buck minster fuller en discovery and consequences

Upload: markobonjak6860

Post on 05-Apr-2018

222 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/2/2019 Buck Minster Fuller En Discovery and Consequences

    1/23

    Review

    Synthesis of aligned carbon nanotubes

    Choon-Ming Seah a, Siang-Piao Chai b, Abdul Rahman Mohamed a,*

    a School of Chemical Engineering, Engineering Campus, Universiti Sains Malaysia, Seri Ampangan, 14300 Nibong Tebal, S.P.S. Pulau Pinang,

    Malaysiab School of Engineering, Monash University, Jalan Lagoon Selatan, 46150 Bandar Sunway, Selangor, Malaysia

    A R T I C L E I N F O

    Article history:

    Received 27 December 2010

    Accepted 25 June 2011

    Available online 30 June 2011

    A B S T R A C T

    Vertically aligned carbon nanotubes (ACNTs) are bundles of carbon nanotubes oriented per-

    pendicular to a substrate, and horizontally aligned CNTs are parallel to the substrate. Their

    dense and orderly arrangement, along with outstanding physical and chemical properties,

    enables ACNTs to be used in various fields. The methods of synthesising ACNTs can be

    classified into single-step and double-step techniques. Thermal pyrolysis and flame syn-

    thesis are the common single-step methods, and both are relatively simple. The double-

    step methods, including catalyst coating and chemical vapour deposition, provide more

    control over the catalyst morphology. This review explores different methods used for -

    ACNT growth, the process parameters that determine the morphology of ACNTs and the

    applications of structured ACNTs.

    2011 Elsevier Ltd. All rights reserved.

    Contents

    1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4614

    2. Single-step methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4615

    2.1. Thermal pyrolysis method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4615

    2.1.1. Parameters determining the properties of ACNTs synthesised by thermal pyrolysis . . . . . . . . . . . . . . 4616

    2.2. Flame synthesis. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4617

    3. Double-step methods. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4619

    3.1. Physical vapour deposition (PVD) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4619

    3.1.1. Pretreatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46203.2. Solution-based catalyst precursors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4621

    3.3. Catalyst . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4623

    3.4. Chemical vapour deposition (CVD) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4624

    3.5. Supergrowth . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4625

    4. Substrates and buffer layers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4625

    5. Alignment of CNTs. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4627

    6. Horizontally aligned carbon nanotubes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4627

    7. Towards mass production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4628

    0008-6223/$ - see front matter

    2011 Elsevier Ltd. All rights reserved.doi:10.1016/j.carbon.2011.06.090

    * Corresponding author: Fax: +60 594 1013.E-mail address: [email protected] (A.R. Mohamed).

    C A R B O N 4 9 ( 2 0 1 1 ) 4 6 1 34 6 3 5

    a v a i l a b l e a t w w w . s c i e n c e d i r e c t . c o m

    j o u r n a l h o m e p a g e : w w w . e l s e v i e r . c o m / l o c a t e / c a r b o n

    http://dx.doi.org/10.1016/j.carbon.2011.06.090mailto:[email protected]://dx.doi.org/10.1016/j.carbon.2011.06.090http://dx.doi.org/10.1016/j.carbon.2011.06.090http://dx.doi.org/10.1016/j.carbon.2011.06.090http://www.sciencedirect.com/http://www.elsevier.com/locate/carbonhttp://www.elsevier.com/locate/carbonhttp://www.sciencedirect.com/http://dx.doi.org/10.1016/j.carbon.2011.06.090http://dx.doi.org/10.1016/j.carbon.2011.06.090http://dx.doi.org/10.1016/j.carbon.2011.06.090mailto:[email protected]://dx.doi.org/10.1016/j.carbon.2011.06.090
  • 8/2/2019 Buck Minster Fuller En Discovery and Consequences

    2/23

    8. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4629

    Acknowledgements. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4629

    References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4629

    1. Introduction

    The discovery of buckminsterfullerene (C60) in 1985 by Krotoet al. [1] has led to an entirely new branch of carbon chemis-

    try. In the early 1990s, another significant breakthrough was

    achieved. Carbon filaments with diameters in the nanometre

    range were observed by Iijima using transmission electron

    microscopy (TEM) [2]. These carbon filaments were called

    carbon nanotubes (CNTs). Two years later, single-walled car-

    bon nanotubes (SWCNTs) were synthesised by Iijima and

    Ichihashi [3] and Bethune et al. [4]. The research on CNTs

    subsequently began in earnest worldwide. Aligned carbon

    nanotubes (ACNTs) were first reported by Thess et al. [5],

    who were able to bundle 70% of the volume of nanotubes into

    crystalline ropes in 1996. In the same year, the Chinese Acad-

    emy of Science reported that a 50 lm-thick film of highlyaligned nanotubes had successfully been grown by chemical

    vapour deposition (CVD) [6]. Ren and Huang[7] first used Plas-

    ma-Enhanced Hot Filament CVD (PE-HF-CVD) to lower the

    growth temperature to below 666 C and used electric field

    as an external force to provoke the alignment. Meanwhile,

    Fan et al. [8] introduced position controlled growth of VACNT

    on porous and plain silicon substrate. They also reported the

    detailed growth and mechanism of alignment of ACNTs in

    their work.

    Vertically aligned CNTs (Fig. 1) are quasi-dimensional car-

    bon cylinders that align perpendicular to a substrate [9]. Ver-

    tically aligned with high aspect ratios [10] and uniform tube

    length made it easy spinning into macroscopic fibres [11].The arrays of ACNT arrays are typically grown from a catalyst

    that is pinned to a substrate, which produces long, high-pur-

    ity nanotubes with sidewalls that are free of catalysts [12]. Be-

    cause of these properties, arrays of ACNT are widely used in

    nanoelectronics or in composite materials as reinforcing

    agents. Large arrays of ACNTs with a high degree of unifor-

    mity in terms of tip radius and height provide excellent field

    emission properties [13,14]. Furthermore, vertically aligned

    CNTs also exhibit a high capability to produce high current

    densities under low operating voltages [15]. CNTs are formed

    from a sheet of graphene, which possesses a strongly aniso-

    tropic structure. Chiral nanotubes could be envisioned in

    which the current-carrying state may have an angularmomentum about the tube axis, making them appropriate

    for flat panel displays. ACNTs can also be reinforced with

    other matrix composites to form anisotropic conductive

    materials [16]. ACNTs synthesised on substrates with a pat-

    terned trench structure enables them to be applied in ad-

    vanced triode-type field emitters [17]. ACNTs have a very

    large surface area and a high thermal conductivity, both of

    which facilitate rapid heat transfer to the surrounding, mak-

    ing them important materials in the construction of solar

    cells [18]. ACNTs have also been used in hydrogen storage,

    as the interior and interstitial surfaces of open-ended CNTs

    have a strong binding energy for adsorbing hydrogen gas mol-

    ecules compared with planar carbon surfaces [19,20]. Alignedmulti-walled carbon nanotubes (MWCNTs) were found to

    possess a higher adsorption rate of hydrogen than non-

    aligned CNTs because of the large inter-nanotube space in be-

    tween the parallel nanotubes [20]. The subnanometre pores of

    ACNTs are suitable for separation of gases and other small

    molecules, such as hydrogen and water.

    ACNTs can be potentially used in self-cleaning applica-

    tions because of their surface property to be hydrophobic

    [21]. Furthermore, ACNTs, possessing larger surface area

    and higher electrical conductivity over entangled CNTs, are

    ideal electrode material for DNA biosensor [22], sensors for

    glucose [23], pH [24] as well as NO2 [25]. A research group from

    Tsinghua University [26] had grown super-aligned CNT arrayswhich are greater in nucleation density, lower CNT diameter

    distribution and better alignment compared to ordinary ACNT

    arrays [26]. These super-aligned CNTs were successfully spun

    into continuous yarns with excellent mechanical and electri-

    cal properties [2628], which can be further developed to a

    touch panel [29], liquid crystal display [30] and transparent

    loudspeaker [31]. The property study shows that the yarns

    possessed mechanical strength greater that 460 MPa. The

    flexibility and strength retained even though the yarns were

    exposed to very high or very low temperatures [32]. Instead

    of yarns, a transparent CNTs sheet also can be drawn in par-

    allel from ACNTS arrays and can be employed to make organ-

    ic light emitting diode [33]. Furthermore, it was shown thatACNTs used in a complementary metal-oxide semiconductor

    (CMOS) integrated circuit may overcome the problem of large

    device-to-device variation when normal CNTs are used [34].

    The ACNT array with super-compressible foam-like behav-

    ior and quick recovery properties are suitable for use as en-

    ergy absorbing coating [35]. In addition, the outstanding

    mechanical properties along with the adhesive strength pos-

    sessed by the entangle structures at the top of arrays can be

    developed to dry adhesive which could withstand up to

    100 N cm1 shear force. The normal adhesion force is low

    and that makes them easy for lifting off [36].

    Since the first ACNT array was reported in 1996 [6], numer-

    ous papers describing the growth of ACNTs have beenFig. 1 A scanning electron microscopy (SEM) image

    showing ACNTs [19].

    4614 C A R B O N 4 9 ( 2 0 1 1 ) 4 6 1 34 6 3 5

    http://www.elsevier.com/locate/carbonhttp://www.elsevier.com/locate/carbonhttp://www.elsevier.com/locate/carbonhttp://www.elsevier.com/locate/carbonhttp://www.elsevier.com/locate/carbonhttp://www.elsevier.com/locate/carbonhttp://www.elsevier.com/locate/carbonhttp://www.elsevier.com/locate/carbon
  • 8/2/2019 Buck Minster Fuller En Discovery and Consequences

    3/23

    published. Over more than a decade, various modifications

    have been made and new techniques have been discovered

    that make it possible to grow high-quality ACNTs with scal-

    able production. Large CNT arrays have successfully been

    grown on different substrates, such as mesoporous silica

    [37], planar silicon substrates [38] and quartz glass plate [16]

    by using transition metals such as Fe, Co, or Ni as the catalyst.

    Because transition metals have non-filled d shells, it enables

    them to adsorb and interact with hydrocarbons. Of all of the

    synthesis methods, CVD has been recognised as the most

    promising method for producing ACNT arrays. Various types

    of CVD have been developed, such as thermal CVD (T-CVD),

    plasma-enhanced CVD (PE-CVD) and floating catalyst CVD

    (FC-CVD).

    2. Single-step methods

    2.1. Thermal pyrolysis method

    Thermal pyrolysis is also known as metalorganic CVD or FC-

    CVD [39]. It is one of the most popular methods for synthes-

    ising dense and aligned forms of CNTs. FC-CVD involves the

    pyrolysis of organometallic precursors such as ferrocene

    [13,38,4053], iron (II) phthalocyanine (FePc) [5458], iron

    pentacarbonyl [59], nickelocene and cobaltocene [60] to nucle-

    ate the growth of nanotubes. Non-carbonaceous compounds,

    such as FeCl3, are also reported to be promising catalyst pre-

    cursors for growing ACNTs [61]. In most cases, a carbon

    source must be added in excess to increase the carbon-

    to-catalyst ratio and prevent high levels of metal impurity

    in the CNTs [52]. Aromatic hydrocarbons such as xylene

    [38,4043], toluene [44], benzene [45,46] and naphthalene

    [47] are often used along with ferrocene because of their

    chemical structure similarities [47] and the fact that most of

    the aromatic hydrocarbons can dissolve ferrocene easily.

    However, the use of heavy hydrocarbons such as aromatic

    hydrocarbons and cyclohexane is not suitable because heavy

    hydrocarbons will deposit on the reactor wall in a low-tem-

    perature zone [53]. As a result, lighter hydrocarbons such as

    acetonitrile [48], ethylene [49,53], acetylene [50] and alkanes

    [51] are commonly used. In addition, tree products such as

    turpentine oil [13] and camphor [52] are also used as carbon

    sources for synthesising ACNTs.

    After many years of research, FC-CVD has been modified

    with the goal of growing ACNTs of better quality and align-

    ment. In general, there are two major types of reactors used

    for FC-CVD, namely double-furnace and single-furnace

    reactors. In the double-furnace setup (Fig. 2), the first furnace

    is responsible for the vaporisation and sublimation of the cat-

    alyst precursor, while the second furnace is kept at a high

    temperature for the catalysts to assemble and nucleate the

    growth of CNTs. One of the drawbacks of this process is the

    steep temperature gradient that exists between the two

    furnaces, which makes it difficult to maintain the same evap-

    oration rate throughout the entire process. As for the conven-

    tional single-furnace setup, only one high-temperature

    furnace is required. The mixture of the catalyst and carbon

    feedstock (liquid phase) is first evaporated using a heater be-

    fore it is introduced into the reactor. The problem with this

    approach lies in controlling the uniformity of the catalyst par-

    ticles inside the reactor. Aerosol-assisted carbon deposition

    has been developed to overcome this shortcoming. Jeong

    et al. [40] used an ultrasonic evaporator to atomise a mixture

    of ferrocene and xylene. The mixture was then carried into a

    single-furnace reactor by a carrier gas. Spray pyrolysis with

    the use of a spray nozzle to atomise the mixture supply com-

    ing into the reactor has also been reported [13,46]. Both meth-

    ods can continuously generate quantitatively controlled

    aerosols in large amounts, ensuring that the carbon source

    and the metal particles are distributed evenly in the reactor.

    The growth of CNTs from FC-CVD has been suggested to

    occur in two different ways. First, the active metals must de-

    posit on the substrate before the growth of CNTs can take

    place. Chen and Yu [55] supported this point with the obser-

    vation of the existence of metal at both ends of the tube, sug-

    gesting the co-existence of the tip-growth and base-growth

    mechanisms. Li et al. [56] found that CNTs could be grown

    from an iron film deposited on substrate that contained dif-

    ferent sizes of iron nanoparticles. The larger iron particles

    were responsible for producing the carbon atomistic species,

    which were required for subsequent growth of the CNTs.

    Meanwhile, the smaller iron particles were more catalytically

    active because of their higher surface energy. The graphite

    layers formed could encapsulate the iron particles of both

    sizes and form a concentric graphitic shell or a semi-spherical

    graphitic shell. The continuous generation of carbon atoms

    by the larger iron particles increased the length of the graph-

    ite layers by forcing the large iron particles up while the small

    iron particles remained on the substrate, which is why both

    ends of the tube contained iron particles. However, the study

    of Huang et al. [62] found that the growth of nanotubes

    started on the active catalyst at the floating stage. These

    authors provided more convincing proof than did the previ-

    ously mentioned authors. In their study, four substrates were

    High temperature furnaceLow temperature furnace

    Carrier gas/ vector gas

    Organometallic and hydrocarbon

    Fig. 2 Scheme showing the double-furnace setup used in the organometallic/hydrocarbon co-pyrolysis process.

    C A R B O N 4 9 ( 2 0 1 1 ) 4 6 1 34 6 3 5 4615

  • 8/2/2019 Buck Minster Fuller En Discovery and Consequences

    4/23

    placed at different locations in the reactor with temperatures

    ranging from 200 C to 900 C. Nanotubes were found on the

    substrate in the temperature range of 200 C to 300 C. It is

    known that this range of temperature is too low for growing

    nanotubes. Furthermore, nanotubes were found on Au-coated

    wafers, but Au is not a good substrate for growing CNTs. Be-

    sides, the U-shaped nanotubes (Fig. 3) are believed to grow

    in the floating phase and then become hooked on the sub-

    strate, forming a U shape as a result of the flowing gas.

    The thermal pyrolysis method is attracting considerable

    attention for the synthesis of ACNTs because of its exception-

    ally low cost and ease of scaling up to mass production

    [47,49,62]. The ACNTs can be grown on flat substrates and

    also on cylindrical substrates such as the wall of a tube reac-

    tor. Moreover, ACNTs are also reported to grow on spherical

    substrates consisting of 50% SiO2, 30% Al2O3, and 20% ZrO2with a diameter of 700 lm (Fig. 4) [53].

    2.1.1. Parameters determining the properties of ACNTs

    synthesised by thermal pyrolysis

    The density and alignment of the CNTs are intimately related

    to various process parameters. The parameters that are

    widely studied by researchers in order to produce

    high-quality, well-aligned CNTs are the growth temperature,

    the catalyst concentration, the feed rate and the period of

    growth [45,63,64].

    The growth temperature is the most crucial parameter for

    determining the properties of CNTs. From our review, the

    common temperatures for the growth of ACNTs are in the

    range of 700950 C. At a relative low temperature, alignment

    could not be obtained because of the incomplete dissociation

    of catalysts precursors and hydrocarbon species [65]. The coa-

    lescence kinetics between the catalyst and the hydrocarbon

    species are not sufficient to guarantee the alignment and

    crystallinity of the nanotubes [13]. Carbon nanofibres (CNFs)

    consisting of defective graphitic qualities such as herringbone

    morphology are always found at low reaction temperatures.

    At a relatively high temperature, the catalyst can lose its cat-

    alytic activity. In addition, the dissolution rate of carbon

    atoms in the catalyst is higher than that of the diffusion

    and precipitation rates at high reaction temperature, which

    causes carbon atoms to accumulate on the surface of cata-

    lysts, forming multi-shelled carbon nanocapsules (MS-CNCs)

    [42].

    The diameter of the ACNTs is strongly correlated to the

    diameter of the catalyst cluster [64]. At higher temperatures,

    the mobility of the catalyst on the substrate is relatively high,

    promoting the coalescence and formation of larger catalyst

    nanoparticles. Hence, a higher temperature encourages the

    growth of CNTs of larger diameter [44]. However, other

    authors have reported that the average diameter of CNTs in-

    creases with temperature in a lower temperature range and

    decreases with temperature in a higher temperature range,

    which is common in the continuous catalyst feeding process.

    Normally, the size and mobility of the catalyst particles in-

    crease with increasing temperature. However, when the CNTs

    start to grow, the floating catalyst will have a more difficult

    time reaching the substrate because it becomes blocked by

    the growth of CNT arrays [63]. The catalyst particles intro-

    duced after this process has occurred will deposit on the tube

    wall and nucleate the formation of smaller diameter nano-

    tubes, resulting in a bimodal diameter distribution. However,

    the bimodal diameter distribution does not exist in the low-

    temperature range [44]. There is another possible explanation

    for this phenomenon, which contrasts with the previous one.

    In the high-temperature range, the energy supply is high,

    which enhances the full decomposition of the catalyst precur-

    sor to form fine catalyst particles that grow CNTs of smaller

    diameter [66,67].

    The CNT growth rate is also linearly proportional to the

    temperature, a feature caused by the increased mobility of

    the floating carbon species. In addition, ACNTs become

    straighter with increasing temperature. Lee et al. [68] studied

    the quality of ACNTs synthesised at different temperatures

    using Raman spectroscopy and high-resolution TEM. They

    found that the degree of crystalline perfection increased line-

    arly with temperature. It is speculated that high temperatures

    will increase the diffusion rate of carbon and create graphitic

    sheets with fewer defects. However, after the optimum tem-

    perature is reached, the growth rate of CNTs decreases with

    increasing temperature. As a result, the carbon deposits, cov-

    ering the entire catalyst surface and deactivating the catalyst.

    The size of the catalyst nanoparticles is the most influen-

    tial factor for controlling the diameter of the CNTs produced

    Fig. 3 U-shaped CNTs (the arrow indicates the flow

    direction) [62].

    Fig. 4 Vertically aligned CNTs grown on a spherical

    substrate approximately 700 lm in diameter [53].

    4616 C A R B O N 4 9 ( 2 0 1 1 ) 4 6 1 34 6 3 5

  • 8/2/2019 Buck Minster Fuller En Discovery and Consequences

    5/23

    [64,69]. The diameter and the yield of CNTs are always

    strongly correlated with the catalyst concentration [40,63]. A

    high catalyst/hydrocarbon ratio will boost the agglomeration

    of the catalyst particles, resulting in the formation of larger

    catalyst particles. When the catalyst/hydrocarbon ratio is

    too high, adverse effects will be observed. The catalyst parti-

    cles agglomerate at a high rate until they becomes too large

    and are no longer active for growing CNTs (low surface-to-vol-

    ume ratio of the catalyst nanoparticles). Nanolumps and

    branched CNTs are favoured in this condition [66]. The CNTs

    found in this condition are generally grown on the smaller-

    sized catalyst particles that escaped from the rapid coales-

    cence [45]. Furthermore, carbon deficiency inhibits the growth

    of new tube walls [70]. If the reaction temperature is high, a

    bimodal diameter distribution will also appear. A higher cat-

    alyst/hydrocarbon ratio results in the metal catalyst sticking

    either on the wall or inside the nanotubes, which affects

    the structural perfection of the tube wall. Longer and thinner

    tubes are reported in a low Fe/C environment because of the

    abundance of carbon and additional active catalyst [71]. The

    smaller catalyst particles are much more active and last long-

    er. At a very low introduction rate of catalyst precursor, the

    yield is less and it is impossible for alignment to occur. Con-

    versely, at a high introduction rate, the catalyst particles coa-

    lesce together and ultimately yield MS-CNCs [46], or the

    catalyst particles may be retained inside the tube and pro-

    mote the growth rate and length of the CNTs [41].

    Tapaszto et al. [63] used a spray nozzle to carry the feed

    (low active solutiontovector gas ratio) at very high flow rate

    into the reactor. This approach contributes to the formation

    of ACNTs with a narrower distribution of diameters. The

    droplets of the liquid reactant involved are small if the liquid

    reactant is sprayed at a very high rate, producing longer tubes

    with a smaller diameter. The droplets decompose more easily

    and promote the growth of ACNTs. In fact, a low retention

    time of the feed causes the decomposition of hydrocarbon

    to be less effective, and pyrolytic coating occurs, which might

    cause the reactor wall to be covered with a sticky layer of con-

    tamination if decomposition of heavy hydrocarbons is

    involved.

    A very high vector gastocarbon source ratio in the reac-

    tor will dilute the concentration of the catalyst precursor

    and carbon sources, producing a low yield of CNTs with poor

    alignment [72]. Meanwhile, Li et al. [56] and Huang et al.[73]

    used a low vector flowrate (

  • 8/2/2019 Buck Minster Fuller En Discovery and Consequences

    6/23

    rod-like probes [7579], grids [8083] or plates [8486]. The use

    of metalorganics such as iron pentacarbonyl [87,88], ferro-

    cene [89] or metal nitrate dissolved in fuel [90,91] has been re-

    ported. So far, ACNTs have only been reported to grow on

    alloy substrates.

    Flame synthesis is a very energy-efficient process because

    the fuel itself is a source of heat and carbon. The temperature

    achieved can be as high as 1600 K, which is hard to achieve

    with CVD in a conventional furnace. For ACNT synthesis over

    a large area, it is more economical to use flame rastering and

    multiple flames to achieve a controllable residence time and

    the desired flame region [84,92]. The production efficiency

    and yield per energy input is lower than that of CVD, making

    this method suitable for industrial production [82]. Addition-

    ally, flame synthesis is a simple one-step method that with-

    out substrate preparation. However, complex substrate

    preparation has also been reported for flame synthesis

    [93,94]. The growth mechanism of ACNTs in the flame synthe-

    sis of hydrocarbon can be divided into three main steps

    [75,82,93,94]. First, the hydrocarbon fuel is pyrolysed in the

    preheated zone to form hydrocarbon species, which will be

    the carbon source for the CNTs, while the metal particles

    form on the surface of the alloy. The hydrocarbon species dif-

    fuses onto the catalyst at an appropriate temperature and is

    then absorbed by the catalyst to nucleate the growth of CNTs.

    The choice of the catalyst is the main consideration for the

    production of the ACNTs. An alloy is always selected because

    of its lower melting point and higher solubility for carbon

    compared with pure metal. From our review, a majority of

    the ACNTs are reported to be grown on alloys containing Ni,

    along with Fe and Cr [77,78,81,84]. It is believed that the for-

    mation of nickel oxide contributes to the growth of CNTs

    [62]. Nevertheless, the alignment still depends on the density

    of the metal nanoparticles, which serve as catalysts. Pan et al.

    [84] studied different kind of alloys using an ethanol flame

    and the authors found that pure Ni and pure Fe only produced

    CNTs and CNFs, respectively. They proposed a hollow-core

    mechanism for Ni. The diffusivity of carbon at the exterior

    surface of Ni is more rapid than in the interior of Ni nanopar-

    ticles, resulting in the growth of CNTs with a hollow core in

    the ethanol flame synthesis. In contrast, a solid-core mech-

    anism applies to Fe because carbon can easily diffuse

    through the Fe particle to form CNFs. Arana et al. [75] also

    found the same outcome. The solubility of carbon in Ni parti-

    cles was higher compared with Fe particles, so carbon would

    precipitate more rapidly in Ni. Hence, long, dense and well-

    aligned CNTs are obtained in the presence of Ni. Xu et al.

    [77,78] studied the growth of CNTs on different alloys for both

    co-flow and counterflow methane diffusion flames. They

    found that ACNTs could only be grown on Ni/Fe/Cr alloys.

    The authors suggested that both Fe and Ni were necessary

    for the growth of ACNTs. ACNTs are also reported to grow

    on Co-coated stainless steel [95]. However, aerosol metal

    organic catalysts show a completely different outcome to that

    mentioned earlier [89,90].

    The regions of the flame in which ACNTs can be grown are

    very limited [83]. Different locations have different carbon

    species concentrations and temperature profiles, which

    determine the morphology of the CNTs formed. The forma-

    tion of CNTs usually happens in the visible orange soot zone

    of a normal diffusion flame [96]. Yuan et al. [83] found that the

    yellowish flame was the best place for the CNTs to grow. They

    suggested that the temperature distinguished the carbon rad-

    ical species that contributed to the formation of CNTs. Higher

    temperatures were promoting the formation of CNTs over

    soot. In another study by Yuan et al. [82], the authors found

    that the yield, diameter and height of CNTs increased with

    the sampling height from the nozzle and the temperature.

    Woo Lee et al. [96] reported that the formation of CNTs on

    a catalytic substrate occurred at a location outside of the soot-

    ing zone and the flame front of inverses diffusion flame. Xu

    et al. [78] confirmed this point by growing CNTs at the tip of

    the flame. However, the CNTs were shorter in length than

    those produced inside the flame. CNTs are commonly found

    in the sooting region of different kinds catalytic probes used

    because of the higher temperatures, which promote the for-

    mation of active catalytic nanoparticles for CNT nucleation.

    The studies of non-premixed diffusion flames show that the

    CNTs grow profusely in the areas near to the centreline of

    the flames but not at the centre because the centre of the

    non-premixed feed is lacking in unsaturated carbon species

    and CO that will contribute to the formation of CNTs [78].

    Counterflow diffusion flames may provide a stable one-

    dimensional reaction zone. There are studies [76,79] based

    on the methane flame model (Fig 6(a)) proposed by Beltrame

    et al. [97] that predict the temperature profile and major car-

    bon species, as shown in Fig 6(b). CNTs are grown 89.5 mm

    from the fuel nozzle. ACNTs are only found when an electric

    field is introduced [76,79]. Merchan-Merchan et al. [76] found

    that highly ordered vertical ACNTs were grown in the region

    8.510.0 mm from the fuel nozzle. Xu et al. [77] successfully

    produced ACNTs with a methane flame seeded with

    acetylene without the electric field. The breakup of the alloy

    surface induced by the carbide formed the catalyst nanopar-

    ticles that were responsible for the growth of the CNTs. The

    high density of the catalytic nanoparticles formed facilitated

    the growth of denser CNTs and provided vertical support for

    the nanotubes. In some other studies [76,79], the counterflow

    diffusion flame was also seeded with acetylene, but no ACNTs

    were synthesised, which might be the reason why Xu et al.

    [77] conducted the synthesis process at a higher temperature

    and why this condition enhanced the formation of catalytic

    nanoparticles that could grow CNTs, as compared with those

    reported in [76,79].

    Flame synthesis is less popular than CVD. There are sev-

    eral shortcomings to this method. The apparatus for the

    flame synthesis, especially the burner, is complicated so that

    the morphology of the flame can be controlled. The gaseous

    fuels must be injected safely and carefully. The main draw-

    back to this method is the poor quality of the ACNTs obtained.

    From our review, the majority of the produced ACNTs are not

    straight, except for those synthesised with the aid of an elec-

    tric field. The bean-sproutlike bundles containing encapsu-

    lated particles at the tips of the nanotubes, as shown in

    Fig. 7, are always found. In addition, a certain amount of CNFs

    is present in the array of ACNTs, which limits the application

    in certain fields. The formation of CNTs in the sooting region

    also enables the deposition of carbon other than graphitic car-

    bon on the wall. The lengths of ACNTs that are synthesised in

    the flame synthesis method are relatively short compared

    4618 C A R B O N 4 9 ( 2 0 1 1 ) 4 6 1 34 6 3 5

  • 8/2/2019 Buck Minster Fuller En Discovery and Consequences

    7/23

    with those synthesised from CVD. The study, optimisationand application of ACNTs synthesised by the flame synthesis

    method is limited. From our review, no articles addressed the

    mechanism of the formation of catalyst nanoparticles, and

    characterisation techniques to determine the crystal phases

    of the catalyst have not been reported. The actual growth

    mechanism of the CNTs synthesised from the flame synthe-

    sis method is still unclear. The poor structure and alignment

    of ACNTs synthesised by this method have seriously limited

    the application of ACNTs. However, the low operating cost

    and scalable production enable ACNTs to be applied as a cat-

    alyst support and composite reinforcing material, which do

    not require ACNTs with perfect alignment and high

    crystallinity.

    3. Double-step methods

    The double-step methods involve coating the active catalyst

    onto a substrate, which has the advantage of controlling the

    morphology and distribution of the catalyst particles. In addi-

    tion, the CNT growth location can be controlled with this ap-

    proach. In general, higher-purity ACNTs are produced by

    double-step methods compared with single-step methods.

    There are two general approaches used to coat the catalyst

    onto the substrates, i.e., through physical vapour deposition

    (PVD) and by a solution-based precursor.

    3.1. Physical vapour deposition (PVD)

    PVD is a vacuum deposition process used to coat a thin film of

    catalyst on a substrate by vaporising the catalyst precursor

    and then condensing it onto the substrate surface. Thin-film

    deposition onto a substrate uses the following sequence:

    First, the catalyst to be deposited is vaporised by physical

    means. The vapour is transported in a low-pressure environ-

    ment to the targeted substrate and condensed to form a thin

    catalyst film [98,99]. This process is one of the most popular

    and most efficient methods for preparing nucleation sites

    for growing ACNTs.

    PVD is commonly used for the thin-film coating in ACNT

    synthesis. The size of the catalyst particles can be controlled

    effortlessly by adjusting the film thickness through the depo-

    sition time [100]. However, PVD is a relatively costly method.

    It requires sophisticated equipment. High vacuum and high

    temperature are required for the catalyst deposition, which

    dramatically increases the energy consumption. However,

    PVD still remains the most popular coating method in the

    field of ACNT synthesis.

    The thickness of the catalyst film is closely related to the

    morphology and alignment of the subsequence growth of

    ACNTs. ACNTs will not grow on a continuous catalyst film.

    The film must be broken down through annealing to create

    nanoparticles, better known as islands. The size of the nano-

    particles determines the density, diameter and number of lay-

    ers in the tube wall of the ACNTs [101]. If the thickness of the

    Fig. 7 An SEM image of bean-sproutlike bundles of well-

    aligned CNTs with catalyst nanoparticles lifting off at the tip

    of the CNTs [82].

    Fig. 6 (a) A schematic of the experiment set-up from Merchan-Merchan et al. [76] using a counterflow diffusion flame

    proposed by Beltrame et al. and (b) the numerical predictions of temperature and major chemical species [79,97].

    C A R B O N 4 9 ( 2 0 1 1 ) 4 6 1 34 6 3 5 4619

  • 8/2/2019 Buck Minster Fuller En Discovery and Consequences

    8/23

    catalyst film is too thin, a sparse distribution of metal islands

    will be produced that results in thin ACNTs of low density

    [102]. Conversely, a thick catalyst film will result in sintering

    of the metal catalyst, which prevents the catalyst film from

    breaking into isolated nanoparticles or islands for growing

    ACNTs [103]. Ho et al. [104] studied the effect of 10, 15 and

    25 nm Ni films on quartz glass. As shown in Fig. 8, the thinner

    film provides a more uniform distribution of small-sized cat-

    alyst nanoparticles. As the thickness increased to 15 nm, the

    nanoparticles started to coalesce and formed elongated

    shapes. For the 25 nm Ni film, the coalescence of the nanopar-

    ticles was more extensive and the voids between the nano-

    particles were covered, just like before the annealing step. It

    was reported that dense and vertical ACNTs can only be

    grown on a substrate with catalyst nanoparticles distributed

    in the pattern shown in Fig. 8(a). The optimum thickness of

    the film required is interdependent with the type of substrate,

    the buffer layer, and the annealing environment. The size of

    the metal islands can be controlled through temperature

    and the time of the annealing process. A thicker film requires

    a longer annealing time at high temperature to break the film

    into smaller islands and vice versa [101].

    Chiu et al. [105] and Wu and Chang [106] studied Fe films

    with thicknesses between 0.3 and 3 nm and showed that

    the diameter of CNTs was directly proportional to the film

    thickness. In addition, CNTs with almost identical diameters

    were grown when the same thickness of Fe was used in both

    studies, although the experiments were conducted under dif-

    ferent conditions. The trend did not change even though the

    range of the film thickness was increased to 50 nm [107]. A

    high percentage of SWCNT arrays were reported for an Fe film

    of 0.6 nm, while double-walled carbon nanotubes (DWCNTs)

    were observed on a film of 5 nm [108]. The height of the

    CNT arrays and the growth rate of the CNTs were also affected

    by the film thickness. The growth rate of the CNTs was more

    rapid when thinner films were used because the reactivity of

    smaller-sized catalyst particles was higher and they are more

    difficult to deactivate [105]. It is well known that the align-

    ment of CNTs is induced by van der Waals forces. The larger

    catalyst nanoparticles will grow CNTs with larger diameters.

    Additionally, the van der Waals force is stronger for a tube

    with a larger diameter. This force will restrain the upward

    growth of the CNTs [106]. A thicker film will result in denser

    ACNT arrays, which inhibits the diffusion of carbon to the cat-

    alyst, resulting in a lower CNT growth rate.

    3.1.1. Pretreatment

    In the pretreatment step, the widely studied parameters are

    the use of reducing gas, the time and the temperature [109

    113]. Ammonia and hydrogen are the common reducing gases

    that have always been applied in dry etching processes. The

    metal film will break into small and more uniform nanoparti-

    cles in the presence of ammonia or hydrogen gas [109,110].

    Ammonia decomposes to hydrogen and nitrogen during the

    pretreatment. Hydrogen reduces the metal oxide that

    provides a nucleation site for growing nanotubes. It is well

    documented that in the initial stage, the average size of nano-

    particles decreases but the density increases with etching

    time [109,110]. Prolonged pretreatment at high temperatures

    enhances the possibility of coalescence with neighbouring

    particles to form larger nanoparticles [110]. The inverse effect

    is shown when the etching time is too long; the catalyst is

    etched by the excited hydrogen generated by either hydrogen

    or ammonia. The catalyst nanoparticles are more crystallised

    when the hydrogen flow rate is high [112]. It has also been re-

    ported that metal nitride will be formed in the presence of

    ammonia or nitrogen during the pretreatment stage, which

    enhances the growth of CNTs [114,115]. However, the nitro-

    gen-to-hydrogen ratio must not be too high, otherwise nitro-

    gen will etch away the catalyst [113].

    Microwave or other sources of power can be applied when

    synthesising ACNTs [116]. The processing pressure in this

    treatment must be high enough to provide sufficient flux im-

    pact from the plasma, so that the film will receive more en-

    ergy and momentum transfer to break the catalyst film into

    nanoparticles [113]. The nanoparticle size increases with the

    microwave power, as does the disorder of the subsequence

    CNTs grown. From Fig. 9, it can be seen that hydrogen plasma

    did a better job than ambient hydrogen at breaking the film

    into fine catalyst nanoparticles.

    Diluted hydrofluoric acid (HF) is widely used in wet etching

    to obtain an uneven surface morphology. Lee et al. [118120]

    used both HF dipping and NH3 pretreatment for Ni and Ni

    Co films on SiO2 substrates. The roughness increased with

    dipping time in HF, while NH3 further etched the surface to

    form small domains inside the metal cluster. Choi et al.

    [121] reported that different surface morphologies were

    obtained at different durations of HF dipping. First, HF etched

    away the metal catalyst and increased the surface roughness.

    Microcracks were formed later. The roughness on the sub-

    strate caused by the etching increased the grip of the catalyst

    Fig. 8 SEM micrographs of deposited Ni films after annealing, with thicknesses of (a)10 nm, (b) 15 nm and (c) 25 nm [104].

    4620 C A R B O N 4 9 ( 2 0 1 1 ) 4 6 1 34 6 3 5

  • 8/2/2019 Buck Minster Fuller En Discovery and Consequences

    9/23

    on the surface. This condition provided a better platform for

    the formation of nanoparticles than did a smooth surface.

    3.2. Solution-based catalyst precursors

    To overcome the shortcoming of PVD, solution-based cata-

    lyst precursors have been developed as an alternative way

    to coat active catalysts on the substrate for growing ACNTs.

    The main techniques used to coat the catalyst on the sub-

    strate with solutions are dip coating, spin coating, spray

    coating and microcontact printing. All of these techniques

    are able to distribute the catalyst nanoparticles evenly on

    the substrate, which is the main criterion for obtaining

    ACNTs. In the dip-coating method, the substrate is dipped

    in a solution containing catalyst precursor at a constant

    speed in order to prevent any judder and rippling on the

    surface of the solution. The substrate is immersed in the

    solution and withdrawn at a slow, uniform speed to obtain

    a uniform coating. The volatile solvent is evaporated, leav-

    ing the metal catalyst on the substrate [122]. A thick cata-

    lyst film will be formed if the withdrawal speed is too

    rapid [123]. Vibration damping tools are sometimes required

    to ensure that the liquid surface remains ripple-free to ob-

    tain a homogeneous thickness across the entire substrate

    [124]. In spin coating, a puddle of the solution is placed

    on the substrate at the axis of rotation. The substrate is ro-

    tated at a very high speed to spin away the excess amount

    of solution through centrifugal force. The inertia of the

    solution is the reason that the solution is ejected radially

    outward. The thickness of the solution decreases at high

    spin speeds. A thick film will result in the formation of lar-

    ger-sized catalyst particles, and a thin film leads to the for-

    mation of smaller-sized but denser particles [125]. The

    solvent used in spin coating is almost the same as the solu-

    tion used in dip coating. In the spray-coating system, the

    precursor is atomised at the nozzle by pressure and then

    directed towards the substrate. Microcontact printing, a soft

    lithography technique, has also been applied in the synthe-

    sis of ACNTs. The stamps for microcontact printing are pre-

    pared by curing poly(dimethyl)siloxane. The printing is

    carried out by placing the stamp on the surface of the sub-

    strate, followed by transferring the catalytic materials onto

    the tops of pillars [126]. Usually, microcontact printing is

    applied in the patterned growth of ACNTs.

    The catalyst solution is crucial for determining the mor-

    phology and topology of ACNTs grown from different coating

    techniques. Various types of solutions are prepared for this

    purpose. The most common solution is an alcoholic solution

    containing metal salts. Alcohol is used because of its high vol-

    atility, and the metal salts are those that are able to be diluted

    easily in alcohol. Metal nitrate and metal acetate, which have

    high solubilities in alcohol, are always selected as the catalyst

    precursors [127]. Murakami et al. [128130], Maruyama et al.

    [131] and Hu et al. [132] used 0.01 wt.% cobalt acetate and

    molybdate acetate in ethanol to grow aligned SWCNTs. Metal

    acetate is also used in spin coating[133,134]. One of the weak-

    nesses of the alcoholic solution is that alcohol has a low

    vapour pressure and a low viscosity, which causes the recrys-

    tallisation of salts and forms non-uniform dispersion of the

    catalyst nanoparticles after the drying process. However,

    ethylene glycol can be added to surmount this obstacle.

    Liquid nitrogen has been utilised to freeze-dry the solution

    and prevent the agglomeration of salts that results from rapid

    evaporation of the alcohol, which maintains the uniform

    dispersion of the catalyst [134]. Mauron et al. [125] found that

    an increase in the concentration of the solution increased the

    diameter of the catalyst nanoparticles but decreased the

    density of the catalyst nanoparticles [125]. However, if the

    concentration is too low, no CNTs are formed or just low-

    density entangled ACNTs will be obtained [135].

    Cho et al. [136] and Choi et al. [127] used a magnetic fluid to

    disperse the catalyst nanoparticles on the substrate. Magnetic

    fluids are stable colloidal suspensions composed of single-

    domain magnetic nanoparticles dispersed in proper solvents

    [137], as shown in Fig. 10. The authors [127,136] used decanoic

    acid as the surfactant that dissolved in the acetone. It was

    slowly added to an iron chlorideammonia solution to ensure

    that the surfactant was able to completely cover the surface

    of the Fe3O4 particles so that the repulsion between particles

    was effective. In this process, the polar heads of the primary

    surfactant adsorb onto the catalyst particles, while the non-

    polar tails are exposed to the solvent. The nonpolar tails of

    the secondary surfactants mount onto the tails of the primary

    surfactants through the van der Waals force, leaving the polar

    heads bound to the ammonia and acetone solvent. The steric

    repulsion prevents them from agglomerating and maintains

    the uniform dispersion after coating. Small catalyst islands

    are formed after heat treatment, and these islands are

    Fig. 9 SEM images of Ni/Si substrate after 10 min (a) annealing in H2 ambient (b) etching by H2 plasma at 700 C followed by

    cooling to room temperature in vacuum [117].

    C A R B O N 4 9 ( 2 0 1 1 ) 4 6 1 34 6 3 5 4621

  • 8/2/2019 Buck Minster Fuller En Discovery and Consequences

    10/23

    responsible for growing ACNTs. Polyvinyl alcohol is some-

    times added to increase the viscosity of the solution in order

    to control the density of the nanoparticles and the film thick-

    ness during spin coating[127].

    The solgel method is applied for admixtures of tetraeth-oxysilane, iron nitrate aqueous solution and ethanol, which

    is used for spray coating [138]. Pan et al. [37] mixed tetra-

    ethyl-ortho-silicate, iron nitrate, ethanol and Pluronic P-

    123 triblock copolymer to develop mesoporous silica with

    the catalyst particles distributed evenly on the substrate.

    The addition of the triblock copolymer improved the wet-

    ting ability of the solution on the substrate. The substrate

    was dip-coated into the mixed solution. Liu et al. [139] used

    block copolymers as micelle catalyst templates. Poly(sty-

    rene-block-acrylic acid) (PS-PAA) was dissolved in toluene,

    stirred and heated to convert all of the polymer material

    to the spherical micelle phase. FeCl3 was then added to

    the solution as the catalyst precursor. PS-PAA is an amphi-philic block copolymer that forms micelles in solution,

    which are capable of self-organising into partially ordered

    structures. Fe ions diffused through the thin PS layer before

    exchanging with the H+ ions of the carboxyl group and

    being effectively bound into the PAA core. A quasi-hexago-

    nal monolayer array was obtained within the PS matrix be-

    cause of the self-assembly of the PS-PAA micelles, as shown

    in Fig. 11(a). The catalyst was uniformly dispersed on the

    substrate for the synthesis of ACNTs in a subsequent CVD

    process. NaOH was added to facilitate metal loading. The

    carboxyl acid group in the PAA core underwent hydration

    during neutralisation, causing swelling of carboxyl-contain-

    ing latex. The volume expansion of the PAA domains led to

    the ruptures illustrated in Fig. 11(b) [140]. The catalyst ions

    exchanged directly with the Na+ on the carboxyl group.

    Various parameters can be varied to enhance the disper-

    sion of the catalyst nanoparticles. The concentration of the

    catalyst will affect the diameter of the nanoparticles formed

    on the substrate. After the catalyst is saturated in the solu-

    tion, the size of the catalyst nanoparticles is determined bythe micelle dimensions. The molecular weight of the PS af-

    fected the spacing between the nanoparticles, while PAA

    dominated the size of the catalyst nanoparticles. Diluting

    the solution with PS homopolymer decreased the density of

    the catalyst nanoparticles. Other colloidal solutions such as

    Co nanoparticles have been prepared by dispersing AOT[-

    bis(2-ethylhexyl)-sulfosuccinate]-stabilised Co nanoparticles

    in toluene, and this approach has been used in spin coating

    [142].

    Ryu et al. [143] used polystyrene nanospheres for shadow

    masks, one of the nanosphere lithography techniques used

    for the fabrication of nano-pitched metallic arrays. An or-

    dered monolayer of nanospheres was spin-coated on the sub-strate, and catalyst solution was coated on top of the

    nanopsheres. Catalyst spots were formed on the substrate

    through the triangle voids between the spheres. The size of

    the nanospheres was used to control the size and density of

    the nanoparticles that formed.

    The main drawback of the solution-based precursor ap-

    proach is that preparing the catalyst precursor is cumber-

    some. It may take hours or days to prepare the solution. In

    addition, the solution tends to accumulate in the notch on

    the substrate. In the spin-coating method, the effect of sur-

    face tension will oppose the uniformity and topography of

    coating, while contact printing is only suitable for coating

    small areas. The catalyst accumulates in recessed areas in

    dip coating [144]. The distribution of catalyst nanoparticles

    is not as uniform as that created by the PVD process.

    Aqueous solution of

    FeCl2 and FeCl3

    NH4OH + primary

    surfactant + acetone

    Secondary

    surfactant

    Secondary surfactantprimary surfactant

    Fig. 10 A schematic representation of the synthesis of surfactant bilayer-stabilised magnetic fluids, using fatty acids as

    primary and secondary surfactants to obtain stable aqueous magnetic fluids. The black and hollow dots represent the polar

    heads of the fatty acids [108].

    PS matrixPS matrix

    PAA domainsCavitated PAA domains

    Fig. 11 A diagram of (a) a non-cavitated thin film on a substrate with metal loaded and (b) the cavitation PAA domain with

    NaOH added [140142] (black dots represent iron chloride).

    4622 C A R B O N 4 9 ( 2 0 1 1 ) 4 6 1 34 6 3 5

  • 8/2/2019 Buck Minster Fuller En Discovery and Consequences

    11/23

    Millimetre-scale growth of ACNTs is seldom reported by this

    method. Although different types of coating methods have

    been proposed, extensive studies have not yet been carried

    out.

    3.3. Catalyst

    The most commonly used active catalysts for growing CNTs

    are magnetic elements such as Fe, Co or Ni. It is reported that

    Co is more appropriate than Fe and Ni for producing better-

    quality CNTs [145]. Yuan et al. [146] found that long and

    straight ACNTs can also be grown using Pt, Pd, Mn, Mo, Cr,

    Sn or Au.

    Magnetic elements have a strong tendency to agglomerate

    on the substrate at high temperatures because of their mag-

    netic properties and their high specific surface energy [147].

    Other metal elements areaddedas stabilisers to prevent exces-

    sive agglomeration of the active elements. Molybdenum is one

    of the proven catalyst stabilisers for the growth of CNTs. Mar-

    uyamas research group usedCoMobimetallic catalyst in syn-

    thesising aligned SWCNTs [129132]. According to the authors,

    the affinity of Mo for oxygen is stronger than that of Co, and

    thus Mo tends to form an oxide at the interface of the catalyst

    and the substrate during the calcination process. It is reported

    that Co will diffuse into MoO3 and form a CoMoOx underlayer.

    During the catalyst reduction, CoO and MoO3 will be reduced

    to Co and MoOy (y 6 2). CoMoOx remains unchanged because

    of its stability. The gooddispersionof Co nanoparticles is attrib-

    uted to the strong interaction between metallic Co and Co-

    MoOx. The chemical state and the morphology of CoMo

    catalysts are shown in Fig. 12. Noda et al. [148] found that the

    yield of CNTs was relatively high when the concentration of

    Co was slightly higher than the concentration of Mo. This find-

    ing supports the mechanism proposed by Maruyamas group.

    We have also reported that the presence of CoMoO4 after calci-

    nation of CoOMoO/Al2O3 catalyst plays the same role as previ-

    ously mentioned [149]. The optimum ratio of CoOx to MoOx is

    8:2 (w/w) [150], which deviated from the findings of Maruyama

    and Noda. In addition, CoAl2O4 and Co3O4 were formed for the

    CoOMoO/Al2O3 catalyst after calcination. It was reported that

    the formation of CoO and Co after partial reduction by hydro-

    gen was more effective for growing CNTs [151]. Meanwhile

    MgMoO4 and CoMoO4 were formed after calcination for Co

    Mo/Mg [152]. However in other studies, when CO was intro-

    duced in CoMo, molybdate dissociated to form molybdenum

    carbide and Co particles of smaller size that could grow

    SWCNTs [153155]. Co ions were embedded in the molybdate-

    like cluster after calcination, and it remained unchanged after

    reduction. Continuously introducing CO reduced the Co parti-

    clesto metallic Co, which subsequently aggregated to form lar-

    ger particles that led to the formation of MWCNTs and CNFs.

    Aligned SWCNTs were also grown on FeMo bimetallic cata-

    lysts. The order of deposition of different types of metals has

    been shown to have an influence on the morphology of the

    ACNTs formed. With Mo deposited on Fe, the catalyst resisted

    poisoning at a high hydrocarbonflux duringCVD. Mowas pres-

    ent in the outer portion of the particle, protecting Fe. Mean-

    while for the case in which Fe was deposited on Mo, Fe easier

    to get poisoned [156].

    The capability of a TiCo bimetallic hybrid catalyst to pro-

    duce aligned SWCNTs was reported in [157]. Ti prevented the

    formation of Co-silicate. Sato et al. [158] proposed a mecha-

    nism for growing CNTs with TiCo catalyst. It was shown that

    TiCoincorporated more carbon, in theform of TiCx, compared

    with Co alone.Furthermore,the meltingtemperature of TiCo

    C was lower than that of CoC, which enhancedthe carbonsol-

    ubility in the catalyst. This feature makes it possible to grow

    ACNTs at lower temperatures. The hybrid bimetallic coating

    (co-sputtering of Ti and Co) outclassed the layer coating of

    the catalyst in yield performance, which is attributable to a

    more uniform dispersion of catalyst [159]. CNTs were also pro-

    duced with Ni/TiO2 [160] and Mn/Ni/TiO2 catalysts [161]. Gunji-

    shima et al. [162] used FeV for the production of aligned

    DWCNTs. The authors claimedthat the incorporation of vana-

    dium was able to increase the activity of Fe and that it initiated

    the growth of CNTs before the catalyst particles started to

    aggregate. Cr is another choice of metal that helps in the syn-

    thesis of ACNTs. The role of Cr is to disperse the active metal

    catalyst uniformly, leading to a better alignment of the grown

    CNTs [147]. A large amount of Cr is required for effective pre-

    vention of metal catalyst aggregation. CNTs can be grown on

    CoV, CoFe, CoNi, CoPt, and CoY. However, ACNTs were

    only found on CoV and CoFe [163].

    In the nucleation stage, the carbon needs to dissolve in the

    catalyst particle until it reaches supersaturation to initiate the

    growth of CNTs [74]. Liu et al. [103] co-sputtered iron and

    graphite on the substrate with the goal of growing ACNTs,

    and they found that the alignment and density of ACNTs

    was better than when using pure iron as a catalyst. The pre-

    saturation of carbon allows CNTs to grow with a high density

    Fig. 12 Schematic representation of the changes in the chemical state and the morphology of CoMo catalysts on quartz

    substrates after (a) calcination and (b) reduction [132].

    C A R B O N 4 9 ( 2 0 1 1 ) 4 6 1 34 6 3 5 4623

  • 8/2/2019 Buck Minster Fuller En Discovery and Consequences

    12/23

    and uniform in the initial phase of growth, which causes

    crowding of the ACNTs and yields ACNTs with a better align-

    ment and orientation.

    3.4. Chemical vapour deposition (CVD)

    Besides FC-CVD, which was discussed in Section 2, there

    are three other conventional CVD methods: thermal CVD

    (T-CVD), hot-filament CVD (HF-CVD) and plasma-enhanced

    CVD (PE-CVD). These methods have been proven to be very

    successful in growing ACNTs. PE-CVD outclasses both

    T-CVD and HF-CVD in terms of the alignment of the synthes-

    ised CNTs because of the applied bias to the substrate [164].

    The use of plasma can significantly reduce the activation en-

    ergy of the growth. The lowest temperature reported for the

    successful growth of ACNTs using PE-CVD was 120 C [165].

    However, the ACNTs were poor crystallinity, short and with

    cone-liked shape.

    Self-bias potential is influenced on the surface of the sub-

    strate under plasma conditions, inducing an electric field dur-

    ing the growth of CNTs. The field forces the CNTs to align to

    the direction of the electric field as they grow [166]. The field

    emission of CNT films may be enhanced by treatment with

    hydrogen plasma. The presence of CH bonds in ACNT sam-

    ples is expected in a hydrogen plasma atmosphere [167]. Be-

    sides that, the period of growth of ACNTs can be controlled

    precisely. The growth can be quenched immediately by

    switching off the power supply for plasma generation [168].

    Another advantage of PE-CVD over T-CVD is that PE-CVD is

    highly efficient at gas decomposition and the concentration

    of reactive species can be controlled [101]. Normally, direct

    current (dc), radio frequency (rf) or microwave excitation are

    used to generate plasma. Plasma deposition is stable and

    highly controllable, leading to reproducible growth condi-

    tions, thus lowering the growth temperature [14]. In dc-PE-

    CVD, the degree of alignment is improved with the plasma

    voltage. Below the plasma excitation potential, no alignment

    is observed [169]. The plasma is ignited if the voltage is high

    enough, and intensive ion bombardment is applied to the

    structure, which initiates the growth of ACNTs. Temperature

    control is another crucial factor for successfully using PE-

    CVD. Low temperatures could limit the decomposition of

    hydrocarbon gas, leading to the formation of amorphous car-

    bon. However, if the temperature is too high, CNTs appear in a

    bundle form, and most of the CNTs possess disordered walls

    [169]. Microwave excitationPE-CVD [116] and rf-PE-CVD [170]

    both show the same outcome under high temperature and

    power. The crystallinity of the CNTs could be seriously dam-

    aged because of the ion bombardment [170]. The negatively

    biased substrate tends to attract and accelerate the positively

    charged hydrogen and hydrocarbon ions generated within the

    plasma. The ions become etchants and cause damage to the

    walls of the CNTs [171]. Carbon soot will deposit at the tips

    of the CNTs if the power applied is too high. To solve this

    problem, a metal plate is used to cover the substrate so as

    to prevent the ions from bombarding the substrate [170]. An-

    other drawback of PE-CVD is that this approach involves com-

    plex equipment setup [172]. Nozaki et al. [173] pointed out

    that two issues need to be addressed to improve the growth

    of CNTs with PE-CVD. One of them is the preparation of

    catalyst nanoparticles that do not coagulate extensively while

    maintaining their catalytic function during PE-CVD. The other

    one is the use of remote plasma, which would restrict the ion

    damage to both the catalyst nanoparticles and the CNTs.

    They proposed atmospheric pressure PE-CVD. Under higher

    gas pressure, ions that accelerated towards a substrate would

    undergo collisions with neutral molecules in the sheath and

    the bombardment energy was lower toward the substrate

    and prevents the damage on the ACNTs grown [174]. Besides,

    Nozaki et al.[175] found that SWCNTs can only be grown in

    the atmospheric PE-CVD.

    In general, HF-CVD and T-CVD are more desirable tech-

    niques than PE-CVD for producing ACNTs because the later

    involves high operating costs and sophisticated equipment

    setup. Furthermore, HF-CVD and T-CVD are suitable for the

    irregular-shaped and multiple substrate coatings that cannot

    be used in PE-CVD [67]. HF-CVD and T-CVD are also free from

    the complications resulting from the large amount of undesir-

    able and uncontrollable radicals created in the plasma during

    the growth of CNTs [78]. A combination of HF-CVD and PE-

    CVD has been studied to capitalise on the advantages of both

    methods [7]. However, it was reported that the CNTs were

    quite similar to the CNTs produced from PE-CVD alone [172].

    Low temperature CVD becomes important when the sub-

    strate involves metallic components, such as in CMOS and

    large-scale-integration (LSI) interconnects, in which these

    components will deform at temperatures above 550 C

    [176]. PE-CVD is not suitable due to the damage caused to

    the CNT wall and that increases the electrical resistivity

    and constrains in the applications in electronic industry.

    Normally, hot filament or feedstock preheating are required

    in a low temperature CVD. The temperature of hot filament

    has to be high enough to decompose carbon source into ac-

    tive components such as radicals to initiate the growth of

    ACNTs [177,178]. Higher preheating temperature will pro-

    duce ACNT arrays of better crystallinity and higher arrays

    [179]. Herringbone-liked structure and CNFs are dominant

    if the preheating temperature [180] or reaction temperature

    [176] is too low. The graphite sheet of CNTs from low tem-

    perature CVD either by HF-CVD or feedstock preheating

    method contains more defects and the yield is low as com-

    pared to CNTs grown in a high temperature condition

    [181,182].

    ACNTs can be grown under low temperature CVD with-

    out the use of feedstock preheating or hot filament. The

    lowest temperature had been reported so far was 350 C

    in a simple T-CVD [183]. The author found that the credit

    belongs to the small catalyst nanoparticles, where nucle-

    ation and growth of CNTs can be initiated on the surface

    of the metal particle with proper shape which avoids the li-

    quid catalyst-carbon eutectic phase that requires higher

    temperature. Meanwhile, Mora et al. [184] found that as

    long as the feedstock is decomposed, CNTs will be grown.

    Using carbon feedstock associated with exothermic decom-

    position will lower the growth temperature. However, the

    main drawback of low temperature synthesis is low quality

    ACNTs produced [158,159,184,185]. The room of study in low

    temperature CVD remains wide and efforts have been put

    for improving its efficiency so that it is comparable with

    conventional CVD.

    4624 C A R B O N 4 9 ( 2 0 1 1 ) 4 6 1 34 6 3 5

  • 8/2/2019 Buck Minster Fuller En Discovery and Consequences

    13/23

    3.5. Supergrowth

    The growth of ACNTs in CVD is always limited by the wrap

    of amorphous carbon on the surface of the catalyst. It is dif-

    ficult to obtain millimetre-scale ACNTs by CVD. Hata and

    co-workers [186] achieved a massive breakthrough by inject-

    ing a small amount of weak oxidiser, i.e., water vapour, dur-

    ing the CVD reaction. The growth rate of ACNTs was

    stunning: an ACNT array 2.5 mm in height was grown within

    10 min, and an ACNT/catalyst weight ratio of above 50,000%

    was achieved. This new CVD method is denoted as super-

    growth. Li et al. [187] almost doubled the height of the ACNT

    forest to 4.7 mm in 2 h of reaction time. The main purpose of

    using a weak oxidant in CVD is to prevent deposition of amor-

    phous carbon over the catalyst nanoparticles and to prolong

    the catalytic lifetime [186188]. The work of Yamada et al.

    [189] shows that the deposition of amorphous carbon was

    the dominant factor of terminating CNTs growth in CVD,

    water vapour and oxidizer played an important role in sus-

    taining supergrowth. Introducing water vapour also increased

    the selectivity towards SWCNTs. According to Amama et al.

    [190], injecting water vapour in the reaction may hold back

    the Ostwald ripening causing large catalyst nanoparticles to

    grow larger due to the sintering effect of smaller catalyst par-

    ticles during annealing and CVD. In addition, the wall struc-

    ture is significantly improved by introducing a small

    amount of water vapour during CVD [191194]. It has been re-

    ported that with the assistance of water vapour, the aspect ra-

    tio of CNTs was greatly increased and a purity of 99.9% and

    above was successfully achieved [186,191,195]. At that level

    of purity, the purification process can be omitted, as it is well

    understood that purification may change the morphology and

    cause defects to the ACNT arrays.

    The effect of water vapour is interdependent on other

    parameters involved in the CVD process. The optimum

    amount of water vapour injected is closely related to the

    hydrocarbon flowrate [196]. Higher flowrates of hydrocarbon

    will deactivate the catalyst by forming encapsulated carbon

    over the active catalyst. Water vapour inhibits this process,

    and thus a very high initial growth rate (IGR) can be achieved.

    From the study reported in [192,194,197], an increase in the

    amount of water vapour introduced increased the height of

    the CNT arrays, increased the CNT growth rate and improved

    the CNT alignment. However, too much water vapour would

    oxidise the CNTs. Balancing the rate of amorphous carbon

    deposition and the rate of amorphous carbon removal is the

    key factor to producing ultralong CNT arrays [198]. The study

    of Futaba et al. [196] showed that the optimum water/ethyl-

    ene ratio is 1/1000. Our review also shows that the majority

    of millimetre-scale ACNT arrays are produced by introducing

    a small quantity of water vapour into the reactor during CVD.

    However, water vapour also has an etching effect on the cat-

    alyst at high temperatures, as reported elsewhere [199].

    The main contribution of the supergrowth method is to

    provide a solid platform to grow millimetre-scale ACNTs.

    There are two different explanations of how the supergrowth

    happens. The first explanation is that the water vapour in-

    creases the IGR. Futaba et al. [196] derived an equation to cal-

    culate the IGR, and they found that the rate of 207 lm/min

    was obtained in their study. Patole et al. [200] also obtained

    an IGR of more than 200 lm/min. However, Li et al. [187] re-

    ported that water vapour did not increase the IGR. Rather, it

    increased the lifetime of the catalyst. From both phenomena,

    one could speculate that a high IGR is usually found in the

    presence of a relatively high water/hydrocarbon ratio. The

    water acts as an oxidising agent to prevent the wrap up of

    the catalyst nanoparticles by amorphous carbon at the initial

    growth stage, maintaining the catalyst in a fresh condition to

    give very high CNT growth rates. When ACNT arrays form,

    ACNTs keep hydrocarbons and water from reaching the cata-

    lyst and water keeps etching away CNTs. Eventually, both

    rates become equal, and the growth stops. For low water/

    hydrocarbon ratios, the IGR is lower, but the growth can last

    longer.

    Supergrowth is no longer limited to water vapour. Other

    growth enhancers such as alcohols, ethers, esters, ketones,

    aldehydes, and even carbon dioxide have been used [188].

    Air had also been used to enhance the growth rate of CNTs.

    With an optimum amount of air introduced in the reaction,

    the CNT growth duration can be extended to more than

    15 h [201]. Ammonia and hydrogen also possess the ability

    to promote the growth of CNTs, but their efficiency is much

    lower than the oxygen-containing compounds. However, it

    has been reported that water vapour has no significant effect

    on the growth of CNTs with MgO single crystal substrates

    [202].

    The research has recently been focused on obtaining

    aligned SWCNT arrays. Almost all of the reported super-

    growth methods use Fe as the active catalyst and alumina

    as the support layer. The optimal thickness of the layers of

    Fe and alumina during PVD has been widely studied. Super-

    growth cannot be achieved if the thickness of Fe on alumina

    is more than 5 nm [203]. Those thicknesses also determine

    the inter-tube spacing between CNTs, which is an important

    factor for supergrowth to take place. Futaba et al. [204] found

    the ACNTs only occupied 3.6% of the array space when a

    1.2 nm Fe film was used as the catalyst.

    The growth of catalyst-free ACNT arrays obtained by the

    supergrowth method solves the purification problem. The

    extreme high purity of ACNTs enables them to be applied in

    various fields such as biology, chemistry and magnetic appli-

    cations. The controllable production of thick DWCNT arrays

    can be applied to create field-emitting materials that possess

    low threshold voltages. The open tip feature and the sparse-

    ness shown in [205] that was created with water to oxidise

    away the cap make them suitable for gas storage and mem-

    brane applications. Our review shows that Fe is the most

    widely used catalyst for supergrowth. Although Co and Ni

    are also used [206,207], no extensive study or optimisation

    has been carried out. Supergrowth seems to be a promising

    method to achieve mass production of ACNTs.

    4. Substrates and buffer layers

    Substrates provide a solid foundation for growing ACNTs. The

    substrate must be able to inhibit the mobility of the catalyst

    particles in order to prevent agglomeration. The lattice

    matching also decides the morphology of the ACNTs formed.

    Silicon wafers are one of the most popular substrates studied

    C A R B O N 4 9 ( 2 0 1 1 ) 4 6 1 34 6 3 5 4625

  • 8/2/2019 Buck Minster Fuller En Discovery and Consequences

    14/23

    for synthesising ACNTs. However, silicon is not an ideal

    substrate. Jung et al. [38] studied the interaction of Fe on plain

    silicon wafers during CVD, and they found that Fe incorpo-

    rated with silicon to form iron silicide (FeSi2) and iron silicate

    (Fe2SiO4) at high temperatures, which are known for their

    non-catalytic activity for the growth of CNTs (Fig. 13(b) and

    (d)). The same problem affects Ni and Co as well. To overcome

    this problem, a buffer layer, also known as the underlayer or

    adhesion layer, is always used to isolate silicon from active

    catalysts and preserve its activity. Buffer layer also promote

    the dispersion of catalyst nanoparticles and increase the sur-

    face roughness for better adhesion of catalyst nanoparticles

    on the substrate [199]. The TEM images in Fig. 13(a) and (c)

    show that the SiO2 layer blocks the catalyst from interacting

    with silicon. A SiO2 layer has been proven to be an ideal sup-

    port for growing ACNTs because of its high surface roughness

    [208].

    Alumina and Al can be used to prevent extensive sintering

    of Fe particles [209]. By increasing the thickness of the Al

    underlayer, the growth of ACNTs can be promoted [210]. Occa-

    sionally, Al or alumina is used together with SiO2. The pres-

    ence of Al enhanced the morphology of CNTs and increased

    the growth rate, as reported by Liu et al. [49]. It was shown

    that catalyst oxidation on the catalyst surface can be reduced

    by the presence of an Al layer, which helps maintain the cat-

    alytic activity [211]. An alloy of Al and catalyst can be formed,

    which increases the activity of the catalyst [49]. Alumina is

    crucial for the dispersion of smaller NiO crystallites [212].

    The role of SiO2 is to prevent the reaction between Al with

    Si before alumina is formed. However, if the aluminium oxide

    layer is too thick, it will bury the catalyst and inhibit the

    growth of CNTs. TungstenTi bimetals and tantalum are suit-

    able adhesion layers as they limit the diffusion of Ni into the

    substrate [213]. Titanium nitride is also a suitable buffer layer

    [112]. Quartz is another substrate often used to replace sili-

    con. In FC-CVD, a quartz tube can be directly used as the sub-

    strate for growing ACNTs. MgO, sapphire [104], alumina

    mullite mixtures, machinable ceramic [49] and Al2O3 fibre

    cloth [214] are also well-known substrates for growing ACNTs.

    Diameter, density and alignment can be controlled by

    manipulating the structural morphology of the substrate.

    Lee et al. [215] anodised the surface of a silicon wafer to make

    it porous in order to control the density and diameter of the

    catalyst particles. Handuja et al. [43] coated a layer of amor-

    phous hydrogenated silicon nitride (a-SiNx:H) on silicon. It

    was then heated in oxygen to form a crystalline silicon oxide

    (SiOx) within the matrix of a-SiNx, which aided the growth of

    ACNTs in terms of length and alignment. The size of the

    SiNx/SiOx clusters and the orientation of the initial catalytic

    centres determine the alignment and diameter of CNTs [43].

    In the solgel method, mesoporous silica thin films can be

    used. Murakami et al. [128] used tetraethyl-ortho-silicate

    (TEOS), ethanol, H2O, HCl and amphiphilic triblock copolymer

    [(C2H2O)106(C3H4O)70(C2H2O)106], which served as a struc-

    ture-directing agent. The catalyst was loaded after the

    film was formed. Xie et al. [216] used tetraethoxysilane

    ((C2H5O)4Si) hydrolysis in an aqueous solution of iron nitrate

    to create a thin film. ACNTs with a specific diameter and dis-

    tribution were controlled through the preparation conditions

    of the iron/silica substrate [128,216].

    Fig. 13 TEM images of a cross section of the substrates: (a) SiO2; (b) Si after CVD; (c) an enlarged picture from the CNT/SiO2interface in (a) showing the presence of gamma-iron particles on the silicon oxide surface and the growth of CNTs from the

    particles formed; and (d) an enlarged area from (b) showing the formation of FeSi2 and Fe2SiO4crystals during CVD processing

    [38].

    4626 C A R B O N 4 9 ( 2 0 1 1 ) 4 6 1 34 6 3 5

  • 8/2/2019 Buck Minster Fuller En Discovery and Consequences

    15/23

    A flat substrate has a low surface area for the growth of

    CNT arrays. Ceramic spherical substrates have been used to

    mass produce ACNTs [53,217]. Lamellar Fe/Mo/vermiculite

    was successfully applied in fluidised bedCVD for mass

    production [10]. Li et al. [93,94] grew ACNTs successfully on

    1D (Fig. 14) and 2D Si substrates with porous AAO

    nanotemplates.

    5. Alignment of CNTs

    Majorityof thearticles agree with the mechanismof alignment

    elucidated by Fan et al. [8] which is simply caused by the van

    der Waals force. The strong interaction of the van der Waals

    force enables the CNTs bound together to form dense ordered

    packing. During the initial stage of CVD, the lack of van der

    Waals forces results in the formation of entangled CNTs. As

    the catalyst film becomes thicker, the interaction between

    the nanotube walls induces the growth of CNTs with a straight

    form that is parallel to the substrate. The steric impediment

    from neighbouring nanotubesdue to the dense arraypromotes

    the aligned growth [119]. The overcrowding of the CNTs in the

    array forces the tubes to grow only vertically [71]. The work of

    Zhang et al. [218] showed the detailed mechanism of align-

    ment from the beginning to the end of the ACNT growth. The

    van der Waals interactions can be utilised as binding energy

    between adjacent CNTs and the CNT yarns can be drawn for

    various application [26,27].

    Magnetic fields, electric fields and voltage biasing are

    widely applied to provoke the alignment of the CNT array.

    PE-CVD is one of the methods that use an electric field to

    force CNTs to grow parallel to the electric field. The high

    anisotropy of the polarisability of CNTs with an elongated

    shape is responsible for the aligned growth in the electric

    field. The polarisability is stronger for short CNTs [219]. The

    open end of the tubes containing charged dangling bonds

    inhibits the closing of the tube end. The interaction between

    charged tube ends and the electric field contributes to the

    alignment as well [220]. Under the high DC plasma and bias

    voltage, ACNTs are grown by pull up force in the electric field

    of the sheath on a substrate. Another reason that this ap-

    proach works is that the tip of the tubes possess a constant

    orientation of catalytic particles. The polarised catalyst parti-

    cles and the induced dipole moment align the CNTs parallel

    to the electric field all the way through the growth of the CNTs

    [76]. The alignment and the thickness of the array increase

    with the applied voltage. In flame synthesis, further increases

    in the voltage form helically coiled, spiral-like CNTs. When a

    strong electric field is employed, it interacts with the induced

    charges at the tips of the CNTs, which generates sprouting,

    thus splitting catalytic droplets to grow CNTs with L, T

    and Y branches [79]. The polarity of the bias or field and

    the substrate are factors that influence the appearance of

    ACNTs. Otherwise, no alignment could be obtained.

    6. Horizontally aligned carbon nanotubes

    Horizontally aligned CNTs are another form of CNTs in which

    the growth is parallel to the substrate. This orientation is very

    important for applications in the electronics fields. Liu and

    co-workers [62,221,222] first proposed a flow-directed growth

    mechanism for a rapid-heating CVD method. A kite mecha-

    nism was proposed. In rapid-heating CVD, the substrate and

    the surrounding gas are heated at different rates. As a result,

    a convection flow is formed by this temperature difference.

    This flow will lift up the nanotubes with catalyst at the tips.

    A quartz tube reactor with a smaller diameter is preferred

    for CVD, which enhances the laminar flow and thus assists

    the floating of the nanotubes [223]. Cu has been used as a

    replacement for Fe because of its low level of interaction with

    Si, which helps the active tip to float [224]. However, there is

    disagreement about this mechanism. Jin et al. [225] pointed

    out a drawback of Lius method. The hotter gas near to the

    wall of the reactor would levitate along the wall, while in cen-

    tre, the cooler gas at the centre of the tube descend down-

    ward, causing a symmetrical gas circulation as a secondary

    flow. This lateral vortex flow distorted the laminar flow and

    decreased the buoyancy effect that uplifts the growing CNTs,

    altering the alignment. They used an ultralow gas feeding rate

    so that the gas heated up gently to prevent secondary flow.

    However, Li et al. [226] found that horizontally aligned CNTs

    were not parallel to the flow of the gas feed. The study of

    Yu et al. [227] also found that the flow of gas had no effect,

    which means that there are other factors that control the

    alignment of CNTs instead of gas flow.

    The highly anisotropic polarisability of CNTs makes it pos-

    sible to use an electric field as an aligning force during

    growth. Normally, an electric field is applied through a pair

    of electrodes. Dai and co-workers [228] grew highly aligned

    suspended SWCNTs under electric fields in the range of 0.5

    2 V/mm. High voltage would break down the SiO2 dielectric

    layer if Si/SiO2 were used. The short metallic CNTs are more

    readily aligned in an electric field than semiconducting CNTs

    because of their considerably higher polarisability. As a result,

    the angular distribution of short SWCNTs is bimodal. Long

    semiconducting CNTs (>1.5 lm) can also be aligned through

    electric fields [229]. Although electric fields are capable of

    aligning the growth of CNTs, Dai [228,230] and Joselevich

    [229,231] found that electric fields are not the dominant factor

    affecting the growth direction. A magnetic field may also

    assist in the alignment of the CNTs. However, magnetic fields

    will interact with the ferromagnetic catalyst instead of the

    CNTs [232,233]. The CNTs appear to arch over the substrate,

    and it is believed that weak magnetic fields are unable to sup-

    port the weight of the catalyst.

    Nanotubes grown in both a flow-directed and field-direc-

    ted manner have a low density and a low degree of alignment.

    Fig. 14 1D sandwich-structured Si-substrate with an AAO

    template. [94].

    C A R B O N 4 9 ( 2 0 1 1 ) 4 6 1 34 6 3 5 4627

  • 8/2/2019 Buck Minster Fuller En Discovery and Consequences

    16/23

    Introducing strong electric fields is not easy [234]. For this rea-

    son, surface-directed growth was developed to address this

    problem. So far, various mechanisms have been proposed.

    Ismach et al. [235] was the first to report atomic-step template

    growth. This is a method that utilises the miscut of a crystal

    to grow aligned SWCNTs. This technique takes advantage of

    stronger van der Waals interactions caused by the large con-

    tact area at the step edge. SWCNTs have been grown on mis-

    cut C-plane sapphire wafers; they grew along the 0.2 nm high

    atomic steps of the vicinal a-Al2O3 [235]. A wake-growth

    mechanism has been proposed in which the catalyst nano-

    particle slides along the atomic step and leaves the growing

    SWCNTs behind. Aligned SWCNTs were also reported to grow

    on miscut single-crystal quartz substrates [236]. However, the

    authors themselves could not confirm whether the step-edge

    or the lattice directs the growth. Artificial step structures can

    be created to guide the growth direction. The surface geome-

    try can be modified to choose the nanotube orientation.

    Yoshihara et al. [237] grew SWCNTs along trenches they cre-

    ated on Si/SiO2 wafers. However, Maehashi et al. [238] found

    that the SWCNTs grew along the edge of terraces and claimed

    that CasimirPolder interactions direct the growth.

    Lattice-oriented growth is the most popular method that

    has been widely discussed. Su et al. [239] observed that

    SWCNTs only oriented in certain directions on the Si(1 0 0)

    and Si(1 1 1) surface. A base-growth mechanism was pro-

    posed in which the catalysts