br />

194. A. S. Barnard, S. P. Russo, I. K. Snook, J. Chem. Php.<br />

195. S. Park, D. Srivastava and K. Cho, J. Nanoscience<br />

Nanotech. 1, 75, 2001.<br />

196. Shenderova O., Brenner D., Ruoff R., submitted.<br />

197. Roundy, D. ; Cohen. M. L, Phys. Rev. B 64, pp.<br />

212103, 2001.<br />

198. Telling RH, Pickard CJ, Payne MC, Field JE, Phys.<br />

Rev. Lett. 84, 5160, 2000.<br />

199. M. S. Dresselhaus, G. Dresselhaus, and P. C. Eklund,<br />

Science of Fullerenes and <strong>Carbon</strong> Nanotubes, Academic<br />

Press, San Diego, 1996.<br />

200. <strong>Carbon</strong> Nanotubes, Preparation and Properties, edited<br />

by T. W. Ebbesen, CRC Press, 1996.<br />

201. Physical properties of carbon nanotubes, R. Saito, G.<br />

Dresselhaus & M. S. Dresselhaus, World Scientific, 1998.<br />

202. <strong>Carbon</strong> Nanotubes: Synthesis, Structure, Properties,<br />

and Applications, edited by M. S. Dresselhaus,G.<br />

Dresselhaus and P. Avouris, Springer-Verlag, 2001.<br />

352


203. M. S. Dresselhaus,G. Dresselhaus, K. Sugihara et al.,<br />

Graphite Fibers and Filaments, Springer-Verlag,<br />

Berlin, 1998.<br />

204. H. Dai, in Top. Appl. Phys. 80, 29–52, 2001.<br />

205. V. L. Kuznetsov, A. N. Usoltseva, A. L. Chuvilin et<br />

al., Phys. Rev. B 64, 235401, 2001.<br />

206. Journet C., Maser W.,Bernier P., Nature 388, 756–<br />

758, 1997.<br />

207. Liu J., Rinzler A., Dai H et al., Science 280, 1253–56,<br />

1998.<br />

208. Kong J., Soh H., Cassel A. et al., Nature 395, 878–<br />

879, 1998.<br />

209. Hafner J., Bronikowski M., Azamian B., et al., Chem.<br />

Phys. Lett. 296, 195–202, 1998.<br />

210. R. R. Bacsa, Ch. Laurent, A. Peigney, W. S. Bacsa,<br />

Th. Vaugien, A. Rousset, Chem. Phys. Lett. 323, 566,<br />

2000.<br />

211. Tans, et al., Nature 393, 49–52, 1998.<br />

212. B. Q. Wei, et al., Appl. Phys. Lett. 79, 1172–1176, 2001.<br />

213. SJ Tans, M H Devoret, H Dai, A Thess, R E Smalley,<br />

L J Geerligs and C Dekker, Nature 386, 474–476,<br />

1997.<br />

214. M. Kociak, A. Yu. Kasumov, S. Guéron et al., Phys.<br />

Rev. Lett. 86, 2416–2419, 2001.<br />

215. S. Sanvito, Y. Kwon, D. Tománek, C. J. Lambert,<br />

Phys. Rev. Lett. 84, 1974–1976, 2000.<br />

216. J. Hone, M. Whitney, A. Zettle, Synthetic Metals 103,<br />

2498, 1999.<br />

217. Berber S., Y. Kwon, D. Tomànek, Phys. Rev. Lett. 84,<br />

2000.<br />

218. Brenner, D. W., Shenderova, O. A., Harrison, J. A.,<br />

Stuart, S. J., Ni, B., Sinnott, S. B., J. Phys. Solid State<br />

14, 783–802, 2002.<br />

219. Peng, H. Y., Zhou, X. T., Lai, H. L., Wang, N., Lee,<br />

S. T. J. Mat. Res. 15, 2020–2026, 2000.<br />

220. Shi, W. S., Zheng, Y. F., Peng, H. Y., Wang, N., Lee,<br />

C. S., Lee, S. T., J. Am. Cer. Soc. 83, pp. 3228–3230,<br />

2000.<br />

221. Yakobson B and Avouris P., Mechanical properties of<br />

carbon nanotubes. In: Dresselhaus, M. S., Dresselhaus,<br />

G., Avouris, Ph. (Eds. ), Topics in Applied Physics,<br />

Springer–Verlag, Heidelberg, Germany, <strong>Carbon</strong><br />

Nanotubes, vol. 80, pp. 287–329, 2001.<br />

222. Yakobson in Handbook of Nanoscience, Engineering<br />

and Technology, CRC Press, 2002.<br />

223. D. Qian, G. J. Wagner, W. K. Liu, R. S. Ruoff, M. F.<br />

Yu, chap. 19 in Handbook of Nanoscience, Engineering<br />

and Technology, CRC Press, 2002.<br />

224. Sinnot, Andrews, Crit. Review. Sol. State Mater. Sci.<br />

p. 145–248, 2001.<br />

225. Erik T. Thostenson, Zhifeng Ren and Tsu-Wei Chou,<br />

Composites Science and Technology 61, 1899–1912,<br />

2001.<br />

226. Hu, J. Q., Lu, Q. K., Tang, K. B., Deng, B., Jiang, R.<br />

R., Qian, Y. T., Yu, W. C., J. Phys. Chem. B 104, pp.<br />

5251–5254, 2000.<br />

227. Zhou G, Duan WH, Gu BL, Chem. Phys. Lett. 333,<br />

344–349, 2001.<br />

228. Ribeiro FJ, Roundy DJ, Cohen ML, Phys. Rev. B 65,<br />

153401 2002.<br />

229. Kudin KN, Scuseria GE, Yakobson BI, Phys. Rev. B<br />

64, 235406, 2001.<br />

230. V. M. Harik, T. S. Gates, M. P. Nemeth, NASA/CR-<br />

2002–211460, ICASE Report No. 2002–7, 2002.<br />

231. G. M. Odegard, T. S. Gates, L. M. Nicholson, K. E.<br />

Wise, NASA TM-2001–210863, Hampton, VA, 2001.<br />

232. D. Sánchez-Portal, E. Artacho, and J. M. Soler, Phys.<br />

Rev. B 59, pp. 12678–12688, 1999.<br />

233. Krishman, E. Dujardin, T. W. Ebbesen, P. N. Yianilos,<br />

and M. M. J. Treacy, Phys. Rev. B 58, 14013, 1998.<br />

234. J. -P. Salvetat, J. -M. Bonard, N. H. Thomson, A. J.<br />

Kulik, L. Forro, W. Benoit, and L. Zuppiroli, Appl.<br />

Phys. A 69, 255, 1999.<br />

235. B. I. Yakobson, C. J. Brabec, and J. Bernholc, Phys.<br />

Rev. Lett. 76 (1996), p. 2511.<br />

236. C. Q. Ru Phys. Rev. B 62 15, 9973, 2000.<br />

237. S. Govindjee and J. L. Sackman, Solid State Commun.<br />

110, 227, 1999.<br />

238. V. M. Harik, Comp. Mat. Sci. 24, 312–326, 2002.<br />

239. P. Zhang, Y. Huang, P. H. Geubelle, P. A. Klein, and<br />

K. C. Hwang, Int. J. Sol. Str. 39, 3893–3906, 2002.<br />

240. M. M. J. Treacy, T. W. Ebbesen, J. M. Gibson. Nature<br />

381, 678–680, 1996.<br />

241. O. Lourie and H. D. Wagner, J. Mater. Res. 13, 2418,<br />

1998.<br />

242. E. W. Wong, P. E. Sheehan, and C. M. Lieber. Science<br />

277, 1971–1975, 1997.<br />

243. Z. W. Pan, S. S. Xie, L. Lu, B. H. Chang, L. F. Sun,<br />

W. Y. Zhou, G. Wang, and D. L. Zhang, Appl. Phys.<br />

Lett. 74, 3152, 1999.<br />

244. Xie SS, Li WZ, Pan ZW, Chang BH, Sun LF, J. Phys.<br />

Chem. Solids 61, 1153–1158 2000.<br />

245. Demczyk et al., Mater. Sci. Eng. A 334, 173–178,<br />

2002.<br />

246. C. Durkan, A. Ilie, M. S. M. Saifullah, and M. E.<br />

Welland, Appl. Phys. Lett. 80, 4244–4246, 2002.<br />

247. J. -P. Salvetat,, Briggs, G. A. D., Bonard, J. -M.,<br />

Bacsa, R. R., Kulik, A. J. Phys. Rev. Lett. 82, 944–<br />

947, 1999.<br />

248. Muster J, Burghard M, Roth S, Duesberg GS,<br />

Hernandez E, Rubio A, J. Vac. Sci. Tech. B, 16, 2796–<br />

2801, 1998.<br />

249. M. F. Yu, O. Lourie, M. J. Dyer, K. Moloni, T. F.<br />

Kelly, and R. S. Ruoff, Science 287, 637, 2000.<br />

250. Poncharal P., Wang L. Z., Ugarte D., de Heer W. A.,<br />

Science 283, 1513, 1999.<br />

251. H. O. Pierson, Handbook of <strong>Carbon</strong>, Graphite, Diamond,<br />

and Fullerenes: Properties, Processing and<br />

Applications, Noyes Publications, Park Ridge, NJ,<br />

1993.<br />

252. G. G. Tibbetts and C. P. Beetz. J. Phys. D 20, 292,<br />

1987.<br />

253. Wang ZL, Gao RP, Poncharal P, de Heer WA, Dai<br />

ZR, Pan ZW, Mater. Sci. Eng. B 16, 3–10, 2001.<br />

254. Yu MF, Files BS, Arepalli S, Ruoff RS, Phys. Rev.<br />

Lett., 84, 5552–5555, 2000.<br />

353


255. G. Van Lier, C. Van Alsenoy, V. Van Doren and P.<br />

Geerlings Chem. Phys. Lett. 326, pp. 181–185, 2000.<br />

256. Zhao QZ, Nardelli MB, Bernholc J, Phys. Rev. B, 65,<br />

144105, 2002.<br />

257. Li F, Cheng HM, Bai S, Su G, Dresselhaus MS, Appl.<br />

Phys. Lett., 77,3161–3163, 2000.<br />

258. D. A. Walters, L. M. Ericson, M. J. Casavant, J. Liu,<br />

D. T. Colbert, K. A. Smith, and R. E. Smalley, Appl.<br />

Phys. Lett. 74, 3803, 1999.<br />

259. M. Endo. J. Mater. Sci. 23, 598, 1988.<br />

260. H. D. Wagner, O. Lourie, Y. Feldman, R. Tenne,<br />

Appl. Phys. Lett. 72, 188, 1998.<br />

261. L. G. Zhou and S. Q. Shi, Comp. Mater. Sci. 23, 166–<br />

174, 2002.<br />

262. O. A. Shenderova, D. W. Brenner, A. Omeltchenko,<br />

X. Su, L. H. Yang, Phys. Rev. B 61, 3877 2000).<br />

263. Belyashko et al., in press, 2002.<br />

264. Xia YY, Zhao MW, Ma YC, et al. Phys. Rev. B 65,<br />

155415, 2002.<br />

265. G. G. Samsonidze, G. Samsonidze, B. Yakobson, Phys.<br />

Rev. Lett. 88, 065501–1, 2002.<br />

266. D. Roundy, private communication.<br />

267. Meng, G. W., Cui, Z., Zhang, L. D., Phillipp, F, J.<br />

Cryst. Growth, 209, pp. 801–806, 2000.<br />

268. Chen KH, Wu CT, Hwang JS, Wen CY, Chen LC,<br />

Wang CT, Ma KJ, J. PhysChem. Sol., 62, 1561–1565,<br />

2001.<br />

269. Zhu YF, Yi T, Zheng B, Cao LL, Appl. Sur. Sci., 137,<br />

83–90, 1999.<br />

270. Deryagin, B. V., Fedoseev, D. V., Luk’yanovich, V.<br />

M., Spitsin, B. V., Ryabov, V. A., Lavrent’ev, A. V.,<br />

J. Cryst. Growth 2(6), 380–384, 1968, Deryagin, B.<br />

V., Builov, L. L., Zubkov, V. M., Kochergina, A. A.,<br />

Fedoseev, D. V. Kristallografiya 14(3), 535–6, 1969,<br />

in Russian.<br />

271. Zamozhskii, V. D., Luzin, A. N, Dokl. Akad. Nauk<br />

SSSR 224(2), 369–72 [Phys. Chem.], 1975, in Russian.<br />

272. Butuzov, V. P., Laptev, V. A., Dunin, V. P.,<br />

Zadneprovskii, B. I., Sanzharlinskii, N. G., Dokl. Akad.<br />

Nauk SSSR 225(1), 88–90, 1975.<br />

273. Field, J. E., The Properties of Natural and Synthetic<br />

Diamond. Academic Press, London, 1992.<br />

274. H. O. Pierson, Handbook of <strong>Carbon</strong>, Graphite, Diamond,<br />

and Fullerenes: Properties, Processing and<br />

Applications, Noyes Publications, Park Ridge, NJ,<br />

1993.<br />

275. W. Lambrecht, C. H. Lee, B. Segall, J. C. Angus, Z.<br />

Li and M. Sunkara, Nature 364, 607, 1993.<br />

276. B. N. Davidson and W. E. Pickett, Phys. Rev. B 49,<br />

14770, 1994.<br />

277. V. L. Kuznetsov, A. L. Chuvilina, Y. V. Butenkoa, S.<br />

V. Stankusb, R. A. Khairulinb and A. K. Gutakovskiic,<br />

Chem. Phys. Lett. 289, 353, 1998.<br />

278. O. Shenderova, D. Areshkin, D. Brenner, Molec.<br />

Simul., in press.<br />

279. Nakaoka N, Watanabe, Phys. Rev. B 65, 155424–<br />

155429, 2002.<br />

280. J. W. Mintmire, B. I. Dunlap, and C. T. White, Phys.<br />

Rev. Lett. 68, 631, 1992.<br />

281. http://itri. loyola. edu/nano.<br />

282. Amato, Fortune Magazine, June 25, 2001.<br />

283. Dolmatov V. Yu., Burkat G. K., J. Superhard Materials<br />

1, 78–86, 2000.<br />

284. Dolmatov V. Y., Diamond Centre, S. Petersburg, Russia,<br />

private communication.<br />

285. A. Kalachev, PlasmaChem GmbH, Mainz/Germany,<br />

private communication.<br />

286. Voznyakovskij A. P., Dolmatov V. Yu. et al. Kauchuk<br />

u rezina, N 6, p. 27–30, 1996 (in Russian).<br />

287. I. V. Timoshkov et al. Galvanotekhnika and obrabotka<br />

poverkhnosti, 7, p. 20–26, 1999 (in Russian).<br />

288. V. M. Volkogon, J. Superhard Materials 4, 57–62,<br />

1998.<br />

289. V. Yu. Dolmatov, L. N. Kostrova, J. Superhard<br />

Materials 3, A. 79–82, 2000.<br />

290. G. P. Bogatyreva and M. A. Marinich, NATO Advanced<br />

Research Workshop on Nanostructured Materials<br />

and Coatings for Biomedical and Sensor Applications,<br />

August 4–8, 2002, Kiev, Ukraine.<br />

291. Bogatyreva GP, Marinich MA, Gvyazdovskaya VL,<br />

Dia. Rel. Mat., 9, 2002–2005, 2000.<br />

292. Corrigan TD, Gruen DM, Krauss AR, Zapol P, Chang<br />

RPH, Diam. Rel. Mater. 11, 43–48, 2002.<br />

293. Krauss AR, Auciello O, Gruen DM, et al., Diam. Rel.<br />

Mater. 10, 1952–1961, 2001.<br />

294. Auciello O, from presentation at NCSU.<br />

295. S. Mitura, A. Mitura, P. Niedzielski, and P. Couvrat,<br />

Chaos, Solitons, Fractals 10, 2165–2176, 1999.<br />

296. Y. Gogotsi, private communication.<br />

297. United States Patent Application 20010047980.<br />

298. Wilson, S. R. Biological aspects of fullerenes, in<br />

Fullerenes: Chemistry, Physics, and Technology,<br />

Kadish, K. M., Ruoff, R. S. (Eds), John Wiley &<br />

Sons, New York, 2000.<br />

299. Da Ros, T. et al., Chem. Commun. 663, 1999.<br />

300. Nanotechnology in Biology: The Good of Small<br />

Things, The Economist, Dec 22, 2001.<br />

301. http://www. csixty. com.<br />

302. M. F. Yu, M. J. Dyer, and R. S. Ruoff, J. Appl. Phys.<br />

89, 4554, 2001.<br />

303. D. Srivastava, D. W. Brenner, J. D. Schall, K. D.<br />

Ausman, M. F. Yu and R. S. Ruoff, J. Phys. Chem. B<br />

103, 4330, 1999.<br />

304. V. Lordi and N. Yao, J. Mater. Res. 15, 2770, 2000.<br />

305. S. J. V. Frankland, A. Caglar, D. W. Brenner, and M.<br />

Griebel, J. Phys. Chem. B 106, 3046, 2002.<br />

306. Lin, A. M. et al., J Neurochem. 72, 1634–1640, 1999.<br />

307. Da Ros, T. et al., J. Org. Chem. 61, 9070, 1996.<br />

308. K. E. Drexler, Nanosystems: Molecular Mashinery,<br />

Manufacturing, and Computation, John Wiley, New<br />

York, 1992.<br />

309. Robert A. Frietas, Jr., Nanomedicine, Volume 1: Basic<br />

Capabilities, Landes Bioscience, 1999.<br />

310. See, for example, R. Tuzun, D. Noid, B. Sumpter,<br />

Nanotechnology 6, 52–63, 1995.<br />

354


311. R. C. Merkle, Prospects in Nanotechnology, Ed. M.<br />

Krummenaeker and J. Lewis, p. 23–49, John Wiley,<br />

New York, 1995.<br />

312. Niemeyer CM, Angew Chem Int Edit 40 (22), 4128–<br />

4158, 2001.<br />

313. Website of <strong>Carbon</strong> Nanotechnologies, Inc. (CNI),<br />

Houston, Texas.<br />

314. I. Brodie and C. A. Spindt, Advances in Electron.<br />

Electron Phys. 83, 1, 1992.<br />

315. V. V. Zhirnov, C. Lizzul-Rinne, G. J. Wojak, R. C.<br />

Sanwald, J. J. Hren, J. Vac. Sci. Technol. B. 19, 87,<br />

2001.<br />

316. V. V. Zhirnov, E. I. Givargizov, J. Vac. Sci. Technol.<br />

12, 633, 1994.<br />

317. C. W. Oh, C. G. Lee, B. G. Park, J. D. Lee, J. H. Lee,<br />

J. Vac. Sci. Technol. B 16 807, 1998.<br />

318. D. S. Y. Hsu and J. Shaw, Appl. Phys. Lett. 80, 118,<br />

2002.<br />

319. K. Matsumoto, S. Kinosita, Y Gotoh, T. Uchiyama, S.<br />

Manalis, C. Quate, Appl. Phys. Lett. 78, 539, 2001.<br />

320. H. I. Lee, S. S. Park, D. I. Park, S. H. Hahm, J. H. Lee,<br />

J. H. Lee, J. Vac. Sci. Technol. B. 16, 762–764, 1998.<br />

321. W. P. Kang, A. Wisitsora-at, J. L. Davidson, IEEE<br />

Electron Dev. Lett. 379, 1998.<br />

322. W. P. Kang, A. Wisitsora-at, J. L. Davidson, M.<br />

Howell, D. V. Kerns, Q. Li, F. Xu, J. Vac. Sci. Technol.<br />

B 17, 740–743, 1999.<br />

323. A. Göhl, A. N. Alimova, T. Habermann, A. L.<br />

Mescheryakova, D. Nau, V. V. Zhirnov, G. Müller, J.<br />

Vac. Sci. and Technol. 17, 670, 1999.<br />

324. K. L. Jensen, R. H. Abrams, R. K. Parker, J. Vac. Sci.<br />

Technol. B 16, 749, 1998.<br />

325. W. Zhu, C. Bower, O. Zhou, G. Kochanski, S. Jin,<br />

Appl. Phys. Lett. 75, 873, 1999.<br />

326. J. T. L. Thong, C. H. Oon, and W. K. Eng, Appl. Phys.<br />

Lett. 79, 2811–2813, 2001.<br />

327. W. A. Mackie, T. Xie, M. R. Matthews, B. P. Routh,<br />

P. R. Davis, J. Vac. Sci. Technol. B 16, 2057–2062,<br />

1998.<br />

328. D. S. Y. Hsu and J. Shaw, Appl. Phys. Lett. 80, 118,<br />

2002.<br />

329. J. -M. Bonard, J. -P. Salvetat, T. Stöckli, L. Forró, A.<br />

Châtelain, Appl. Phys. A 69, 245, 1999.<br />

330. L. Nilsson, O. Groening, P. Groening, and L.<br />

Schlapbach, Appl. Phys. Lett. 79, 1036–1038, 2001.<br />

331. V. V. Zhirnov and J. J. Hren, MRS Bull. 23, 42, 1998.<br />

332. W. B. Choi, D. S. Chung, J. N. Kang, H. Y. Kim, Y.<br />

W. Jin, I. T. Han, Y. H. Lee, J. E. Jung, N. S. Lee, G.<br />

S. Park, J. M. Kim, Appl. Phys. Lett. 75, 3129, 1999.<br />

333. N. S. Lee, D. S. Chung, I. T. Han, J. H. Kang, Y. S.<br />

Choi, H. Y. Kim, S. H. Park, Y. W. Jin, W. K. Yi, M.<br />

J. Yun, J. E. Jung, C. J. Lee, J. H. You, S. H. Jo, C.<br />

G. Lee, J. M. Kim, Diamond Related Mater. 10, 265,<br />

2001.<br />

334. J-M. Bonard, T. Stockli, O. Noury, and A. Chatelain,<br />

Appl. Phys. Lett. 78, 2775–2777, 2001.<br />

335. H. Sugie, M. Tanemura, V. Filip, K. Iwata, and F.<br />

Okuyama, Appl. Phys. Lett. 78, 2578–2580, 2001.<br />

336. J. G. Leopold, O. Zik, E. Cheifets, D. Rosenblatt, J.<br />

Vac. Sci. Technol. A 19, 1790, 2001.<br />

337. C. Bower, D. Shalom, W. Zhu, D. Lopez, G. P.<br />

Kochanski, P. L. Gammel, S. Jin, IEEE Trans. Electron.<br />

Dev. 49, 1478, 2002.<br />

338. Y. Wei, C. Xie, K. A. Dean, B. F. Coll, Appl. Phys.<br />

Lett. 79, 4527, 2001.<br />

339. J. A. Hutchby, G. I. Bourianoff, V. V. Zhirnov, J. E.<br />

Brewer, IEEE Circuits Devices Mag. 18, 28, 2002.<br />

340. J. Kong, J. Cao, H. Dai, Appl. Phys. Lett. 80, 73, 2002.<br />

341. C. Zhou, J. Kong, E. Yenilmez, and H. Dai, Science<br />

290, 1552, 2000.<br />

342. J. Kong and H. Dai, J. Phys. Chem. 105, 2890, 2001.<br />

343. V. Derycke, R. Martel, J. Appenzeller, and Ph. Avouris,<br />

Appl. Phys. Lett. 80, 2773, 2002.<br />

344. S. J. Tans, A. R. M. Verschueren, C. Dekker C, Nature<br />

392, 49, 1998.<br />

345. R. Martel, T. Schmidt, H. R. shea, T. Hertel, Ph.<br />

Avouris, Appl. Phys. Lett. 73, 2447, 1998.<br />

346. P. G. Collins, M. Arnold, Ph. Avouris, Science 292,<br />

706, 2001.<br />

347. S. J. Wind, J. Appenzeller, R. Martel, V. Derycke, Ph.<br />

Avouris, Appl. Phys. Lett. 80, 3817, 2002.<br />

348. W. B. Choi, J. U. Chu, K. S. Jeong, E. J. Bae, J. W.<br />

Lee, J. J. Kim, J. O. Lee, Appl. Phys. Lett. 79, 3696,<br />

2001.<br />

349. A. Javey, Q. Wang, A. Ural, Y. Li, H. Dai., Nano<br />

Letters 2, 929, 2002.<br />

350. A. Bachtold, P. Hadley, T. Nakanishi, C. Dekker,<br />

Science 294, 1317, 2001.<br />

351. S. Heinze, J. Tersoff, R. Martel, V. Derycke, J.<br />

Appenzeller, Ph. Avouris, Phys. Rev. Lett 89, 6801,<br />

2002.<br />

352. O. A. Shenderova, B. L. Lawson, D. Areshkin and D.<br />

W. Brenner, Nanotechnology 12, 191, 2001.<br />

353. N. Hamada, S. Sawada, and A. Oshiyama, Phys. Rev.<br />

Lett. 68, 1579, 1992.<br />

354. C. T. White, D. H. Robertson and J. W. Mintmire,<br />

Phys. Rev. B 47, 5485, 1993.<br />

355. L. Chico, V. H. Crespi, L. X. Benedict, S. G. Louie,<br />

and M. L. Cohen, Phys. Rev. Lett. 76, 97,1996.<br />

356. V. Ivanov, J. B. Nagy, and Ph. Lambin, Chem. Phys.<br />

Lett. 223, 329, 1994.<br />

357. D. W. Brenner, J. D. Schall, J. P. Mewkill, O. A.<br />

Shenderova, S. B. Sinnott, J. Br. Interplan. Soc. 51,<br />

137, 1998.<br />

358. C. T. White and T. N. Todorov, Nature 393, 240,<br />

1998.<br />

359. J. Bernholc, D. Brenner, M. Buongiorno Nardelli, V.<br />

Meunier and C. Roland, Ann. Rev. Mater. Res. 32,<br />

347, 2002.<br />

360. R. Heyd, A. Charlier, and E. McRae, Phys. Rev. B 55<br />

6820, 1997.<br />

361. L. Yang, and J. Han, Phys. Rev. Lett. 85, 154, 2000.<br />

362. A. N. Andriotis, M. Menon, D. Srivastava, L.<br />

Chernozatonskii, Phys. Rev. B 65, 165416, 2002.<br />

363. A. N. Andriotis, M. Menon, D. Srivastava, L.<br />

Chernozatonskii, Phys, Rev. Lett. 87, 66802, 2001.<br />

355


364. M. Terrones, F. Banhart, N. Grobert, J. C. Charlier, H.<br />

Terrones H, and P. M. Ajayan, Phys. Rev. Lett. 89,<br />

75505, 2002.<br />

365. D. Goldhaber-Gordon and I. Goldhaber-Gordon,<br />

Molecular electronics — Momentous period for<br />

nanotubes, Nature 412, 6847, 2001.<br />

366. S. Iijima, C. Brabec, A. Maiti, J. Bernholc, Structural<br />

flexibility of carbon nanotubes, J. Chem. Phys. 104,<br />

2089, 1996.<br />

367. M. R. Falvo, G. J. Clary, R. M. Taylor, V. Chi, F. P.<br />

Brooks, S. Washburn, and R. Superfine, Nature 389<br />

582, 1997.<br />

368. S. B. Sinnott and A. Garg, Phys. Rev. B 60, 13786,<br />

1999.<br />

369. J. A. Harrison, S. J. Stuart, D. H. Robertson, C. T.<br />

White, J. Phys. Chem. 101, 9682, 1997.<br />

370. H. J. Dai, J. H. Hafner, A. G. Rinzler, D. T. Colbert,<br />

and R. E. Smalley, Nature 384, 147, 1996.<br />

371. G. Gao, T. Cagin, and W. A. Goddard, Nanotechnology<br />

9, 184, 1998.<br />

372. L. X. Benedict, N. G. Chopra, M. L. Cohen, A. Zettl, S.<br />

G. Louie, V. H. Crespi, Chem. Phys. Lett. 286, 490, 1998.<br />

373. Maiti, C. J. Brabec, and J. Bernholc, Phys. Rev. Lett.<br />

70, 3023–3026, 1993.<br />

374. M. I. Heggie, M. Terrones, B. R. Eggen et al., Phys.<br />

Rev. B 57, 13339, 1998.<br />

375. J. P. Lu and W. Yang, Phys. Rev. B 49, 11421, 1994.<br />

376. D. Tománek, W. Zhong, and E. Krastev, Phys. Rev. B<br />

48, 15461, 1993.<br />

377. Y. K. Kwon and D. Tomanek, Phys. Rev. B 58, 16<br />

001, 1998.<br />

378. Y. -K. Kwon, Y. H. Lee, S. -G. Kim, P. Jund, D. Tománek,<br />

and R. E. Smalley, Phys. Rev. Lett. 79, 2065, 1997.<br />

379. J. –C. Charlier and J. –P. Michenaud, Phys. Rev. Lett.<br />

70, 1858–1861, 1993.<br />

380. G. Gao, Cagin,T., Goddard,W. A., Nanotechnology<br />

9, 184–91, 1998.<br />

381. J. Tersoff R. S. Ruoff, Phys. Rev. Lett. 73, 676–679, 1994.<br />

382. M. J. López, A. Rubio, J. A. Alonso, L. -C. Qin, and<br />

S. Iijima, Phys. Rev. Lett. 86, 3056, 2001.<br />

383. Yoshida M., Osawa E., Fullerene Sci. Tech. 1, 55,<br />

1993.<br />

383. International Symposium, Detonation Nanodiamonds–<br />

2003. http://www.ioffe.rulnanodiamond.<br />

356

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!