16
REVIEW ARTICLE Low-temperature plasmas in carbon nanostructure synthesis Igor Levchenko a) Plasma Nanoscience Laboratories, CSIRO Materials Science and Engineering, P.O. Box 218, Lindfield, New South Wales 2070, Australia and Plasma Nanoscience@Complex Systems, School of Physics, The University of Sydney, Sydney, New South Wales 2006, Australia Michael Keidar Department of Mechanical and Aerospace Engineering, The George Washington University, Washington, DC 20052 Shuyan Xu Plasma Sources and Applications Centre, NIE, and Institute of Advanced Studies, Nanyang Technological University, 637616 Singapore Holger Kersten Institute of Experimental and Applied Physics, University Kiel, D-24098 Kiel, Germany Kostya (Ken) Ostrikov Plasma Nanoscience Laboratories, CSIRO Materials Science and Engineering, P.O. Box 218, Lindfield, New South Wales 2070, Australia and Plasma Nanoscience@Complex Systems, School of Physics, The University of Sydney, Sydney, New South Wales 2006, Australia (Received 4 July 2013; accepted 3 September 2013; published 20 September 2013) Plasma-based techniques offer many unique possibilities for the synthesis of various nanostructures both on the surface and in the plasma bulk. In contrast to the conventional chemical vapor deposition and some other techniques, plasma-based processes ensure high level of controllability, good quality of the produced nanomaterials, and reduced environmental risk. In this work, the authors briefly review the unique features of the plasma-enhanced chemical vapor deposition approaches, namely, the techniques based on inductively coupled, microwave, and arc discharges. Specifically, the authors consider the plasmas with the ion/electron density ranging from 10 10 to 10 14 cm 3 , electron energy in the discharge up to 10 eV, and the operating pressure ranging from 1 to 10 4 Pa (up to 10 5 Pa for the atmospheric-pressure arc discharges). The operating frequencies of the discharges considered range from 460 kHz for the inductively coupled plasmas, and up to 2.45 GHz for the microwave plasmas. The features of the direct-current arc discharges are also examined. The authors also discuss the principles of operation of these systems, as well as the effects of the key plasma parameters on the conditions of nucleation and growth of the carbon nanostructures, mainly carbon nanotubes and graphene. Advantages and disadvantages of these plasma systems are considered. Future trends in the development of these plasma-based systems are also discussed. [http://dx.doi.org/10.1116/1.4821635] I. INTRODUCTION Carbon nanostructures and nanomaterials such as gra- phene 1,2 and especially one-dimensional nanotubes, 3,4 nano- cones, 5,6 nanorods, 7,8 nanotips, 9,10 nanowires, 11,12 etc., feature many unique mechanical, optical, magnetic, and electrical properties, which offer superior opportunities for the modern applications ranging from photovoltaic cells, 13,14 energy stor- age and conversation devices, 15,16 memory devices, 17,18 nano- electronics 19,20 and supercapacitors 21,22 to nanomechanical and biomedical devices, 2325 gas and biosensors, 26,27 electron- emitting panels, 28,29 polymer–carbon reinforced composites, 30 etc. Presently, carbon nanostructures and nanomaterials are commonly fabricated using catalyzed chemical vapor deposition (CVD) technique. 3133 This method is simple, but a significant lack of controllability in the shape, size, density, and other characteristics of the nanostructures produced by the CVD motivates researchers to look for the alternative techniques. Besides, the structure of the carbon nanomaterials, such as the level of structural defects, vacancies, and impur- ities, is often beyond the limits required for the effective use of the nanostructures in the functional devices. Presently, plasma-enhanced CVD (PECVD) is a subject of a strong interest from academic and industry sectors because of its capacity to produce the carbon nanostructures of high quality in a controllable environment. 3436 Owing to a high level of controllability intrinsic to the plasma-based nanofabri- cation, 37,38 the catalytic and catalyst-free PECVD is a tech- nique particularly promising for the fabrication of various carbon nanostructures including nanotubes and graphene. 39,40 a) Author to whom correspondence should be addressed; electronic mail: [email protected] 050801-1 J. Vac. Sci. Technol. B 31(5), Sep/Oct 2013 2166-2746/2013/31(5)/050801/16/$30.00 050801-1 Author complimentary copy. Redistribution subject to AIP license or copyright, see http://jvb.aip.org/jvb/copyright.jsp

REVIEWARTICLE Low-temperature plasmas in carbon … · 2016. 5. 9. · high energy of the particles supplied to the growth zone, bal-anced influx of carbon material to the growing

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

  • REVIEW ARTICLE

    Low-temperature plasmas in carbon nanostructure synthesis

    Igor Levchenkoa)

    Plasma Nanoscience Laboratories, CSIRO Materials Science and Engineering, P.O. Box 218, Lindfield, NewSouth Wales 2070, Australia and Plasma Nanoscience@Complex Systems, School of Physics, The Universityof Sydney, Sydney, New South Wales 2006, Australia

    Michael KeidarDepartment of Mechanical and Aerospace Engineering, The George Washington University, Washington, DC20052

    Shuyan XuPlasma Sources and Applications Centre, NIE, and Institute of Advanced Studies, Nanyang TechnologicalUniversity, 637616 Singapore

    Holger KerstenInstitute of Experimental and Applied Physics, University Kiel, D-24098 Kiel, Germany

    Kostya (Ken) OstrikovPlasma Nanoscience Laboratories, CSIRO Materials Science and Engineering, P.O. Box 218, Lindfield, NewSouth Wales 2070, Australia and Plasma Nanoscience@Complex Systems, School of Physics, The Universityof Sydney, Sydney, New South Wales 2006, Australia

    (Received 4 July 2013; accepted 3 September 2013; published 20 September 2013)

    Plasma-based techniques offer many unique possibilities for the synthesis of various nanostructures

    both on the surface and in the plasma bulk. In contrast to the conventional chemical vapor

    deposition and some other techniques, plasma-based processes ensure high level of controllability,

    good quality of the produced nanomaterials, and reduced environmental risk. In this work, the

    authors briefly review the unique features of the plasma-enhanced chemical vapor deposition

    approaches, namely, the techniques based on inductively coupled, microwave, and arc discharges.

    Specifically, the authors consider the plasmas with the ion/electron density ranging from 1010 to

    1014 cm�3, electron energy in the discharge up to �10 eV, and the operating pressure ranging from1 to 104 Pa (up to 105 Pa for the atmospheric-pressure arc discharges). The operating frequencies of

    the discharges considered range from 460 kHz for the inductively coupled plasmas, and up to

    2.45 GHz for the microwave plasmas. The features of the direct-current arc discharges are also

    examined. The authors also discuss the principles of operation of these systems, as well as the

    effects of the key plasma parameters on the conditions of nucleation and growth of the carbon

    nanostructures, mainly carbon nanotubes and graphene. Advantages and disadvantages of these

    plasma systems are considered. Future trends in the development of these plasma-based systems

    are also discussed. [http://dx.doi.org/10.1116/1.4821635]

    I. INTRODUCTION

    Carbon nanostructures and nanomaterials such as gra-

    phene1,2 and especially one-dimensional nanotubes,3,4 nano-

    cones,5,6 nanorods,7,8 nanotips,9,10 nanowires,11,12 etc., feature

    many unique mechanical, optical, magnetic, and electrical

    properties, which offer superior opportunities for the modern

    applications ranging from photovoltaic cells,13,14 energy stor-

    age and conversation devices,15,16 memory devices,17,18 nano-

    electronics19,20 and supercapacitors21,22 to nanomechanical

    and biomedical devices,23–25 gas and biosensors,26,27 electron-

    emitting panels,28,29 polymer–carbon reinforced composites,30

    etc. Presently, carbon nanostructures and nanomaterials are

    commonly fabricated using catalyzed chemical vapor

    deposition (CVD) technique.31–33 This method is simple, but

    a significant lack of controllability in the shape, size, density,

    and other characteristics of the nanostructures produced by

    the CVD motivates researchers to look for the alternative

    techniques. Besides, the structure of the carbon nanomaterials,

    such as the level of structural defects, vacancies, and impur-

    ities, is often beyond the limits required for the effective use

    of the nanostructures in the functional devices.

    Presently, plasma-enhanced CVD (PECVD) is a subject of

    a strong interest from academic and industry sectors because

    of its capacity to produce the carbon nanostructures of high

    quality in a controllable environment.34–36 Owing to a high

    level of controllability intrinsic to the plasma-based nanofabri-

    cation,37,38 the catalytic and catalyst-free PECVD is a tech-

    nique particularly promising for the fabrication of various

    carbon nanostructures including nanotubes and graphene.39,40a)Author to whom correspondence should be addressed; electronic mail:

    [email protected]

    050801-1 J. Vac. Sci. Technol. B 31(5), Sep/Oct 2013 2166-2746/2013/31(5)/050801/16/$30.00 050801-1

    Author complimentary copy. Redistribution subject to AIP license or copyright, see http://jvb.aip.org/jvb/copyright.jsp

    http://dx.doi.org/10.1116/1.4821635http://dx.doi.org/10.1116/1.4821635mailto:[email protected]://crossmark.crossref.org/dialog/?doi=10.1116/1.4821635&domain=pdf&date_stamp=2013-09-20

  • Plasmas are commonly perceived as the “fourth state of

    the matter.” The plasmas represent a fully or partially ion-

    ized gas, i.e., ions, electrons, and neutral species.

    Importantly, the densities of the positive and negative par-

    ticles in the plasma are equal, i.e., the plasma is a quasineu-

    tral medium. One more important feature of the plasma is

    the interaction among the charged particles leading to the

    collective behavior of the whole ensemble. In technological

    applications, mainly a low-temperature plasma is used, with

    the ion energy not exceeding one hundred electron-volts (of

    the order of one million Kelvin). A high-temperature plasma

    with the ion energy exceeding 100 eV is mainly a subject of

    the controlled fusion research and is thus outside the scope

    of this review. Another way to classify the plasma is to com-

    pare the ion thermal energy ei with the energy of electronsee. The plasma is termed “equilibrium” if ei � ee, and“nonequilibrium,” if ei� ee. Both equilibrium and nonequi-librium low-temperature plasmas are used in industry and

    will be discussed here.

    An industry-scale production of the carbon nanostructures

    such as nanotubes and graphenes is another problem where

    the use of the low-temperature plasmas may be a possible so-

    lution.41,42 Indeed, the production of carbon nanotubes and

    graphene requires specific conditions such as sufficiently

    high energy of the particles supplied to the growth zone, bal-

    anced influx of carbon material to the growing nanostructure,

    and a specially prepared (fragmented and activated) cata-

    lyst.43,44 In the conventional CVD process, these conditions

    are ensured by the use of very high (typically about

    800–1200 �C) process temperature. In the plasma-basedprocesses, these conditions can be easily met by the proper

    selection of the process parameters such as density, electron

    temperature, and surface potential. The requirement of the

    high plasma density makes the arc plasma especially promis-

    ing for the large-scale production of the carbon

    nanomaterials.45

    One more important point is the environment-benign na-

    ture of the plasma-based techniques. In fact, many plasma-

    based technological processes are conducted in a confined

    reactor chamber where a controlled atmosphere is used (Fig.

    1). In vacuum-based methods, air cannot be used as a pro-

    cess environment due to too high oxygen content, and thus,

    the controlled gas mixture is created and maintained in the

    chamber during treatment, and then exhausted through the

    filters, in a controlled manner. Thus, the human respiratory

    tract remains isolated from the harmful volatile compounds,

    and exposure to toxic gas or liquid chemicals is reduced.

    Also, humans remain isolated from any expose to the nano-

    materials which are kept in the reactor chamber during the

    growth process.46

    In our paper, we review the capability of the plasmas pro-

    duced by the inductively coupled (ICP), microwave (MWP),

    and arc discharge systems for the fabrication of various car-

    bon nanostructures, mainly carbon nanotubes and graphenes.

    We will discuss some of the principles of operation of these

    systems, as well as the key plasma parameters and their

    effects on the conditions of nucleation and growth of carbon

    nanostructures. Advantages and disadvantages of these

    plasma systems will be considered. Future trends in the de-

    velopment of these plasma-based systems will also be

    addressed.47,48

    This review is mainly based on the results obtained in col-

    laborative efforts of researchers from CSIRO (Australia),

    George Washington University (USA), the University of

    Kiel (Germany), and Nanyang Technological University,

    Singapore. Specializing in different plasma systems and

    processes, these research groups developed several practical

    techniques through collaborative efforts. The demonstration

    of the usefulness of international collaborations is one of the

    aims of this work. Along with this, a large body of the results

    obtained by other researchers is discussed.

    II. INDUCTIVELY COUPLED AND MICROWAVEPLASMA SYSTEMS

    A. Principle of operation and plasma parameters

    The chemical vapor deposition is a process widely used

    for the fabrication of various surface-bound nanomaterials

    such as nanotubes, nanorods, nanocones, nanotips, and nano-

    wires. In this process, a solid material is deposited onto a

    heated substrate from gaseous precursors. Chemical reac-

    tions in the gas environment are used to synthesize the target

    nanomaterial on the surface. In many cases, CVD technique

    involves a complex chain of the processes including dissoci-

    ation of the carbon-containing precursors (such as methane,

    acetylene, ethylene, ethanol, carbon dioxide, etc.), formation

    of larger molecules in the gas environment and on the surfa-

    ces of the substrate and catalyst, diffusion of various species

    over the substrate and catalyst surfaces, saturation of the cat-

    alyst,49,50 nucleation of the nanostructures on the cata-

    lyst51,52 or directly on the substrate surface.53 As a result,

    growth of the pattern of nanostructures occurs due the mate-

    rial flux directly from the gas phase or by the surface diffu-

    sion,54 and further incorporation of atoms into the growing

    structure.55

    FIG. 1. (Color online) Plasma-based technological nanofabrication is con-

    ducted in a confined reactor chamber. The worker breathing zone remains

    isolated from the harmful volatile compounds, and no toxic gas or liquid

    chemicals can be exposed to humans. No radiation is present outside the

    chamber, no harmful gas leakage and harmful by-products. Reprinted with

    permission from Han et al., J. Phys. D: Appl. Phys. 44, 174019 (2011).Copyright 2011, IOP Publishing Ltd.

    050801-2 Levchenko et al.: Low-temperature plasmas in carbon nanostructure synthesis 050801-2

    J. Vac. Sci. Technol. B, Vol. 31, No. 5, Sep/Oct 2013

    Author complimentary copy. Redistribution subject to AIP license or copyright, see http://jvb.aip.org/jvb/copyright.jsp

  • Due to many drawbacks of the CVD technique such as

    low controllability with respect to the nanostructure proper-

    ties, position, size, and shape, as well as low growth rate and

    need for a very high surface temperature, the plasma-based

    CVD process was proposed. This process relies on the

    plasma environment as an extra dimension to control

    the processes related to the nucleation and growth of the

    surface-bound nanostructures. The plasmas differ from neu-

    tral gases by the presence of free electrons and various ion-

    ized species which in turn could be in various excited

    states.56 As a result, the plasma may be affected by an exter-

    nal electric or magnetic field, and hence, the ion and electron

    fluxes may be effectively controlled. Indeed, numerous stud-

    ies on the plasma-based fabrication of various nanostructures

    have demonstrated unprecedented level of control over the

    material fluxes, heating, and eventually structure and proper-

    ties of the nanostructures grown on the plasma-immersed

    surfaces.57 Among others, the possibility to grow large pat-

    terns of the vertically aligned nanostructures is one of the

    major advantages of PECVD, as compared to the conven-

    tional CVD technique.58 Other benefits of the plasma envi-

    ronment include, but not limited to, the lower temperatures

    required for nucleation and growth of the nanostructures, as

    well as much higher growth rates.

    The low-frequency inductively coupled plasma reac-

    tors59,60 are successfully used for the fabrication of various

    surface-bound nanostructures such as nanorods,61 nano-

    cones,36 nanowires,62 graphene,63 and large patterns of nano-

    dots.64 Schematics of the two typical ICP reactors are shown

    in Fig. 2. The most frequently used reactors consist of a vac-

    uum chamber and a planar inductive coil installed on the

    upper lid of the chamber, as shown in Fig. 2(a), or a coil in-

    stalled over the confinement tube [Fig. 2(b)]. Also, various

    multispiral configurations may be used.59

    In contrast to the capacitively coupled plasmas, the induc-

    tively coupled discharge is sustained by the alternating radio

    frequency electromagnetic field generated by the induction

    coil installed on the dielectric lid or over the dielectric con-

    finement tube. An alternating electromagnetic field is used

    for the heating and electron-impact ionization of the gas,

    eventually sustaining the steady discharge. The ICP operated

    in the H-mode typically features a much higher plasma den-

    sity, as compared to capacitively coupled plasmas.59

    The microwave plasma reactors used in the nanofabrica-

    tion operate at typical microwave frequencies such as 2.45

    GHz. The schematic of the MWP setup is similar to that of

    ICP.53 The reactor chamber is connected to the microwave

    generator (magnetron) with the help of a waveguide, and

    thus, the microwave energy is supplied through the quartz

    window to the chamber where the discharge is sustained.

    Both ICP and MWP reactors are usually equipped with

    the vacuum and gas supply systems capable of maintaining

    the required pressure and gas composition in the reaction

    chamber. To increase the energy of ions supplied to the

    growing nanostructures, a negative biasing potential can be

    applied to the substrate holder. Also, the reactors are usually

    equipped with a diagnostic system consisting of various

    types of probes, to enable measurements of the most impor-

    tant parameters such as the electron energy and the plasma

    density. The gas pressure in the chamber is usually moni-

    tored with the help of Pirani and Penning gauges, as well as

    capacitance manometers.

    It should be also stressed that the surface temperature is

    one of the key parameters in nanofabrication.65,66 In the

    plasma-based systems, the surface temperature can be con-

    trolled even without external heating, by regulating the

    energy flux from the plasma.67 Some typical characteristics

    of the ICP and MWP reactors are summarized in Table I.

    FIG. 2. (Color online) (a) Schematic of the ICP reactor with planar coil on top of the vacuum chamber. (b) Schematic of the ICP reactor with the coil installed

    over the confinement chamber. (c) Photo of the discharge in the confinement chamber of the setup shown in (b).

    050801-3 Levchenko et al.: Low-temperature plasmas in carbon nanostructure synthesis 050801-3

    JVST B - Microelectronics and Nanometer Structures

    Author complimentary copy. Redistribution subject to AIP license or copyright, see http://jvb.aip.org/jvb/copyright.jsp

  • B. Growth of vertical carbon nanostructures in ICP

    Owing to high plasma density and ion energy, the ICP

    reactors are effective for the growth of various vertically

    aligned carbon nanostructures, such as carbon nanocones.

    An example of such nanostructures successfully grown in

    the ICP reactor with the planar inductive coil is shown in

    Figs. 3 and 4.36 The nanocones were grown on a lightly

    doped n-Si(100) substrate, which was coated with a nickel

    catalyst (layer of 30 nm thick deposited by the magnetron

    sputtering) before treatment in the ICP reactor. The catalyst-

    coated specimens were installed in the reactor chamber on

    the sample holder, and first pretreated by the discharge

    ignited in argon for 30 min. Next, a discharge in a mixture of

    argon and hydrogen was used to prepare the catalyst layer

    for nanostructure nucleation. Specifically, during this stage,

    the metal catalyst layer was fragmented into separate islands,

    suitable for the nucleation of separate nanocones. After pre-

    paring catalytically active metal islands, methane was sup-

    plied to the reactor chamber, and the growth of nanocones

    started. During the growth, no external heating was used and

    the growing nanostructures were heated by the plasma flux,

    which was strong enough to maintain the surface tempera-

    ture of about 500 �C when the biasing potential of 300 V wasapplied.36 During the process, the plasma with the density of

    1012 cm�3 was sustained, when the radio-frequency power of

    0.1 W� cm�1 was applied to the upper coil.As it can be seen in Fig. 3, dense patterns of the carbon

    nanocones with the height of 300–800 nm were grown. For

    the short deposition time (5 min), very nonuniform pattern

    was formed, whereas after 10 min into the process, the uni-

    formity was significantly improved, and quite uniform

    patterns eventually formed after 20 min of deposition [Fig.

    3(c)]. Along with the uniformity, the shape of the nanocones

    also changed and the nanostructures became much thinner.

    The Raman characterization technique has shown that the ra-

    tio of G and D bands increased with the deposition time, i.e.,

    the crystalline structure was improved and the amount of

    amorphous carbon decreased with the process time.36 The

    XRD technique has also evidenced the best crystalline struc-

    ture of the samples grown for the longer time. Thus, the

    plasma-related processes can significantly enhance the qual-

    ity of the synthesized nanostructures.

    C. Catalyst-free growth of vertical carbon nanotubesin microwave plasma

    Microwave plasmas can be successfully used for the

    growth of large arrays of vertically aligned carbon nano-

    tubes.53 Normally, the plasma-based fabrication of carbon

    nanomaterials requires metal catalyst to initiate the nuclea-

    tion and sustain the nanostructure growth. Fabrication of

    metal-free nanotubes not containing any metal particles rep-

    resents a strong interest for the nanoelectronic applications.

    The nucleation and growth of carbon nanomaterials can be

    achieved without a metal catalyst, but this is done by a com-

    plex postprocessing,68 which leads to the degradation of the

    nanotube properties and damage to the substrate.69

    Therefore, the metal-free fabrication of the carbon nanotubes

    is not practical for the industrial-scale production.70 It should

    be noted that the carbon nanomaterials could be nucleated

    on nonmetallic catalysts (e.g., on semiconductor nanopar-

    ticles), but only low-quality structures could be produced in

    such process.71,72 The nanotubes catalyzed and grown with-

    out metal catalyst are usually not vertically aligned and have

    irregular pattern structure. Nevertheless, the microwave

    plasma technique enables an efficient catalyst-free nuclea-

    tion and growth of dense, vertically aligned arrays of long

    nanotubes on silicon wafers.53

    In the experiments on catalyst-free fabrication of carbon

    nanotubes,53 several different experimental variations were

    used with respect to the plasma/gas environments, plasma

    location relative to the substrate, and presence of the special

    notch and dot patterns on the surface. Specifically, the fol-

    lowing configurations were used: gas contacting the surface

    with/without any written features; remote plasma and

    TABLE I. Typical parameters of ICP and MWP reactors used for the

    nanofabrication.

    Parameter ICP MWP

    Discharge power range (kW) 0.01–10 0.1–10

    Plasma density (cm�3) 1010–1013 1010–1011

    Electron energy in discharge (eV) 1–5 5–10

    Operating pressure range (Pa) 1–1000 10–10 000

    Typical frequency 0.46–13.56 MHz 2.45 GHz

    Power density (W� cm�3) 0.01–0.5 1–2Substrate temperature range ( �C) 100–500 100–800

    FIG. 3. SEM images of carbon nanocones grown in ICP reactor for different growth times: (a) 5 min; (b) 10 min; and (c) 20 min. Reprinted with permission

    from Tsakadze et al., Carbon 45, 2022 (2007). Copyright 2007, Elsevier Ltd.

    050801-4 Levchenko et al.: Low-temperature plasmas in carbon nanostructure synthesis 050801-4

    J. Vac. Sci. Technol. B, Vol. 31, No. 5, Sep/Oct 2013

    Author complimentary copy. Redistribution subject to AIP license or copyright, see http://jvb.aip.org/jvb/copyright.jsp

  • surface with/without written features, and plasma contacting

    the surface with/without written features. The nanotube

    nucleation in the neutral gas environment and in the

    remotely located plasma was not observed for any surfaces,

    and only the plasma contacting the mechanically patterned

    surface (parallel linear notches and spots were mechanically

    written on silicon) has produced the nanotubes. The micro-

    wave discharge was sustained for 3–5 min in a mixture of

    methane and nitrogen at a pressure of typically 10 Torr, and

    the surface temperature was maintained in the range of

    700–800 �C by the plasma heating. These arrays wereformed in a very fast process with the growth rates up to

    �50 lm/min. These growth rates are much higher comparedto the processes conducted in the absence of metal catalyst.

    Figure 5 illustrates the fabricated nanotubes. Scanning

    electron microscopy (SEM) has shown that a high-density

    forest of highly aligned nanotubes precisely replicated the

    notch pattern, including complex patterns consisting of lin-

    ear notches and spots applied directly over the notch. The

    transmission electron microscopy (TEM) and Raman charac-

    terizations of the nanotube structure have clearly shown the

    absence of catalyst particle at the closed end tip of the nano-

    tubes, thus revealing the base-led growth mode.73,74 The

    nanotubes reached 80 nm in diameter.

    The mechanism of the nanostructure growth without

    metal catalyst particles was explained based on the key role

    of nanoscale features on the surface.53 The nanoscale fea-

    tures were created by mechanical writing of the pattern.

    Since the carbon solubility in silicon is low75 and very high

    growth rate was observed, one can assume that the bulk dif-

    fusion of carbon in silicon did not play a major role, and a

    vapor–liquid–solid mechanism was not involved. Initially,

    the tip of a Si nanoscale feature was heated up by the

    plasma. Then, the heated tip reshapes76 due to the rearrange-

    ment of material. Next, the carbon atoms form closed chains

    in the steps on the silicon surface. Finally, the nanotubes nu-

    cleate from these chains and start growing. When the nano-

    tube reaches the maximum length, the tip of nanotube

    closes. Hence, in this process, the nucleation of carbon nano-

    tubes occurs by the minimization of surface energy at the

    steps formed on silicon surface,77 where the adsorbed atoms

    can be considered as “partially dissolved.” Since the silicon

    surface is covered with a large number of nanoscale surface

    features, the resulting array of long multiwall nanotubes is

    very dense and formed only on the mechanically patterned

    areas.

    The plasma-related effects are essential for the above pro-

    cess since the plasma-related heating and high rate of mate-

    rial delivery are of the primary importance here. Therefore,

    the catalyst-free, dense arrays of long vertically aligned mul-

    tiwall nanotubes can be grown using a low-temperature

    microwave discharge on the silicon wafers where the arbi-

    trary pattern was mechanically applied. Schematic of the

    nanotube growth on the surface features is shown in Fig. 6.

    Note that the nanotube patterns fabricated using micro-

    wave plasmas feature a very high density. In contrast, the

    use of neutral gas-based CVD process results in the forma-

    tion of nanotube patterns of a significantly lower density, as

    evidenced by SEM images shown in Fig. 7(a). After post-

    treatment in acetone, such patterns collapse by the action of

    capillary forces [Figs. 6(b) and 6(c)].78

    FIG. 4. SEM (a), (b) and TEM (inset) images of single-crystalline carbon

    nanocones grown in ICP reactor. Reprinted with permission from

    Levchenko et al., Appl. Phys. Lett. 91, 113115 (2007). Copyright 2007,American Institute of Physics.

    FIG. 5. (Color online) Scheme of patterns writing and SEM micrographs of

    nanotubes grown on the features of different shapes and sizes. (a) Process of

    writing linear and dot patterns on a Si substrate. (b) Top-view of nanotubes

    on two closely positioned dot features. (c) Tilted view of a high-density

    nanotube array on a dot feature. In the central area, the nanotubes can be

    a few hundred microns long. (d), (e) Dense array of vertically aligned

    nanotubes on a linear feature, at different magnification. Reprinted with

    permission from Kumar et al., Carbon 50, 325 (2011). Copyright 2011,Elsevier Ltd.

    050801-5 Levchenko et al.: Low-temperature plasmas in carbon nanostructure synthesis 050801-5

    JVST B - Microelectronics and Nanometer Structures

    Author complimentary copy. Redistribution subject to AIP license or copyright, see http://jvb.aip.org/jvb/copyright.jsp

  • D. Controllable growth of graphene flakes in ICP

    Graphene was discovered recently and now it attracts

    great attention of the researchers due to exceptional mechan-

    ical, electrical, optical, and other properties. Graphene can

    be a very promising material for many applications in nanoe-

    lectronics, energy conversion, sensing, various nanodevices,

    supercapacitors, nanoelectromechanical systems, and many

    others.79,80

    Many applications require graphene nanosheets in the

    form of flakes consisting of a few graphene monolayers.

    These nanosheets should be separated and securely attached

    to the surface of the substrate suitable for the specific appli-

    cation (e.g., doped silicon). Besides, it is highly desirable to

    precisely control the position and orientation of the graphene

    nanosheets during the fabrication.81–85 The vertically aligned

    graphene nanosheets expose open edges which ensure many

    unique features, related to the high reactivity of the reactive

    edges of graphene monolayers, which can enhance electric

    field, attach atoms and biological species.86–88 The known

    fabrication techniques capable of producing high-quality

    graphenes (such as chemical solution methods,89 arc dis-

    charges,90 and CVD) could not ensure precise control of the

    graphene alignment and orientation. None of these techni-

    ques is able to produce high-quality vertically aligned gra-

    phenes on the wafer surface. Instead, only unbound (like in

    arc discharges) or flat-oriented (by chemical methods) gra-

    phenes can be produced.

    The low-temperature ICP plasmas can be useful for the

    fabrication of large arrays of well-separated graphene nano-

    sheets, vertically aligned and bound on the wafer surface.

    Moreover, the plasma-based process demonstrates an excel-

    lent controllability of the graphene orientation and place-

    ment on the surface. In the typical experiments, the plasma-

    produced nanosheets were perfectly oriented along the

    substrate center–substrate edge direction in the pretty wide

    (as compared with the nanosheet size) substrate edge zone.

    These nanosheets rested on the silicon nanograss (a dense

    array of nanoscaled silicon cones on the silicon surface91),

    fabricated from the silicon substrate in a separate plasma-

    based process.92 SEM, TEM, and Raman techniques were

    used to characterize the produced nanosheets. A perfect

    alignment, orientation, and the few-layer internal structure

    of the nanosheets were confirmed. These perfect nanostruc-

    tures can be used for the development of advanced nanodevi-

    ces that require graphene directly integrated into a Si-based

    nanodevice platform.

    The fabrication process involved two main stages. At the

    first stage, silicon nanograss was fabricated on the silicon

    substrate by processing the wafers for 120 min in a radio-

    frequency ICP discharge93 ignited in H2 gas at a pressure of

    5.0 Pa. The discharge was sustained by supplying

    500–600 W RF power to the chamber. The substrate temper-

    ature was maintained in the range of 350–400 �C, and thewafer was biased negatively by applying the �250–300 Vpotential relative to the grounded chamber walls. The nano-

    grass consisted of sharp self-organized Si nanoconical struc-

    tures, typically of about 1 lm in height, with the basediameter up to 100 nm. The produced pattern was fairly

    dense and uniform (see more details on the nanograss struc-

    ture and characterization elsewhere.92) Then, the nanograss

    was used as a catalyst-free platform for the fabrication of

    self-organized vertically aligned graphene nanosheets. The

    wafer surface was activated for 3 min using radio-frequency

    ICP plasmas ignited in nitrogen (gas pressure of 4 Pa and RF

    power of 400 W). The growth process was initiated by

    FIG. 6. (Color online) Plausible mechanism of nucleation and growth of mul-

    tiwalled carbon nanotubes on a silicon nanohillock in the plasma. (a) Si

    nanohillock is locally heated by the plasma. (b) Carbon atoms form chains

    on the heated Si nanohillock. (c) Carbon chains close, nanotube walls nucle-

    ate. (d) Nanotubewalls grow in the open-end mode and eventually close.

    Reprinted with permission from Kumar et al., Carbon 50, 325 (2011).Copyright 2011, Elsevier Ltd.

    FIG. 7. SEM images of the three-dimensional structures formed from CVD-grown carbon nanotube patterns and postprocessing with acetone, produced in

    Plasma Nanoscience Labs at CSIRO (Yick et al., unpublished). (a) Side view; (b), (c) Top view.

    050801-6 Levchenko et al.: Low-temperature plasmas in carbon nanostructure synthesis 050801-6

    J. Vac. Sci. Technol. B, Vol. 31, No. 5, Sep/Oct 2013

    Author complimentary copy. Redistribution subject to AIP license or copyright, see http://jvb.aip.org/jvb/copyright.jsp

  • adding methane to the discharge. During the growth, a dis-

    charge power increased to 800–1000 W was used, along with

    the higher (up to 5–10 Pa) gas pressure and surface bias of

    �50 V. The wafer was heated by the plasma irradiation only,and the substrate temperature did not exceed 500 �C. After10 min into the process, well-aligned vertical graphene nano-

    sheets (VGNS) were fabricated (Fig. 8).

    Thus, the low-temperature ICP and microwave plasmas

    are efficient tools for the fabrication of various surface-

    bound nanostructures such as carbon nanotubes, nanocones,

    and graphene flakes. Notably, some very important features

    not possible in the gas-based fabrication techniques were

    achieved in plasma-based processes, namely, an efficient

    catalyst-free growth of carbon nanotubes, fabrication of the

    complex structures of nanotubes on the mechanically written

    patterns, and highly controllable fabrication of the few-layer

    vertically aligned graphene nanosheets.

    E. Synthesis of carbon nanoparticlesin radio-frequency plasma

    Along with other plasma-grown carbon nanostructures, the

    nanoparticles grown directly in the volume of process cham-

    ber are of a particular interest due to possible precise deposi-

    tion onto specific locations using custom-shaped electric

    fields immediately after nucleation and growth in the gas

    phase.94 The plasma technique allows efficient nucleation and

    growth of the carbon nanoparticles from carbon-containing

    gases.95,96 In typical experiments,97 a specially designed setup

    was used in which an asymmetric radio-frequency discharge

    at 13.56 MHz is sustained. Argon and acetylene are typical

    gases that are used as a background gas and a reactive precur-

    sor, respectively. The power of the discharge may be varied in

    a broad range, reaching 50 W. Under these conditions, the

    plasma density reaches �109 cm�3.The cloud of nucleated carbon nanoparticles can be made

    visible by illuminating them with a laser beam, as shown in

    Fig. 9(a). Furthermore, a void (dust-free region) can be

    observed between the two clouds. These plasma-grown nano-

    particles precipitate to the lower electrode after acquiring an

    excess mass precluding them from levitation. Finally, they

    can be collected on the electrode surface. Figure 9(b) is an

    SEM image of nanoparticles found on the electrode surface.

    The nucleation and growth of nanoparticles in low-

    temperature plasma proceeds via the following mecha-

    nism.98 After the initial phase, the nuclei of a few nanometer

    size move in the plasma, collide, and coagulate.99 Finally,

    they form larger particles with the size of several tens of

    nanometers.100,101 The density of such particles typically

    reach (1–5)� 103 cm�3 (up to 104 cm�3 for higher dischargepowers). Since the larger negatively charged nanoparticles

    have larger collision cross-sections, they also accumulate

    neutral radicals and positive ions, and thus grow up to sev-

    eral hundred nanometers. The size distribution of the nano-

    particles tends to be monodisperse.100,102 It should be also

    mentioned that the density of nanoparticles can reach

    FIG. 8. (Color online) Different orientation of the VGNSs (SEM images)

    grown near the edges, corners, and in the central area of the silicon substrate

    covered with nanograss (optical image in the center). The VGNSs grown in

    the central area of the substrate (lower right panel) show a random orienta-

    tion. The graphenes grown near the substrate edge align in the edge-center

    direction. Reprinted with permission from Kumar et al., Appl. Phys. Lett.100, 053115 (2012). Copyright 2012, American Institute of Physics.

    FIG. 9. (Color online) (a) Photo of illuminated clouds of nanoparticles in the

    plasma. The region free of nanoparticles is clearly visible; (b) Typical car-

    bon nanoparticles collected on the electrode surface (SEM image). The scale

    bar in the inset is 200 nm. Reprinted with permission from Hundt et al., J.Appl. Phys. 109, 123305 (2011). Copyright 2011, American Institute ofPhysics.

    050801-7 Levchenko et al.: Low-temperature plasmas in carbon nanostructure synthesis 050801-7

    JVST B - Microelectronics and Nanometer Structures

    Author complimentary copy. Redistribution subject to AIP license or copyright, see http://jvb.aip.org/jvb/copyright.jsp

  • �107 cm�3 in the pulsed symmetric Ar-C2H2 capacitivelycoupled discharges; in this case, the charged nanoparticles

    can significantly affect the plasma properties.103

    F. Growth of nanostructures on plasma-exposedsurfaces—theoretical examination

    When examining the growth of nanostructures on the

    plasma-exposed surface, the two main groups of the physical

    processes should be distinguished, namely, surface processes

    and plasma processes. While the surface processes are mainly

    responsible for the formation of the internal structure of the

    nanostructures and the morphology on nanostructure patterns,

    the plasma processes are mainly responsible for the material

    delivery to the substrate surface and surfaces of the growing

    nanostructures. Nevertheless, these groups of the processes

    are tightly interrelated and should be examined in the frame-

    work of the integrated model. The most important processes

    occurring on the plasma-exposed surfaces are the surface dif-

    fusion between growing nanostructures, evaporation of the

    adsorbed particles from the wafer and nanostructure surfaces

    to the space and 2D vapor, and attachment of the particles to

    the growing nanostructures.104 The aim of the surface model

    is to calculate the balance of material fluxes on the surface of

    the nanostructures, rate of the nanostructure growth and, even-

    tually, an internal structure of the nanostructures.

    The material balance on the individual nanostructure is

    wi ¼ wþP þ wþS � w�E � w�N ; (1)

    where wi is the total flux of material to the individual nano-structure, wþP is the flux of material to the nanostructure fromthe plasma, wþS is the flux of material adsorbed on the surfaceto an individual nanostructure, w�E is the evaporation from thenanostructure, and w�N is the flux of material evaporated fromthe nanostructure to the two-dimensional vapor. A similar

    relation may be written for the energy balance, taking into

    account various energy contributions such as radiation, power

    transfer by electrons, ions and neutral species, and other

    energy sources including possible exothermic and endother-

    mic reactions on the surface.67,105 Then, the flux to the surface

    of the individual nanostructure from the plasma wþP is

    wþP ¼ðS

    jSNdsþ SNwan; (2)

    where jSN is the ion flux from plasma to the nanostructure, SNis the area of the nanostructure surface, and wan is the flux ofneutral particles to the nanostructure from the plasma. The

    flux wþS can be obtained by solving the diffusion equation forthe density of atoms g(x, y, t) adsorbed on the substrate50,104

    @g@t¼ D @

    2g@x2þ @

    2g@y2

    !þ win � wvpi; (3)

    where win is the total flux of atoms and ions to the substratefrom the plasma, wvpi is the flux of evaporation, and D is thesurface diffusion coefficient. The evaporation flux wvpi is

    50

    wvpi ¼ ð1=k2aÞ �0 expð�eSe=kTÞ; (4)

    where eSe is the energy of evaporation from the surface, T isthe surface temperature, ka is the lattice constant, �0 is thefrequency of lattice atom oscillations, and k is theBoltzmann’s constant.

    The total flux to the substrate from the plasma, win, is

    win ¼ ð1=k2aÞðS

    jSdsþ wa; (5)

    where jS is the density of the ion flux to the substrate, and wais the flux of neutral species to the substrate from the

    plasma.

    Hence, the total flux of adsorbed atoms to the individual

    nanostructure can be calculated as50

    wþS ¼ �Lm

    k3aqmD@g@x

    � �; (6)

    where L is the perimeter of nanostructure, m is the adatommass, and qm is the nanostructure material density. The sur-face diffusion coefficient can be calculated as

    D ¼ k2a�04

    expð�ed=kTÞ; (7)

    where ed is the surface diffusion activation energy.The flux of atoms evaporating from the surface of an indi-

    vidual nanostructure w�E into the 3D vapor is calculated as

    w�E ¼ ðSi=kaÞ �0 expð�ea=kTÞ; (8)

    where Si is the surface area of an individual nanostructure, andea is the evaporation energy. The two-dimensional flux w

    �N from

    the individual nanostructure is calculated similarly to Eq. (8).

    To describe the growth dynamics of various nanostruc-

    tures, different forms of the growth equations should be

    used. The nanocones may be described with the following

    set of equations:36

    ð@V=@rÞ dr ¼ k3aðwþS � w�N Þ dt (9)

    and

    ð@V=@hÞ dh ¼ k3aðwþP � w�E Þ dt; (10)

    where V is the volume, r is the base radius, and h is theheight of the nanocone.

    The most important feature of the plasma–surface inter-

    actions influencing the kinetics of material supply to the

    plasma-immersed body is the formation of a sheath above

    the surface.106 The ions enter the sheath with the Bohm ve-

    locity Vb¼ (kTe/mi)1/2 (assuming that the sheath is colli-sionless), where Te is the electron temperature and mi isthe ion mass. In the most common case of the thick sheath

    (for kTe � US, where US is the surface potential), thesheath width can be estimated as kS � kDð2US=kTeÞ, where

    050801-8 Levchenko et al.: Low-temperature plasmas in carbon nanostructure synthesis 050801-8

    J. Vac. Sci. Technol. B, Vol. 31, No. 5, Sep/Oct 2013

    Author complimentary copy. Redistribution subject to AIP license or copyright, see http://jvb.aip.org/jvb/copyright.jsp

  • kD ¼ ðe0kTe=neÞ1 2=

    is the Debye length, e is the electroncharge, e0 is the dielectric constant, and n is the plasmadensity.

    The electric field in the vicinity of the nanostructure is

    very strong due to very large surface curvature. Thus, the

    density of the electric charges in the plasma–surface sheath

    can be neglected, and the electric field can be found from the

    Laplace equation Du¼ 0, with the boundary condition ofequipotentiality over the entire conductive surface

    /ðx; y; zÞjSURF ¼ US: (11)

    Note that we are considering mainly carbon nanostruc-

    tures, so the surface can be assumed conductive in most

    cases, e.g., related to graphitic carbons.

    Thus, the model describes material supply through the

    plasma–surface sheath to the wafer surface, and further ma-

    terial supply to the individual nanostructures. As it can be

    seen from the above analysis, the nanostructure growth on

    the plasma-exposed surface is determined by many process

    parameters including the plasma density, electron tempera-

    ture, surface temperature, surface diffusion energy, and

    some other parameters. On one hand, this enables many

    interesting possibilities for the growth control by managing

    material supply to the wafer surface and each nanostructure.

    On the other hand, this makes the control very complicated.

    Also, the presence of the surface–plasma sheath limits the

    rate of material supply. Nevertheless, it well exceeds that of

    the neutral gas-based process, and thus the growth rate of the

    plasma-fabricated nanostructures much exceeds the growth

    rates observed in the neutral gas-based process. However,

    the sheath makes it possible to precisely control (via the

    plasma parameters) the material supply to the surface, and

    thus to form a defect-free crystalline structure of the nano-

    structures. This is especially important during formation of

    carbon nanotubes and graphenes, which should in some

    cases exhibit a perfect structure. In general, the energetic

    ions can indeed produce defects in the crystalline structure.

    Nevertheless, in the typical low-temperature plasma proc-

    esses of interest here, the probability of generation of

    ion-induced defects is not high. Indeed, the characteristic

    energies of the onset of defect formation in carbon nanotubes

    and graphenes are quite high and reach �10 eV for thehighly crystalline structures.107 This energy is comparable

    with the typical cross-sheath potential in high-density low-

    temperature plasma processes, e.g., inductively coupled plas-

    mas.55 On the other hand, the ions bombarding the growing

    structure on the surface lose their energy for many processes,

    such as activation and removal of hydrogen usually terminat-

    ing the dangling hydrogen bonds, heating the surface, acti-

    vating atoms of the crystalline structure, and others. As a

    result, most of the impinging ions are not able to produce

    significant defects. Instead, the accelerated ions help activat-

    ing the growing surface and to eventually produce the highly

    crystalline structure. Indeed, highly crystalline carbon struc-

    tures are effectively produced in the plasma.37

    III. ARC DISCHARGE SYSTEMS

    A. Principle of operation and plasma parameters

    The arc discharge in vacuum or controllable gas atmos-

    phere is one of the most promising techniques for the large-

    scale production of various nanostructures with minimal

    structural defects. This method combines a large yield with

    relatively high controllability.108,109

    A typical setup for the arc-assisted nanofabrication is

    shown in Fig. 10. In this system, the temperature of the an-

    ode surface may reach 3000–3500 K.110 An intense erosion

    of electrode material in the discharge spots produces dense

    thermally equilibrium plasma.111 In the discharge core, the

    gas temperature may reach 5000 K. Therefore, the nanostruc-

    tures growing in the plasma can be effectively heated up to

    the temperatures required for the efficient crystal growth.

    The application of an external magnetic field (e.g., by instal-

    lation of the additional magnetic coil or permanent magnet

    in the discharge zone, see Fig. 10) significantly increases the

    plasma density and temperature.114 In this process called

    magnetically enhanced synthesis, the plasma density is

    increased by the two effects: first, by magnetic confinement

    that restricts the plasma in the area close to the electrodes,

    and second, by magnetizing electrons and thus creating con-

    ditions for more effective ionization of the neutral gas

    atoms.112 In turn, the plasma temperature increases in the

    magnetic field due to stronger electric field in the magne-

    tized plasma. This results in the increased heating of the

    nanostructures in the plasma, and hence, an increased flux of

    material to the surface. As a result, fast nucleation and

    growth of the nanostructures is provided. Typical character-

    istics of the arc discharge are summarized in Table II.

    FIG. 10. (Color online) Schematic of the experimental setup and photo of the

    discharge (inset). Graphene flake-containing material was collected from the

    magnet surface (collection area), close to the discharge. A simulated two-

    dimensional map of the magnetic field is shown on the background of the

    discharge. Reprinted with permission from Levchenko et al., Carbon 48,4556 (2010). Copyright 2010, Elsevier Ltd.

    TABLE II. Typical parameters of arc plasmas used for the nanofabrication.

    Discharge power range (kW) 0.1–10

    Plasma density (cm�3) 1011–1014

    Electron energy in discharge (eV) 0.5–1.0

    Discharge voltage (V) 30–100

    Discharge current (A) 50–100

    050801-9 Levchenko et al.: Low-temperature plasmas in carbon nanostructure synthesis 050801-9

    JVST B - Microelectronics and Nanometer Structures

    Author complimentary copy. Redistribution subject to AIP license or copyright, see http://jvb.aip.org/jvb/copyright.jsp

  • B. Graphene production in arc discharge

    In a typical arc-based process,113 the graphene flakes are

    produced using carbon anode and cathode installed in the non-

    magnetic vacuum chamber. The anode is a rod with a hole in

    the center. The hole is filled with a mixture of graphitic car-

    bon and catalyst (for example, Y-Ni powders in 1:4 ratio)

    with the average particle size up to 1 mm. A discharge-

    enhancing cube-shaped permanent magnet installed in the

    chamber creates a nonuniform magnetic field of about 1 kG in

    the interelectrode gap. An interelectrode gap may be varied in

    the range of 1–10 mm. A typical topography of the magnetic

    field is shown in Fig. 10. The process is conducted in a helium

    gas at a pressure of 200–500 Torr.114 More details on the pro-

    cess and setup design can be found elsewhere.113

    Since the gas temperature in the magnetically enhanced

    plasma may reach 5000 K, an intense evaporation of the

    metal catalyst particles emitted to the plasma from the elec-

    trode is sustained.115 In the remote zones of the discharge,

    the gas temperature falls down, the metal catalyst renu-

    cleates into small (several nm in diameter) particles, which

    slowly grow in the plasma. When the temperature falls

    below 1500 K, nucleation and growth of the graphene on the

    small newly nucleated catalyst particles starts.

    Graphene can be also produced in the catalyst-free arc dis-

    charge process. In this case, a hydrogen-containing environment

    should be used to terminate dangling bonds and thus prevent the

    graphene sheets from rolling into nanotubes. Hydrogen can

    effectively reduce oxides and carbides in the plasma, so the pro-

    cess conducted in the hydrogen-enriched environment may

    result in graphene doping with some dopants. On the other hand,

    the catalyzed process in a noble gas environment makes the tech-

    nique highly flexible with respect to any graphene compositions.

    The deposition of the graphene nucleated and grown in

    the arc discharge-based process can be controlled by using a

    specially selected metal catalyst. In the typical experiments,

    a mixture of the two transition metals was used. Yttrium,

    which is a paramagnetic metal capable to easily form car-

    bides and thus enable quick and efficient nucleation of gra-

    phene, and nickel, which is a ferromagnetic metal with the

    Curie temperature of about 350 �C, were used as cata-lysts.112,113 The metal catalyst nanoparticles are nonmag-

    netic in the hot zone of the discharge. Then, the catalyst

    temperature decreases below the Curie point outside of the

    hot discharge zone, the catalyst particles become ferromag-

    netic and change the trajectory in the external magnetic field.

    Finally, these particles with the graphene flakes attached to

    them deposit onto the specific surfaces of the collecting mag-

    net. Therefore, a custom-designed catalyst composition led

    to the graphene deposition onto the magnet surface.108,115

    Figure 11 illustrates the results of characterization of the

    typical graphene flakes collected from the surfaces of the

    collection magnet. The size of the flakes reaches 2000 nm,

    FIG. 11. (Color online) SEM, TEM, AFM, and Raman characterization of the pristine [(a) and (b)] and processed (c)–(g) samples collected from the magnet

    surfaces. Representative SEM images of the carbon nanoparticles containing graphene flakes [(a) and (b)]; SAED pattern of the carbon material (c); TEM

    image of the folding graphene layers (d); 2D [(e) and (f)] and 3D (g) AFM images of two graphene flake samples. Three profiles of the AFM image illustrate

    the flake structure containing several graphene layers (total thickness 2–3 nm); Representative Raman spectra of the graphene flake samples (h). G-peak at

    �1582 cm1 and symmetrical 2D-peak at �2650 cm1 demonstrate the presence of a relatively small number of graphene layers in the flakes. Reprinted withpermission from Levchenko et al., Carbon 48, 4556 (2010). Copyright 2010, Elsevier Ltd.

    050801-10 Levchenko et al.: Low-temperature plasmas in carbon nanostructure synthesis 050801-10

    J. Vac. Sci. Technol. B, Vol. 31, No. 5, Sep/Oct 2013

    Author complimentary copy. Redistribution subject to AIP license or copyright, see http://jvb.aip.org/jvb/copyright.jsp

  • and the flakes consist of typically up to 10 layers. Using the

    atomic force microscopy (AFM), the flakes of about one

    micron with a fairly uniform height in the 1–5 nm range

    (Fig. 11), i.e., of 3–15 graphene layers were revealed. The

    Raman characterization of the collected flakes has demon-

    strated the presence of a D-peak at around 1325 cm�1, which

    indicates relatively small amount of the structural defects in

    sp2 bonds. The G-peak at 1582 cm�1 and highly symmetrical2D-peak at 2650 cm�1 evidence the occurrence of a rela-

    tively small number of graphene layers.

    The production yield of the graphene flakes in the arc dis-

    charge can be estimated by the anode ablation rate reaching

    (3–5)� 10�3 kg�min�1.108 Besides, the higher productionyield can be reached by increasing the arc discharge current,

    since the carbon supply rate increases linearly with the cur-

    rent.110 Therefore, the graphene production can be scaled up

    to 1� 10�3 kg� h�1m�2 when the arc current of 100–150 Ais used.

    C. Synthesis and separation of graphene and carbonnanotubes in arc discharge

    The arc discharge plasma can be also used for the large-

    scale production and simultaneous separation of the carbon

    nanotubes and graphene flakes.116 In this case, a specially

    designed discharge enhancing/separation magnetic unit

    (DESMU) was used to collect the carbon nanostructures.

    Also, the nanostructures were collected from the chamber

    walls. In Fig. 12, one can see the schematic of the setup,

    photographs of the discharge, as well as topography of the

    magnetic field produced by DESMU and SEM images of the

    typical carbon nanomaterials—ropes of nanotubes and gra-

    phene fragments. The nanotubes and graphene flakes were

    deposited on the magnet surfaces only, whereas lacey carbon

    was found only on the chamber walls.

    The size of the graphene flakes grown in such process

    reaches 2500 nm, with the crystallographic faceting clearly

    visible on some of the flakes. The flakes are usually covered

    with amorphous carbon. The typical size of the metal par-

    ticles found in the TEM images reached 10 nm. SEM charac-

    terization of the deposits was used to roughly estimate the

    production rate, which reaches 1� 10�4 m2� h�1 of gra-phene for the discharge current of 50 A, or 0.02 cm2 h�1A�1.

    The numerous multiwalled carbon nanotubes were also

    found in samples collected elsewhere on the chamber walls.

    D. Growth of nanostructures in arc plasmas—theoretical examination

    Now we will discuss the probable mechanism of the gra-

    phene flakes and carbon nanotubes growth of in the plasma of

    the magnetic field-enhanced arc. The theoretical and experi-

    mental studies have proven the following scenario for the for-

    mation of large multilayered flakes in the plasma. At the first

    stage, a single layer graphene flake nucleates on the metal cat-

    alyst particle and grows to a larger radius without formation

    of any additional layers. At the second stage, the single layer

    graphene flake reaches a critical size and a new layer nucle-

    ates, thus forming a two-layer flake. Next, the base layer and

    a newly nucleated second graphene layer grow together and

    after some time a new (third) graphene layer is formed on the

    top. This process stops when the flake is deposited on the col-

    lection surface. It is quite evident that the main condition for

    the formation of a large graphene flake consisting of a few

    graphene layers is the low density of carbon atoms adsorbed

    on the flake during the growth in the plasma. When the atom

    density on the flake is low, the deposited carbon atom evapo-

    rates from the flake growing in the plasma before another

    atom deposits onto the flake. As a result, no nucleation of a

    new layer is possible, and the graphene flake grows as a

    single-layer fragment. If the flux of carbon atoms from the

    plasma to the surface of the graphene flake exceeds the flux of

    carbon evaporation, the carbon adatoms are accumulated on

    the flake surface, the density of adatoms increases and a new

    graphene layer nucleates. The critical graphene flake size can

    be estimated by taking into account the main processes on the

    graphene flake growing in the plasma. Specifically, the carbon

    flux to the flake from the plasma, carbon flux from the flake

    by surface diffusion to the flake edges, carbon evaporation

    from the flake to the plasma bulk, and removal of the carbon

    adatoms by direct impact of the gas molecules deposited onto

    the flake from the plasma should be considered.

    The flux of carbon material to the surface of a single-

    layer graphene flake from the plasma can be estimated as

    WS ¼ VincS; (12)

    where Vi is the ion velocity, nc is the density of carbon ionsin plasma, and S is the graphene flake surface area.Assuming S¼ pR2 and the ion velocity equals to the Bohmvelocity VB¼ (kTe/mi)

    1=2, where R is the graphene flake ra-dius, Te is the electron temperature, mi is the carbon ionmass, and nc¼ np where np is the plasma density, one canobtain

    WS ¼ pR2npffiffiffiffiffiffiffiffiffiffiffiffiffiffikTe=mi

    p: (13)

    Next, the flux of carbon adatoms to the flake edges, We,can be estimated using the random walk distance relation

    R2¼DS t, where DS is the surface diffusion coefficient, and tis the characteristic diffusion time. The surface diffusion

    coefficient can be written as

    DS ¼ ½k2a�0=4� � expð�ed=kTÞ; (14)

    where ed is the surface diffusion activation energy, and T isthe graphene flake surface temperature. Combining the two

    above described formulas one can obtain the expression for

    the characteristic time of adatom diffusion on the graphene

    flake

    t ¼ ½4R2=k2a�0� � expðed=kTÞ: (15)

    Since the flux of carbon adatoms to the flake edges is

    �esc¼ 1/t, one can obtain

    We ¼ ½k2a�0=4R2� � expð�ed=kTÞ: (16)

    050801-11 Levchenko et al.: Low-temperature plasmas in carbon nanostructure synthesis 050801-11

    JVST B - Microelectronics and Nanometer Structures

    Author complimentary copy. Redistribution subject to AIP license or copyright, see http://jvb.aip.org/jvb/copyright.jsp

  • Substituting an expression for the lattice oscillation fre-

    quency �0¼ 2kT /h, where h is the Planck’s constant, onecan finally obtain an estimate for the We as

    We ¼ ½kTk2=2hR2� � expð�ed=kTÞ: (17)

    The carbon adatoms are also removed from the graphene

    flake by the impact of gas molecules depositing from plasma

    with an energy exceeding the carbon evaporation energy. The

    frequency of this process is �C¼V0 n0 k2� exp(�ea/kT0), whereV0¼ (2kT0/m)

    1=2 is the gas atom velocity, m is gas atom mass, eais the surface evaporation energy, n0¼P0/(kT0), P0 and T0 arethe gas density, pressure, and temperature. Thus, one can obtain

    Wi ¼ k2P0ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi2=mkT0

    p� expð�ea=kT0Þ: (18)

    The most important parameter of the graphene flake, the

    critical radius, can be calculated by the balance between the

    four main fluxes, i.e., WS¼WP þWe þWi, where WP is theevaporation. It can be concluded from the above model that

    the graphene flake temperature, plasma density, and the elec-

    tron temperature in the plasma are the key parameters which

    specify the size of a single layer graphene flake. The calcula-

    tions made for the typical arc plasma density of 1019 m�3 and

    the two typical electron temperatures of 0.1 and 0.5 eV have

    shown that the above model reasonably predicts the size of the

    graphene flakes grown in the arc discharge-based process.108

    FIG. 12. (Color online) Experimental setup, photo of the plasma reactor and discharge, and SEM micrographs of representative graphene flakes. (a), (b)

    Representative SEM images of the carbon deposit collected from different collection areas. Ropes of carbon nanotubes found on the top and side surfaces of

    the DESMU, in the areas close to the discharge; graphene layers found on the top and side surfaces of the DESMU, in the areas remote from the discharge. An

    effective separation of the two different carbon nanostructures was ensured. (c) Schematic of the experimental setup. (d) Photograph of the experimental setup.

    (e) Schematic of the mutual position of the cube-shaped magnet, anode and cathode, and the computed 2D map of the magnetic field (field strength of 1.2 kG

    in the discharge gap was optimized for the highest yield of both graphene and nanotubes. (f) Consecutive photographs of the discharge development in the

    nonuniform magnetic field. At the first moments after ignition (t1), the discharge is localized between the cathode and anode. Later (t2, t3, and t4), plasma

    plume becomes unsymmetrical and pulls (due to the interaction with magnetic field) toward the magnet of the discharge enhancing/separation magnetic unit.

    Reprinted with permission from Volotskova et al., Nanoscale 2, 2281 (2010). Copyright 2010, The Royal Society of Chemistry.

    050801-12 Levchenko et al.: Low-temperature plasmas in carbon nanostructure synthesis 050801-12

    J. Vac. Sci. Technol. B, Vol. 31, No. 5, Sep/Oct 2013

    Author complimentary copy. Redistribution subject to AIP license or copyright, see http://jvb.aip.org/jvb/copyright.jsp

  • Similar dependencies can be obtained for the growth of the

    carbon nanotubes in arc discharge plasmas, assuming that the

    nanotubes grow on the partially molten metal catalyst particle

    supplied to the plasma from the ablated electrode (Fig. 13).

    The metal particle immersed in the plasma is a subject of the

    significant heating and ablation, which reduce the particle size

    and eventually lead to the formation of an external liquid

    shell. The carbon atoms deposit on the surface of the particle,

    diffuse through the liquid shell, and eventually incorporate

    into the nanotube. The ion flux at the sheath border–nanotube

    surface supplies carbon atoms to the nanotube. Upon recombi-

    nation, carbon adatoms migrate about the nanotube surface

    and eventually reach the molten catalyst shell or re-evaporate

    to the plasma bulk. Presently, the two growth scenarios are

    accepted: the vapor–liquid-solid117,118 and solid–liquid–-

    solid.119 Both scenarios include the diffusion of the carbon

    atoms in the molten shell, and thus the carbon supply to the

    external catalyst surface is a decisive factor that determines

    the growth kinetics of the nanostructure. To evaluate the car-

    bon supply to the catalyst, one should use a diffusion model

    described in the Sec. II F devoted to the growth of carbon

    nanostructures on the surface.

    To understand the experimental fact of the nanotube length

    increase in the magnetically enhanced discharge, one should

    take into account that at higher plasma densities (up to for

    5� 1018 m�3) the sheath around the nanotube is narrow, andthe ion focusing becomes much stronger [Fig. 13(b)]. The

    number of ions deposited directly on the catalyst surface

    increases, and the adatoms should pass a longer distance along

    the nanotube. As a result, the re-evaporation loss decreases,

    and the length increases.113 At lower plasma density, the

    sheath between the plasma and surface of the nanotube is

    large, and the ion flux is distributed nearly uniformly on the

    surface. The catalyst immersed in the plasma adsorbs some

    fraction of the total ion flux deposited on the nanotube, and

    the rest atoms re-evaporate from the nanotube surface.

    IV. CONCLUDING REMARKS

    A. Unique features of the plasma environment

    As compared with the conventional neutral gas-based

    process environment, the plasma-based process offers sev-

    eral features important for the nanofabrication. The principal

    advantage is very high (and highly controllable) energy den-

    sity in the process area, which allows effective control over

    the energy and material supply to the nanostructures.

    Moreover, high energy density ensures unique possibilities

    for the bulk synthesis of various precursors, which can be

    used for the further deposition onto the nanostructures grow-

    ing on the surfaces, and for the nanostructure nucleation and

    growth in the plasma bulk. Normally, an external heating of

    the growth surfaces can be avoided in the plasma-based pro-

    cess, thus keeping the substrate material intact.

    One more significant feature of the plasma-based tech-

    nique is the involvement of the charged particles (ions and

    electrons) and strong electric field in the process. As a result,

    proper selection of the process parameters such as the

    plasma density, electron temperature, and surface potential

    could be a powerful tool for the precise control of the mate-

    rial deposition onto nanostructures. Magnetic fields could

    also be used for controlling the process parameters (plasma

    density and temperature) and trajectories of the growing

    nanostructures moving in the plasma bulk. These ways of

    control are impossible in the neutral gas-based techniques.

    A very high plasma density in the arc discharge makes it

    very promising for the large scale, bulk production of vari-

    ous nanoparticles and nanomaterials, including highly

    demanded nanotubes and graphene. In contrast, neutral proc-

    esses are usually based on the surface nucleation and growth,

    which are substantially two-dimensional and hence, exclude

    most of the process environment from participation in the

    nanoscale synthesis and processing.

    One more important feature of the plasma-based tech-

    nique is the possibility to selectively produce specific carbon

    allotropes during the growth. Indeed, several carbon allotro-

    pies are important for various applications. In some special

    cases, a complex combined platform involving two or more

    allotropes should be created in a single process (e.g., amor-

    phous carbon and dense pattern of carbon nanotubes for the

    immobilization and protection of the biocatalysts).120 A low-

    temperature plasma-based process provides unique possibil-

    ities to fabricate such sophisticated platforms. Moreover, a

    possibility to fabricate and separate several carbon allotropes

    (e.g., carbon nanotubes and graphenes) was discussed above.

    It should be noted that it is not easy to specify the key pa-

    rameter for every specific carbon allotropy to emerge in any

    FIG. 13. (Color online) (a) Growth of single-walled nanotube (SWNT) on

    molten metal catalyst particle in a plasma. Carbon flux from the plasma is

    nonuniformly distributed about nanotube and catalyst particle surface.

    Carbon adatoms diffuse to the catalyst end, and then incorporate into nano-

    tube structure through the molten catalyst shell. The nanotube diameter is

    determined by the size of the molten catalyst core. Reprinted with permis-

    sion from Keidar et al., Appl. Phys. Lett. 92, 043129 (2008). Copyright2008, American Institute of Physics. (b) Distribution of carbon ion flux on

    nanotube surface with the plasma density and nanotube length as parame-

    ters. Reprinted with permission from Keidar et al., J. Appl. Phys. 103,094318 (2008). Copyright 2008, American Institute of Physics.

    050801-13 Levchenko et al.: Low-temperature plasmas in carbon nanostructure synthesis 050801-13

    JVST B - Microelectronics and Nanometer Structures

    Author complimentary copy. Redistribution subject to AIP license or copyright, see http://jvb.aip.org/jvb/copyright.jsp

  • particular system. Nevertheless, the plasma environment

    allows precise control over the most important parameters

    such as the gas and surface temperatures and the rates of ma-

    terial delivery and removal, and eventually, to adjust the syn-

    thesis process for the specific allotrope. In particular, high-

    purity graphenes were produced by controlling the plasma

    parameters such as the electron temperature (e.g., raising it

    up to �2 eV) using an external magnetic field and henceincreasing both the surface temperature and the material

    delivery rate; the same setup without magnetic field was ca-

    pable to produce only the carbon nanotubes.108 The process

    conducted at lower discharge current and low electron tem-

    perature produces only amorphous carbon.

    B. Surface-based versus bulk growth of thenanostructures in plasmas

    Along with the common features such as a high energy

    density and involvement of charged species and electric fields,

    various kinds of plasmas offer different parameters and hence

    are better suited for either surface-based (inductively coupled

    and microwave discharges) or bulk (arc discharge) growth of

    the nanostructures. The main differences in the ICP/MWP

    and arc plasmas are the operating pressure and plasma density

    (see Tables I and II). The density of arc plasmas may reach

    �1014 cm�3 versus �1012–1013 cm�3, which is common forICP discharges, and the highest operating pressure for the arc

    discharge reaches 1 atm and even higher for some arc types.

    As a result, quite different conditions for the nucleation and

    growth are present in ICP/MWP and arc plasmas. Relatively

    low material flux in ICP/MWP systems is a provision for their

    preferential use in the surface-based techniques where accu-

    mulation of the material on the substrate surface is a prerequi-

    site for the nucleation and the following growth of

    nanostructures. In this case, the growth is due to material

    fluxes delivered directly from the plasma and also from the

    surface accumulating all material supplied from the plasma-

    surface sheath. Other applications of the ICP/MW plasmas

    involve treatment of the nanostructures, modification of their

    crystalline structures, and postprocessing of the surface-bound

    nanostructures for their activation, etc.

    In contrast, the arc plasmas appear to be promising for the

    nucleation and growth of the nanostructures in the bulk. In this

    case, material is supplied from the plasma and accumulated

    only on the surface of the growing nanostructures (no material

    is supplied from the plasma-immersed substrate by surface dif-

    fusion), and thus, the high density of the plasma is the key rea-

    son why the bulk nanosynthesis in arc discharge is effective.

    Along with this, arc plasmas can also be used for the fabrica-

    tion of nanostructured and hard films and coatings, e.g., nano-

    crystalline diamond. Moreover, process control using magnetic

    field is one more feature of the arc discharge, which makes it

    possible to simultaneously fabricate and separate different

    nanostructures directly in the process environment.

    C. Plasma benefits for other nanomaterials

    Though this review paper is mainly focused on the fabrica-

    tion of carbon-based nanostructures in the low-temperature

    plasma environment, the plasma benefits for the fabrication of

    other (noncarbonous) nanomaterials are also worth of mention-

    ing. Among others, the ability of plasma-based processes to pro-

    duce highly crystalline materials such as single-crystal silicon

    nanostructures121 and single-crystalline SiC/AlSiC core-shell

    nanostructures122,123 should be stressed. Very high rate of

    energy supply to the growth area and highly controllable mate-

    rial supply ensure the formation of the surface-based nanostruc-

    tures of the highest quality with the high level of

    morphodimensional selectivity.124,125 The electric-field driven

    self-organization on solid surfaces is one more unique feature of

    the plasma-based processes.64,126,127 In particular, the material

    fluxes on the surface, which are the main driving force of the

    self-organization of surface-based nanostructures such as quan-

    tum dots and nanocones, can be effectively controlled by the

    electric field and eventually, by the plasma parameters.128–130

    D. Future trends in the development of plasma-basedsystems

    Analysis of the plasma-based techniques for nanoscale

    synthesis demonstrates that the plasma has several features

    as a nanofabrication environment such as good controllabil-

    ity and high energy and material densities in the process vol-

    ume. Owing to these features, many unique nanostructures

    have been synthesized. It is quite natural to suppose that the

    controllability and high output of the plasma-based systems

    can be further improved. Specifically, the following trends in

    the development of the plasma-based systems for nanofabri-

    cation should be noted:

    1) Enhancing controllability of the existing systems by the

    development of new control tools, instruments, and meth-

    ods enabling sophisticated management of power and

    particle fluxes to the substrates and nanostructures;

    2) Development of new systems ensuring even higher level

    of controllability, especially in managing the precise dep-

    osition of the fabricated nanostructures onto surfaces, as

    well as plasma–chemical synthesis of new compounds.

    As an outlook it should be pointed out that one of the

    most critical areas where research is needed is basic under-

    standing of the plasma-based technique by utilizing

    advanced experimental techniques and numerical

    simulations.131,132

    ACKNOWLEDGMENTS

    This work was partially supported by CSIRO’s OCE

    Science Leadership Research Program, CSIRO Sensors and

    Sensor Network TCP, and the Australian Research Council.

    1K. S. Novoselov, D. Jiang, F. Schedin, T. J. Booth, V. V. Khotkevich, S.

    V. Morozov, and A. K. Geim, Proc. Natl. Acad. Sci. U.S.A. 102, 10451(2005).

    2I. W. Frank, D. M. Tanenbaum, A. M. van der Zande, and P. L. McEuen,

    J. Vac. Sci. Technol. B 25, 2558 (2007).3S. Iijima, Nature (London) 354, 56 (1991).4M. Keidar, Y. Raitses, A. Knapp, and A. Waas, Carbon 44, 1022 (2006).5H. Park, S. Choi, S.l Lee, and K. H. Koh, J. Vac. Sci. Technol. B 22,1290 (2004).

    050801-14 Levchenko et al.: Low-temperature plasmas in carbon nanostructure synthesis 050801-14

    J. Vac. Sci. Technol. B, Vol. 31, No. 5, Sep/Oct 2013

    Author complimentary copy. Redistribution subject to AIP license or copyright, see http://jvb.aip.org/jvb/copyright.jsp

    http://dx.doi.org/10.1073/pnas.0502848102http://dx.doi.org/10.1116/1.2789446http://dx.doi.org/10.1038/354056a0http://dx.doi.org/10.1016/j.carbon.2005.10.008http://dx.doi.org/10.1116/1.1740760

  • 6S. Kumar, I. Levchenko, M. Keidar, and K. Ostrikov, Appl. Phys. Lett.

    97, 151503 (2010).7R. Ritikos, S. A. Rahman, S. M. A. Gani, M. R. Muhamad, and Y. K.

    Yap, Carbon 49, 1842 (2011).8M. Mannsberger, A. Kukovecz, V. Georgakilas, J. Rechthaler, F. Hasi,

    and G. Allmaier, Carbon 42, 953 (2004).9Z. Tsakadze, K. Ostrikov, J. D. Long, and S. Xu, Diam. Relat. Mater. 13,1923 (2004).

    10M. C. Kan, J. L. Huang, J. C. Sung, and D. F. Lii, J. Vac. Sci. Technol. B

    21, 1216 (2003)11A. Kumar, D. K. Avasthi, A. Tripathi, L. D. Filip and J. D. Carey, J. C.

    Pivin, J. Appl. Phys. 102, 044305 (2007).12Y. C. Sui, B. Z. Cui, R. Guardian, D. R. Acosta, L. Martinez, and R.

    Perez, Carbon 40, 1011 (2002).13D. J. Norris and E. S. Aydil, Science 338, 625 (2012).14C. Li, Y. Chen, Y. Wang, Z. Iqbal, M. Chhowalla, and S. Mitra, J. Mater.

    Chem. 17, 2406 (2007).15P. V. Kamat, J. Phys. Chem. Lett. 4, 1727 (2013).16P. Avouris, Chem. Phys. 281, 429 (2002).17A. Star, Y. Lu, K. Bradley, and G. Grener, Nano Lett. 4, 1587 (2004).18T. Rueckes, K. Kim, E. Joselevich, G. Y. Tseng, C. -L Cheung, and C.

    M. Lieber, Science 289, 94 (2000).19D. Li and R. B. Kaner, Science 320, 1170 (2008).20P. J. Wessely, F. Wessely, E. Birinci, B. Riedinger, and U. Schwalke,

    J. Vac. Sci. Technol. B 30, 03D114 (2012).21J. R. Lake, A. Cheng, S. Selverston, Z. Tanaka, J. Koehne, M.

    Meyyappan, and B. Chen, J. Vac. Sci. Technol. B 30, 03D118 (2012).22M. Peckerar et al., J. Vac. Sci. Technol. B 29, 011008 (2011).23R. A. Barton, J. Parpia, and H. G. Craighead, J. Vac. Sci. Technol. B 29,

    050801 (2011).24V. V. Deshpande, H.-Y. Chiu, H. W. Ch. Postma, C. Miko, L. Forro, and

    M. Bockrath, Nano Lett. 6, 1092 (2006).25K. Bazaka, M. V. Jacob, R. J. Crawford, and E. P. Ivanova, Acta

    Biomater. 7, 2015 (2011).26K. Bradley, J. Gabriel, A. Star, and G. Gruner, Appl. Phys. Lett. 83, 3821

    (2003).27P. G. Collins, K. Bradley, M. Ishigami, and A. Zettl, Science 287, 1801 (2000).28F. Li, Y. Huang, Q. Yang, Z. Zhong, D. Li, L. Wang, S. Song, and C.

    Fan, Nanoscale 2, 1021 (2010).29T. Manabe, S. Nitta, S. Abo, F. Wakaya, and M. Takai, J. Vac. Sci.

    Technol. B 31, 02B110 (2013).30Y. J. Kim, T. S. Shin, H. D. Choi, J. H. Kwon, Y. C. Chung, and H. G.

    Yoon, Carbon 43, 23 (2005).31G. G. Jernigan et al., J. Vac. Sci. Technol. B 30, 03D110 (2012).32A. K. Chakraborty, J. Jacobs, C. Anderson, C. J. Roberts, and M. R. C.

    Hunt, J. Appl. Phys. 100, 084321 (2006).33J. Zheng, R. Yang, L. Xie, J. Qu, Y. Liu, and X. Li, Adv. Mater. 22, 1451

    (2010).34M. J. Behr, E. A. Gaulding, K. A. Mkhoyan, and E. S. Aydil, J. Vac. Sci.

    Technol. B 28, 1187 (2010).35W.-H. Chiang and R. M. Sankaran, Nat. Mater. 8, 882 (2009).36Z. L. Tsakadze, I. Levchenko, K. Ostrikov, and S. Xu, Carbon 45, 2022

    (2007).37K. Ostrikov, J. D. Long, P. P. Rutkevych, and S. Xu, Vacuum 80, 1126

    (2006).38I. Levchenko, U. Cvelbar, M. Modic, G. Filipič, X. X. Zhong, M.

    Mozetič, and K. Ostrikov, J. Phys. Chem. Lett. 4, 681 (2013).39K. Ostrikov, E. C. Neyts, and M. Meyyappan, Adv. Phys. 62, 113 (2013).40K. Ostrikov, Rev. Mod. Phys. 77, 489 (2005).41H. Wang and J. J. Moore, J. Vac. Sci. Technol. B 28, 1081 (2010).42H. W. Wei, K. C. Leou, M. T. Wei, Y. Y. Lin, and C. H. Tsai, J. Appl.

    Phys. 98, 044313 (2005).43S.-F. Lee, Y.-P. Chang, and L.-Y. Lee, J. Vac. Sci. Technol. B 26, 1765

    (2008).44I. Levchenko, S. Y. Huang, K, Ostrikov, and S. Xu, Nanotechnology 21,

    025605 (2010).45K. S. Subrahmanyam, L. S. Panchakarla, A. Govindaraj, and C. N. R.

    Rao, J. Phys. Chem. C 113, 4257 (2009).46Z. J. Han et al., J. Phys. D: Appl. Phys. 44, 174019 (2011).47D. H. Seo, S. Kumar, A. E. Rider, Z. Han, and K. Ostrikov, Opt. Mat.

    Express 2, 700 (2012).48I. Levchenko, K. Ostrikov, M. Keidar, and U. Cvelbar, J. Phys. D: Appl.

    Phys. 41, 132004 (2008).

    49I. Levchenko and K. Ostrikov, Nanotechnology 19, 335703 (2008).50I. Levchenko and K. Ostrikov, J. Phys. D: Appl. Phys. 40, 2308 (2007).51D. M. Nothern and J. M. Millunchick, J. Vac. Sci. Technol. B 30, 060603

    (2012).52S. S. Coffee, S. K. Stanley, and J. G. Ekerdt, J. Vac. Sci. Technol. B 24,

    1913 (2006).53S. Kumar, I. Levchenko, K. Ostrikov, and J. A. McLaughlin, Carbon 50,

    325 (2012).54I. Levchenko, K Ostrikov, and A. B. Murphy, J. Phys. D: Appl. Phys. 41,

    092001 (2008).55I. Levchenko, K. Ostrikov, M. Keidar, and S. Xu, J. Appl. Phys. 98,

    064304 (2005).56P. Jiang, S. Kar, and Y. Zhou, Phys. Plasmas 20, 012126 (2013).57I. Levchenko, K. Ostrikov, M. Keidar, and S. Xu, Appl. Phys. Lett. 89,

    033109 (2006).58I. Levchenko, K. Ostrikov, D. Mariotti, and A. B. Murphy, J. Appl. Phys.

    104, 073308 (2008).59S. Xu, K. N. Ostrikov, Y. Li, E. L. Tsakadze, and I. R. Jones, Phys.

    Plasmas 8, 2549 (2001).60T. Okumura, Phys. Res. Intern. 2010, 164249 (2010).61H.-W. Huang, C.-C. Kao, T.-H. Hsueh, C.-C. Yu, C.-F. Lin, J.-T. Chu,

    H.-C. Kuo, and S.-C. Wang, Mater. Sci. Eng. B 113, 125 (2004).62B. J. M. Hausmann, M. Khan, Y. Zhang, T. M. Babinec, K. Martinick,

    M. McCutcheon, P. R. Hemmer, and M. Loncar, Diam. Relat. Mater. 19,621 (2010).

    63L. V. Nang and E.-T. Kim, J. Electrochem. Soc. 159, K93 (2012)64I. Levchenko, K. Ostrikov, K. Diwan, K. Winkler, and D. Mariotti, Appl.

    Phys. Lett. 93, 183102 (2008).65H. Deutsch, H. Kersten, S. Klagge, and A. Rutscher, Contrib. Plasma

    Phys. 28, 149 (1988).66H. Kersten and G. M. W. Kroesen, J. Vac. Sci. Technol. A 8, 38 (1990).67H. Kersten, H. Deutsch, H. Steffen, G. M. W. Kroesen, and R. Hippler,

    Vacuum 63, 385 (2001).68T. W. Ebbesen, P. M. Ajayan, H. Hiura, and K. Tanigaki, Nature 367,

    519 (1994).69V. Derycke, R. Martel, M. Radosavljevic, F. M. Ross, and P. Avouris,

    Nano Lett. 2, 1043 (2002).70X. M. H. Huang, R. Caldwell, L. Huang, S. C. Jun, M. Huang, M. Y.

    Sfeir, and S. P. O’Brien, J. Hone, Nano Lett. 5, 1515 (2005).71B. Liu, W. Ren, L. Gao, S. Li, S. Pei, C. Liu, C. Jiang, and H. M. Cheng,

    J. Am. Chem. Soc. 131, 2082 (2009).72S. Huang, Q. Cai, J. Chen, Y. Qian, and L. Zhang, J. Am. Chem. Soc.

    131, 2094 (2009).73P. J. F. Harris, Carbon 45, 229 (2007).74X. Li, A. Cao, Y. J. Jung, R. t. Vajtai, and P. M. Ajayan, Nano Lett. 5,

    1997 (2005).75P. Laveant, P. Werner, G. Gerth, and U. Gosele, Sol. State Phenom.

    82–84, 189 (2002).76S. Helveg, C. Lopez-Cartes, J. Sehested, P. L. Hansen, B. S. Clausen, J.

    R. Rostrup-Nielsen, F. Abild-Pedersen, and J. K. Norskov, Nature 427,426 (2004).

    77J. Pezoldt, Y. V. Trushin, V. S. Kharlamov, A. A. Schmidt, V. Cimalla,

    and O. Ambacher, Nucl. Instrum. Methods Phys. Res. B 253, 241 (2006).78Z. Han, B. Tay, C. Tan, M. Shakerzadeh, and K. Ostrikov, ACS Nano 3,

    3031 (2009).79J. M. Tirado, D. Nezich, X. Zhao, J. W. Chung, J. Kong, and T. Palacios

    J. Vac. Sci. Technol. B 28, C6D11 (2010)80R. M. Westervelt, Science 320, 324 (2008).81A. K. Geim, Science 324, 1530 (2009).82M. Hiramatsu, K. Shiji, H. Amano, and M. Hori, Appl. Phys. Lett. 84,

    4708 (2004).83K. Ostrikov, J. Phys. D: Appl. Phys. 44, 174003 (2011).84Y. F. Huang et al., Nature Nanotechnol. 2, 770 (2007).85S. Kumar and K. Ostrikov, Nanoscale 3, 4296 (2011).86D. Mariotti, V. �Svrček, and D.-G. Kim, Appl. Phys. Lett. 91, 183111 (2007).87S. Hacohen-Gourgy, I. Diamant, B. Almog, Y. Dubi, and G. Deutscher,

    Appl. Phys. Lett. 99, 172108 (2011).88A. E. Rider, S. Kumar, S. A. Furman, and K. Ostrikov, Chem. Commun.

    48, 2659 (2012).89N. Behabtu et al., Nature Nanotechnol. 5, 406 (2010).90M. Keidar, J. Phys. D 40, 2388 (2007).91J. Shieh, S. Ravipati, F.-H. Ko, and K. Ostrikov, J. Phys. D: Appl. Phys.

    44, 174010 (2011).

    050801-15 Levchenko et al.: Low-temperature plasmas in carbon nanostructure synthesis 050801-15

    JVST B - Microelectronics and Nanometer Structures

    Author complimentary copy. Redistribution subject to AIP license or copyright, see http://jvb.aip.org/jvb/copyright.jsp

    http://dx.doi.org/10.1063/1.3502562http://dx.doi.org/10.1016/j.carbon.2011.01.006http://dx.doi.org/10.1016/j.carbon.2003.12.049http://dx.doi.org/10.1016/j.diamond.2004.06.010http://dx.doi.org/10.1116/1.1574045http://dx.doi.org/10.1063/1.2767227http://dx.doi.org/10.1016/S0008-6223(01)00230-5http://dx.doi.org/10.1126/science.1230283http://dx.doi.org/10.1039/b618518ehttp://dx.doi.org/10.1039/b618518ehttp://dx.doi.org/10.1021/jz400902shttp://dx.doi.org/10.1016/S0301-0104(02)00376-2http://dx.doi.org/10.1021/nl049337fhttp://dx.doi.org/10.1126/science.289.5476.94http://dx.doi.org/10.1126/science.1158180http://dx.doi.org/10.1116/1.4711128http://dx.doi.org/10.1116/1.4712537http://dx.doi.org/10.1116/1.3524906http://dx.doi.org/10.1116/1.3623419http://dx.doi.org/10.1021/nl052513fhttp://dx.doi.org/10.1016/j.actbio.2010.12.024http://dx.doi.org/10.1016/j.actbio.2010.12.024http://dx.doi.org/10.1063/1.1619222http://dx.doi.org/10.1126/science.287.5459.1801http://dx.doi.org/10.1039/b9nr00401ghttp://dx.doi.org/10.1116/1.4790518http://dx.doi.org/10.1116/1.4790518http://dx.doi.org/10.1016/j.carbon.2004.08.015http://dx.doi.org/10.1116/1.3701700http://dx.doi.org/10.1063/1.2360774http://dx.doi.org/10.1002/adma.200903147http://dx.doi.org/10.1116/1.3498737http://dx.doi.org/10.1116/1.3498737http://dx.doi.org/10.1038/nmat2531http://dx.doi.org/10.1016/j.carbon.2007.05.030http://dx.doi.org/10.1016/j.vacuum.2006.01.025http://dx.doi.org/10.1021/jz400092mhttp://dx.doi.org/10.1080/00018732.2013.808047http://dx.doi.org/10.1103/RevModPhys.77.489http://dx.doi.org/10.1116/1.3497030http://dx.doi.org/10.1063/1.1993776http://dx.doi.org/10.1063/1.1993776http://dx.doi.org/10.1116/1.2981084http://dx.doi.org/10.1088/0957-4484/21/2/025605http://dx.doi.org/10.1021/jp900791yhttp://dx.doi.org/10.1088/0022-3727/44/17/174019http://dx.doi.org/10.1364/OME.2.000700http://dx.doi.org/10.1364/OME.2.000700http://dx.doi.org/10.1088/0022-3727/41/13/132004http://dx.doi.org/10.1088/0022-3727/41/13/132004http://dx.doi.org/10.1088/0957-4484/19/33/335703http://dx.doi.org/10.1088/0022-3727/40/8/S11http://dx.doi.org/10.1116/1.4754563http://dx.doi.org/10.1116/1.2221318http://dx.doi.org/10.1016/j.carbon.2011.07.060http://dx.doi.org/10.1088/0022-3727/41/9/092001http://dx.doi.org/10.1063/1.2040000http://dx.doi.org/10.1063/1.4789986http://dx.doi.org/10.1063/1.2222249http://dx.doi.org/10.1063/1.2996272http://dx.doi.org/10.1063/1.1343887http://dx.doi.org/10.1063/1.1343887http://dx.doi.org/10.1155/2010/164249http://dx.doi.org/10.1016/j.mseb.2004.07.004http://dx.doi.org/10.1016/j.diamond.2010.01.011http://dx.doi.org/10.1149/2.082204jeshttp://dx.doi.org/10.1063/1.3012572http://dx.doi.org/10.1063/1.3012572http://dx.doi.org/10.1002/ctpp.2150280206http://dx.doi.org/10.1002/ctpp.2150280206http://dx.doi.org/10.1116/1.576403http://dx.doi.org/10.1016/S0042-207X(01)00350-5http://dx.doi.org/10.1038/367519a0http://dx.doi.org/10.1021/nl0256309http://dx.doi.org/10.1021/nl050886ahttp://dx.doi.org/10.1021/ja8093907http://dx.doi.org/10.1021/ja809635shttp://dx.doi.org/10.1016/j.carbon.2006.09.023http://dx.doi.org/10.1021/nl051486qhttp://dx.doi.org/10.4028/www.scientific.net/SSP.82-84.189http://dx.doi.org/10.1038/nature02278http://dx.doi.org/10.1016/j.nimb.2006.10.058http://dx.doi.org/10.1021/nn900846phttp://dx.doi.org/10.1116/1.3516649http://dx.doi.org/10.1126/science.1156936http://dx.doi.org/10.1126/science.1158877http://dx.doi.org/10.1063/1.1762702http://dx.doi.org/10.1088/0022-3727/44/17/174003http://dx.doi.org/10.1038/nnano.2007.389http://dx.doi.org/10.1039/c1nr10860chttp://dx.doi.org/10.1063/1.2805191http://dx.doi.org/10.1063/1.3657146http://dx.doi.org/10.1039/c2cc17326chttp://dx.doi.org/10.1038/nnano.2010.86http://dx.doi.org/10.1088/0022-3727/40/8/S18http://dx.doi.org/10.1088/0022-3727/44/17/174010

  • 92J. Shieh, F. J. Hou, Y. C. Chen, H. M. Chen, S. P. Yang, C. C. Cheng,

    and H. L. Chen, Adv. Mater. 22, 597 (2010).93S. Kumar, I. Levchenko, Q. J. Cheng, J. Shieh, and K. Ostrikov, Appl.

    Phys. Lett. 100, 053115 (2012).94M. Wolter, M. Haass, T. Ockenga, J. Blazek, and H. Kersten, Plasma

    Process. Polym. 6, S620 (2009).95Ch. Hollenstein, Plasma Phys. Controlled Fusion 42, R93 (2000).96S.-H. Hong and J. Winter, J. Appl. Phys. 100, 064303 (2006).97M. Wolter, I. Levchenko, H. Kersten, and K. Ostrikov, Appl. Phys. Lett.

    96, 133105 (2010).98Y. Watanabe, J. Phys.