You are on page 1of 32

Submitted to The Journal of Physical Chemistry

Carbon Nanotube Growth by Catalytic Chemical Vapour


Deposition: A General Kinetic Model

Journal: The Journal of Physical Chemistry

Manuscript ID: jp-2009-06893m

Manuscript Type: Article

Date Submitted by the


21-Jul-2009
Author:

Complete List of Authors: Monzon, Antonio; University of Zaragoza, Chemical and


Environmental Engineering
Romeo, Eva; University of Zaragoza, Chemical and Environmental
Engineering
Latorre, Nieves; University of Zaragoza, Chemical and
Environmental Engineering
Cazaña, Fernando; University of Zaragoza, Chemical and
Environmental Engineering
Royo, Carlos; University of Zaragoza, Chemical and Environmental
Engineering
Villacampa, Jose; University of Zaragoza, Chemical and
Environmental Engineering
Ubieto, Teresa; University of Zaragoza, Chemical and
Environmental Engineering

ACS Paragon Plus Environment


Page 1 of 31 Submitted to The Journal of Physical Chemistry

1
2
3
4 Carbon Nanotube Growth by Catalytic Chemical Vapour Deposition: A General
5
6 Kinetic Model
7
8 N. Latorre, E Romeo, F. Cazaña, T. Ubieto, C. Royo, J.I. Villacampa and A. Monzón*
9
10 Institute of Nanoscience of Aragon. Department of Chemical and Environmental
11
12 Engineering. University of Zaragoza. 50009 Zaragoza. Spain.
13
14
15
16
Abstract
17 A General Kinetic Model has been developed that includes all the relevant steps
18
19 involved in CNT growth: hydrocarbon decomposition, catalytic nanoparticle
20
21 carburization, carbon diffusion, CNT nucleation and growth, catalyst deactivation and
22
23 self-regeneration, and/or growth termination by the effect of steric hindrance.
24 Here we emphasize the importance of a suitable kinetic description of all the
25
26 stages, in particular the initial carburization-nucleation and the growth cessation. We
27
28 have discussed the different mechanisms proposed to explain the critical step of CNT
29
30 nucleation and have used an autocatalytic kinetic equation to describe it. The two
31
32
parameters involved in this autocatalytic equation allow a very good fit of the initial
33 induction period usually observed during the growth of CNTs. In addition, rigorous
34
35 formulations of the main causes of CNT growth cessation (catalyst deactivation and
36
37 steric hindrance) have been proposed.
38
39 The General Model as developed is shown to be a potentially versatile tool of
40 general application. In this paper we have applied it to fit data obtained in our lab, and
41
42 also to super growth of VA-SWNT experimental data published in the literature. In all
43
44 cases the values obtained for the kinetic parameters have realistic physical meaning in
45
46 good agreement with the mechanism of CNT formation.
47
48
49
50
51
52
53 Corresponding author: Prof. A. Monzon,
54 e-mail: amonzon@unizar.es, Tel: +34 976 761157; Fax.: +34 976 762142
55
56
57
58
59
60

1
ACS Paragon Plus Environment
Submitted to The Journal of Physical Chemistry Page 2 of 31

1
2
3 1. Introduction.
4
5
6 The attractive possibilities of using the outstanding properties1,2 of carbon
7 nanotubes and other new carbonaceous nanomaterials in a broad range of new
8
9 applications3-5 is motivating substantial research effort in practically all the fields of
10
11 nanoscience and nanotechnology. However, extensive use of these materials requires
12
13 the development of scalable and selective production processes. Catalytic chemical
14
15
vapour deposition (CCVD) has become probably the main technique for the synthesis of
16 carbon nanotubes, including selective production of single-walled nanotubes (SWNTs)
17
18 using many carbon sources and catalysts6,7. In addition to the usual CCVD method, that
19
20 uses large surface area porous supported metallic catalysts8-10, the production of layers
21
22 of vertically aligned VA-SWNT is assuming increasing importance11,12. VA-SWNTs
23 can also be produced by CCVD if the catalyst composition and the operating parameters
24
25 are optimized for the synthesis conditions13,14. Furthermore, this technique can be
26
27 improved, for example combining a dip-coat catalyst loading process15 with the alcohol
28
29 catalytic chemical vapour deposition (ACCVD) method12,16-18. Other methods used are
30
31
water-assisted CVD13,19, oxygen assisted CVD20, point-arc microwave plasma CVD21,
32 molecular-beam synthesis22, and hot-filament CVD23. These new advances have
33
34 succeeded in significantly increasing the overall yield, but improvements in SWNT
35
36 quality and control over chirality are still necessary, particularly when considering
37
38 electrical and optical applications18.
39
40 Even though the formation and growth mechanisms of carbon nanotubes by
41
42 CCVD have been extensively studied in the past24-35, there is no general agreement
43 about what the critical steps are. Most authors propose that this mechanism includes the
44
45 stages of hydrocarbon (or another carbon source such as CO) decomposition over the
46
47 metal surface, carbon diffusion through the particles27,31-35 and/or atomic carbon surface
48
49 transport36,37 and finally carbon precipitation forming CNTs. Although this form of
50
51
carbon accumulation allows the catalyst to maintain its activity for an extended period
52 of time, catalyst deactivation can occur through the formation of encapsulating carbon
53
54 over the surface of the metal particles34,36. The deactivation phenomenon can be
55
56 reversible as a consequence of gasification of this type of carbon by oxygen19, water38,39
57
58 or hydrogen40-43 that can be added, or be present, in the feed. Additionally to catalyst
59 deactivation, other causes of carbon growth cessation such as steric hindrance44 or
60
defect diffusion to the growth front45 have been considered. Other phenomena such as

2
ACS Paragon Plus Environment
Page 3 of 31 Submitted to The Journal of Physical Chemistry

1
2
3 initial catalyst activation can also occur if the catalyst is not previously reduced before
4
5 the reaction44,46.
6
7 The physical-chemical description of the steps involved in the CNT growth
8
9 process has usually been tackled by kinetic models that take into account only some of
10
11 the stages (e.g. nucleation and initial growth,47). Other proposed models describe the
12
13 process in steady state31,33,48, without considering that it is strongly time-dependent.
14
15
These models are usually used to fit experimental data of: i) stationary carbon
16 formation rate as a function of the operating conditions40,41,33,48; ii) evolution over time
17
18 of the carbon formation rate49,50,51; and iii) evolution over time of mass (or length of
19
20 CNTs) of carbon accumulated15-18,42-44,52.
21
22 Obviously, a rigorous description of all these stages necessarily implies obtaining
23
24 very complex mathematical models with too many parameters, which eventually
25
26 hinders their application for the analysis of kinetic data. On the other hand, the simplest
27
models are frequently used owing to the fact that they are easy to apply and understand,
28
29 although obviously these models are unable to describe the complete process accurately.
30
31 A compromise solution is to find models that consider the main critical stages of the
32
33 process without excessively increasing its mathematical complexity and therefore the
34
35
number of parameters.
36
37 In any case, it is necessary to have a sufficient quantity of precise experimental
38
39 data, for which in situ techniques are especially suitable. Among the in situ techniques,
40 the most commonly used to follow the growth of carbon nanotubes are: i)
41
42 environmental HR-TEM35-37,52-55; ii) optical absorbance15-18,45; iii) RAMAN scattering56;
43
44 iv) time-resolved reflectivity (TRR) of laser beams52,57; v) gas analysis by mass
45
46 spectrometry of time-evolved gases44,58, vi) thermogravimetric techniques42,43,51,59,60. In
47
addition, SEM images taken after different reaction times have also been used to follow
48
49 the kinetics of VA-SWCNT growth61,62.
50
51
52
Recently, our group has developed kinetic models to investigate the growth of
53 carbon nanotubes. These models were successfully applied to studying the data obtained
54
55 with in situ thermogravimetric systems in reactions with different carbon sources and
56
57 catalysts42-44,59-63. In the different examples studied, we have considered the steps of
58
59 deactivation-regeneration43,60, catalyst activation, or growth termination by steric
60 hindrance44.

3
ACS Paragon Plus Environment
Submitted to The Journal of Physical Chemistry Page 4 of 31

1
2
3 In spite of these efforts, an adequate description of the critical initial induction
4
5 period of time has not been attained. In fact, the evolution of the rest of the steps is
6
7 determined by this initial period, when some surface carburization of the metallic
8
9 nanoparticles and carbon nanotube nucleation takes place. It has been observed that the
10
length of this period is strongly dependent on the operating conditions42,56. Thus, it has
11
12 been found that the induction period becomes longer as the partial pressure of the
13
14 carbon source diminishes42,56,58,60,63, indicating that the initial carburization-nucleation
15
16 step is controlled by the feeding of carbon atoms. With the aim of obtaining a more
17
18
complete description of the growth process, in this work we present a kinetic model
19 including all the above considerations, considering especially the initial carburization-
20
21 nucleation step and the growth termination, by catalyst deactivation and/or by the effect
22
23 of steric hindrance. However, we have tried to reduce the number of empirical equations
24
25 using parameters with a physical-chemical meaning. In order to show that the developed
26
model could be a versatile tool for general application, we have used it not only to fit
27
28 data from our lab, but also other data as for example super growth of VA-SWNT
29
30 experimental data.
31
32
33
34
35 2. Kinetic model of CNT growth.
36
37 A) Hydrocarbon (carbon source) decomposition over the catalytic surface.
38
39 As an example, here we take the case of catalytic methane decomposition; one of
40
41 the most commonly used CCVD process. Other carbon sources can also be considered
42
43 in the same way. The mechanism of methane decomposition can be expressed through
44
the following individual steps32,33:
45
46
47 CH 4 + l ⇔ CH 4 − l (i)
48
49
CH 4 − l + l →
rds
CH 3 − l + H − l (ii)
50
51
52 CH 3 − l + l ⇔ CH 2 − l + H − l (iii)
53
54
55 CH 2 − l + l ⇔ CH − l + H − l (iv)
56
57 CH − l + l ⇔ C − l + H − l (v)
58
59
60 2 H − l ⇔ H 2 + 2l (vi)

4
ACS Paragon Plus Environment
Page 5 of 31 Submitted to The Journal of Physical Chemistry

1
2
3 Atoms of carbon and gaseous hydrogen are then released over the catalyst surface.
4
5 In this case it is assumed that the number of active sites “l” involved in the controlling
6
7 step (named as m) of the methane decomposition is 2 (m=2)32,33. The step of reversible
8
9 formation of encapsulating coke, which deactivates the catalyst, goes through parallel
10
reactions of condensation-oligomerization, and can be expressed as64:
11
12
13 h(CH x − l ) ⇔ Plh ↓ + h 2 H 2 (vii)
14
15
16
Plh is the encapsulating coke and h is the number of active sites involved in the
17 above reaction of coke formation64.
18
19
20
B) Formation of surface carbide and carbon nanotube nucleation.
21
22 According to the mechanism proposed by Alstrup29, after hydrocarbon decomposition
23
24 the remaining carbon atoms react with the metallic nanoparticles at the surface forming
25 a metastable carbide, that in the reaction conditions decomposes leaving carbon atoms
26
27 at the metallic subsurface. After this decomposition-segregation step, the carbon atoms
28
29 are introduced inside the metal particles29, determining in this way the value of the
30
31 carbon concentration at the carbide-metallic nanoparticle interphase. However, for this
32
stage Puretzki et al.52 consider that the carbon atoms are dissolved on the metallic
33
34 nanoparticles forming a highly disordered ‘molten’ layer on their surface. Because of
35
36 much higher carbon diffusivity in the disordered layer compared to the ordered solid
37
38 phase, the carbon atoms diffuse along this layer and precipitate into a nanotube. Helveg
39
40 et al,36, based on density-functional theory calculations and on “in situ” TEM
41 observations, propose that the growth of CNFs on a Ni-Mg-Al catalyst involves surface
42
43 diffusion of both carbon and nickel atoms. Thus, the graphene layer nucleation and
44
45 growth is explained as a dynamic formation and restructuring of mono-atomic step
46
47 edges at the nickel surface.
48
49
50 In the case of SWNT growth, in a recent paper44 we have considered that the carbon
51
52 atoms enter into the metallic nanoparticles through the clean surface (interface 1), and
53
54 that they leave the metallic phase through interface 2 when forming the SWNT. The
55
56
driving force for the surface or bulk diffusion from interface 1 to interface 2 is the
57 difference in chemical potential between the two interfaces. Comparing these two
58
59 diffusion pathways, Lin et al.37 consider that the surface diffusion is dominant due to its
60
lower activation barrier, as a consequence of a lower coordination number. After the

5
ACS Paragon Plus Environment
Submitted to The Journal of Physical Chemistry Page 6 of 31

1
2
3 carbon concentration has reached a certain threshold, nucleation of ordered forms of
4
5 carbon (e.g., hexagons) occur forming nuclei of graphene caps38,65,66. The development
6
7 of the graphene caps is driven by the need to minimize the energy associated with the
8
9 nucleation of a graphene layer when constrained to grow and form a nanotube37.
10
In any case, the nucleation stage can be expressed as:
11
12
13 n(C − l ) → nCs + nl (viii)
14
15 where CS is the concentration of surface carbide and has units of (g.C/g. cat). Supposing
16
17 that the process of nucleation follows an autocatalytic kinetics44,67 the rate of
18
19 carburization-nucleation can be expressed as:
20
21 dCS
22 rS = =ψS ⋅ (1+ KS ⋅ CS ) ⋅ (CSm −CS ) (1)
23 dt
24
25 In this equation, the autocatalytic contribution can be related to an autoassembled
26
27 process67 that, for example, occurs during the formation of the caps of the SWNT38,65,66.
28
29 The term ψS represents the intrinsic kinetic function of carburization, and for a given
30
catalyst its value depends on the reaction conditions. CSm represents the maximum
31
32 surface carbide concentration attainable on the surface of metallic particles at the side
33
34 gas phase (interface 1), and this value determines the thickness of the carbide layer.
35
36 This thickness has been associated52 to the number of layers of the MWNT precipitated
37
38
from the metallic particles. On the other hand, the value of the parameter KS determines
39 the weight of the autocatalytic effect on the carburization kinetics.
40
41
42 Assuming that the reactor operates under differential conditions, i.e. ψS and KS are
43
44
constants at all the reactor positions, the integration of the above expression gives the
45 evolution of CS over time:
46
47
(1 − exp(−α ⋅ t ))
48 CS = CSm ⋅ ; α = ψ S ⋅ (1 + K S ⋅ CS ) (2)
49 (1 + K S ⋅ exp(−α ⋅ t )) m

50
51
52 Therefore, the rate of surface carburization can also now be expressed as:
53
54 dCS CS m ⋅ α ⋅ (1 + K S ) ⋅ exp(−α ⋅ t ))
55 rS = = (3)
56
dt (1 + K S ⋅ exp(−α ⋅ t )) 2
57
58 From the above equation it is deduced that rS passes through a maximum time
59
60 equal to:

6
ACS Paragon Plus Environment
Page 7 of 31 Submitted to The Journal of Physical Chemistry

1
2
3
t S = ln( K S ) α (4)
4
5
6 Taking the value of tS, equation 2 can be rewritten as:
7
8 (1 − exp(− t τ S )) 1
9 CS = C S m ⋅ ; τS = (5)
10 (1 + exp(− (t − t S ) τ S )) α
11
12 The term τS can be considered as the lifetime of the carburization process.
13
14
15 If the value of KS is zero, or its value is very low compared to ψS, the autocatalytic
16
17 contribution is negligible, and the kinetics of carburization-nucleation is described by a
18 first order law:
19
20
dC S
21 = ψ S ⋅ (CS m − CS ) ⇔ C S = C Sm (1 − exp(−ψ S ⋅ t )) (6)
22 dt
23
24
In this case, rS does not pass through a maximum, decreasing monotonically over
25
26 time. Finally, if the carburization-nucleation of the metallic nanoparticles occurs very
27
28 fast, the value of ψS will be very high, and CS will attain the value of CSm almost
29
30 instantaneously, i.e. CS=CSm.
31
32 B) Rate of carbon nanotube growth.
33
34
This stage begins with the formation-precipitation of the carbon nanotubes. As
35
36 was mentioned above, when carbon concentration has reached a certain threshold,
37
38 nucleation occurs forming graphene caps38,65,66. This fact generates the interface CNT-
39
40 metal, interface 2, and the carbon flux is maintained because the nanotube structure
41
42
provides a thermodynamic sink for the carbon, and as a result the carbon concentration
43 at the interface 2 is kept low. The rate of the diffusion-precipitation process determines
44
45 the rate of formation of carbon nanotubes. In the case of methane decomposition the
46
47 stoichiometry is CH4(g)↔C(s)+2H2(g), and therefore the rates of methane conversion,
48
49 CNT formation, and hydrogen production are directly related by:
50
51 (−rCH 4 ) t = (rC ) t = (rH 2 ) t 2 (7)
52
53
54 In addition, the mass balance for the reactant (methane in this case) in a plug-flow
55
reactor, assuming pseudo-steady state, can be expressed as:
56
57
58 dX CH4
59 − (rCH4 )t = 0 (8)
d (W ( FCH4 )0 )
60

7
ACS Paragon Plus Environment
Submitted to The Journal of Physical Chemistry Page 8 of 31

1
2
3 In the above expression XCH4 represents the methane conversion, (FCH4)0 the mass
4
5 flowrate of methane fed to the reactor and W the catalyst weight. The solution of this
6
7 differential equation allows the calculation of methane conversion, hydrogen production
8
9 and of the total amount of carbon nanotubes produced along the reactor and over time.
10
Bearing in mind the deactivation produced the formation of encapsulating coke,
11
12 equation 8 can be expressed in terms of catalyst activity, a,64,68:
13
14
15 dX CH 4
(−rCH 4 )t = = (rC )t = (rC ) 0 ⋅ a ; a = (rC )t (rC ) 0 (9)
16 d (W ( FCH 4 ) 0 )
17
18
19 The term (rC)0 corresponds to the CNT growth rate, in the absence of deactivation,
20
21 which can be expressed in terms of difference of chemical potential force between the
22
interfaces 1 and 225,26,32,33,44:
23
24
25 dmC
26
(rC ) 0 = = kC ⋅ (CS − C F ) (10)
dt t =0
27
28
29 The term k C is the effective coefficient of carbon transport, has units of time-1,
30
31 and depends on the average size of the metallic crystallites, the metallic exposed area,
32
33 and the carbon atom diffusivity on the metallic nanoparticles25,26,32,33,44:
34
35 C) Catalyst deactivation
36
37 The main causes of catalyst deactivation are fouling by encapsulating coke,
38
39 sintering (or thermal aging), and poisoning68. Therefore, the deactivation rate,
40
41 rd = − da dt , must be described according to an appropriate deactivation kinetic model
42
43 related to the deactivation cause68-71. Thus, if the formation of encapsulating coke is
44
45 partially reversible, the catalyst does not suffer a complete deactivation, maintaining a
46
47
residual level of activity at steady state. In these conditions, the net rate of activity
48 variation can be expressed as71:
49
50
da
51 rd = − = ψ d ⋅ a d −ψ r ⋅ ( a d m − a ) (11)
52 dt
53
54 The terms ψ d and ψ r are respectively the “deactivation and regeneration kinetic
55
56
functions”, and both also depend on the operating conditions. If the activity decay is
57
58 irreversible, ψr=0, the deactivation rate will be64,68:
59
60
da
rd = − =ψ d ⋅ ad (12)
dt

8
ACS Paragon Plus Environment
Page 9 of 31 Submitted to The Journal of Physical Chemistry

1
2
3 It should be noted that, in equation 11 or 12, the values of the kinetic orders d and
4
5 dm depend on the reaction mechanism, i.e. the number of sites involved in the
6
7 reactions71:
8
9 d = (m + h − 1) m ; d m = (m − 1) m (13)
10
11
12 Thus, only values with a real physical meaning for m and h, i.e. 1 or 2, were
13
14 assessed. In fact, cases involving 3 or more active sites in an elemental step were not
15 considered since they are quite improbable72. This means that m and h are not fitting
16
17 parameters, and therefore their values must be selected before the data fitting.
18
19
D) Steric hindrance.
20
21
22 For the SWNT it has also been considered that the growth on porous catalysts can be
23
24
slowed down as a consequence of the steric hindrance44. In this case, as carbon
25 accumulates within the internal porous structure, physical hindrance for the free
26
27 enlargement of the growing nanotubes may occur. This fact constrains SWNT growth,
28
29 delaying the insertion of new carbon atoms at interface metal-SWNT. The degree of
30
31 interaction of a growing SWNT with the catalyst support, and/or with other nanotubes,
32 depends on the catalyst pore sizes. The larger the available space on the catalyst, the
33
34 less hindered will be the insertion of new carbon atoms at the interface. Similar
35
36 considerations have been made to explain the deactivation of catalytic carbonaceous
37
38 materials used to produce hydrogen by methane decomposition73,74.
39
40
We assume here that this hindrance effect can be expressed as a potential function of the
41 accumulated mass of carbon:
42
43
44 C F = ξ H ⋅ mCp (14)
45
46 In this equation, the term ξH is called the “hindrance factor”. In addition, the
47
48 value of the hindrance kinetic order, p, determines if this effect is progressive over time,
49
50 p>1, or decays over time, p<1. As commented above, the values of both parameters
51
52 depend on the catalyst texture.
53
54 Finally, substituting equations 2, 10 and 14 into equation 9, the amount of CNTs
55
56 formed can be calculated integrating the following expression:
57
58 dmC jC (1 − exp(−α ⋅ t ))
59 + ξ HC ⋅ mCp ⋅ a = 0 ⋅a (15)
60 dt (1 + K S ⋅ exp(−α ⋅ t ))

9
ACS Paragon Plus Environment
Submitted to The Journal of Physical Chemistry Page 10 of 31

1
2
3
Where the terms ξHC and jC0 are defined as:
4
5
6 ξ HC = ξ H ⋅ kC ; jC0 = C S m ⋅ kC (16)
7
8
9
jC0 has units of (g C/g cat.min) and can be considered as the intrinsic CNT growth
10 rate for the fresh catalyst. Similarly, the term ξHC represents the intrinsic constant rate of
11
12 steric hindrance, and has units of (g cat(1-p)/g C(1-p).min). The differential equations 11
13
14 and 15 can be considered as a general kinetic model for studying the growth rate and the
15
16 amount of CNTs formed over the reaction time, as a function of the operating
17
18
conditions, and of the potential effects of catalyst deactivation and/or steric hindrance.
19 The parameters involved are: ξHC and p, accounting for the steric hindrance, ψd and ψr
20
21 relating to catalyst deactivation and jC0, that is the intrinsic growth rate. This general
22
23 model can be simplified into many particular cases; some of them are described in the
24
25 following paragraphs.
26
27 On the other hand, one of the most original aspects of this model is the
28
29 description, by means of parameters ψS and KS in equations 1 and 2, of the
30
31 carburization-nucleation stage. Figures 1 to 3 show some examples of the effect of these
32 parameters on the CNT growth. In Figures 1 and 2 it can be seen that as the value of KS
33
34 increases, the induction period becomes longer, indicating that the more demanding the
35
36 autocatalytic effect, the more time is needed to achieve the initial stage of nucleation. In
37
38 contrast, if KS is zero or takes a low value, the induction period is not appreciable
39
40
because the carburization-nucleation steps occur very quickly.
41
42 As regards the catalyst deactivation, this causes a diminution in the reaction rate
43
44 and then on the mass of CNTs accumulated over the catalyst. Furthermore, the
45 maximum rate of reaction appears later, and with a lower value (Figure 1). In relation to
46
47 the evolution of mC, an increase in KS produces a decrease in the amount of CNTs
48
49 formed because the catalyst deactivation occurs before the carburization-nucleation step
50
51 has been completed and the nanoparticles can transport the carbon atoms. On the other
52
hand, if there is no catalyst deactivation, after the initial induction period the reaction
53
54 rate increases until it reaches a constant final value (Figure 2). Consequently, the
55
56 induction period mC subsequently increases linearly over the reaction time (Figure 2). In
57
58 Figure 3 the results relating to the impact of the parameter ψS are presented. Thus, an
59
60 increase in ψS causes a shortening of the induction period and simultaneously an
increase in the maximum reaction rate. Both phenomena are the result of the fact that

10
ACS Paragon Plus Environment
Page 11 of 31 Submitted to The Journal of Physical Chemistry

1
2
3 the termination of the carburization-nucleation stage occurs before the deactivation of
4
5 the catalyst is advanced. The results in Figures 1 to 3 correspond to the case where there
6
7 is no effect of steric hindrance (ξHC=0). However, qualitatively similar results will be
8
9 obtained if the growth termination is caused by steric hindrance, or by both causes
10
11 acting simultaneously.
12
13 In the following sections we present the equations corresponding to some
14
15 particular cases of the general model developed here.
16
17 Case 1. CNT growth termination by catalyst deactivation alone.
18
19 If the CNT growth is not hindered by any steric impediment, ξHC=0, or the value
20
21 of CF is very low compared with CS, equation 15 can be simplified to:
22
23
dmC jC (1 − exp(−α ⋅ t ))
24 rC = = 0 ⋅a (17)
25 dt (1 + K S ⋅ exp(−α ⋅ t ))
26
27
28 The value of the catalyst activity, a, is calculated from equation 11, and then the
29
fitting parameters in this case are ψd, ψr, jC0, KS and ψS. The explicit equation for the
30
31 activity vs. time relationship depends on the values of m and h. For example, in the
32
33 simplest case, m=h=1, the deactivation rate and the catalyst activity are given by the
34
35 following expressions:
36
37 da
38 rd = − = ψ d ⋅ a −ψ r ⋅ (1 − a ) (18a)
39
dt
40
41 a = aS + (1 − aS ) exp(−(ψ d +ψ r ) ⋅ t ) ; aS = ψ r (ψ d +ψ r ) (18b)
42
43
44
The term aS represents the residual activity of the catalyst.
45
46 Now, mC must be calculated integrating numerically the following equation:
47
48 dmC jC (1 − exp(−α ⋅ t ))
49 = 0 ⋅ (aS + (1 − aS ) exp(−(ψ d +ψ r ) ⋅ t ) ) (19)
50 dt (1 + K S ⋅ exp(−α ⋅ t ))
51
52 If the deactivation is irreversible, ψr=0, and the carburization-nucleation step is
53
54 very rapid, i.e. KS 0 and ψS is very high, then CS=CSm, In these conditions equation 19
55
56 results in the most simple case:
57
58 dmC
59 = jC0 ⋅ exp(−ψ d ⋅ t ) (20a)
60 dt

After integration, it is obtained that:

11
ACS Paragon Plus Environment
Submitted to The Journal of Physical Chemistry Page 12 of 31

1
2
3
jC0
4 mC (t ) = (1 − exp(−ψ d ⋅ t ) ) (20b)
5 ψd
6
7
8
Equation 20b is formally equal to that used to describe many studies of the
9 production of layers of vertically aligned single-walled nanotubes, VA-SWNT,
10
11 including the so-called super-growth proces17-19, 45, 61:
12
13
  t 
14 H (t ) = β ⋅ τ 0 1 − exp −   (21)
15
16
  τ0 
17
18 The equivalence between the kinetic parameters of both models is given by:
19
20 H (t ) ∝ mC (t ) ; β ∝ jC ; τ0 = 1ψ d
(22)
0
21
22 H max = β ⋅τ 0 ∝ jC0 ψ d
23
24
25 In summary, equation 21 can be considered as a particular case of the general
26 model described in this work. It is obvious that for the cases described by equations 15
27
28 and 20, there are many others that can be easily deduced and applied for each specific
29
30 particular study. For example, if the carburization-nucleation step is described by
31
32 equation 6, instead of considering that CS=CSm, the following intermediate case is
33
obtained:
34
35
36  (1 − exp(−(ψ d + ψ S ) ⋅ t )) (1 − exp(−ψ S ⋅ t )) 
37 mC = jC 0  −  (23a)
38  ψ d +ψ S ψS 
39
40 rC = jC 0 (1 − exp(−ψ S ⋅ t )) ⋅ exp(−ψ d ⋅ t ) (23b)
41
42
43 These equations will be applied later in the analysis of VA-SWNT supergrowth
44
45 data.
46
47 Case 2. CNT growth cessation by steric hindrance only.
48
49 If the catalyst does not suffer deactivation, i.e. a=1, and the growth cessation is
50
51 only caused by steric hindrance, rC is calculated as:
52
53
dmC jC (1 − exp(−α ⋅ t ))
54 + ξ HC ⋅ mCp = 0 (24)
55 dt (1 + K S ⋅ exp(−α ⋅ t ))
56
57
58
In the above equation the fitting parameters are ξHC, p, jC0, KS and ψS. Similarly to
59 the above case, if p=1, the carburization-nucleation step occurs very quickly, i.e. KS
60
0, ψS is very high and CS=CSm, equation 24 is simplified to:

12
ACS Paragon Plus Environment
Page 13 of 31 Submitted to The Journal of Physical Chemistry

1
2
3 dmC
4 + ξ HC ⋅ mC = jC0 (25a)
5 dt
6
7 The analytical solution of the above equation is:
8
9
jC0
10 mC (t ) = (1 − exp(−ξ FC t )) (25b)
11 ξFC
12
13
14 The above equations are identical to equations 20a and 20b, replacing ψd by ξHC.
15
16 This result also indicates that by means of kinetic studies alone it is not easy to
17 distinguish what is the true underlying reason, or reasons, for CNT growth decay. In
18
19 fact, both causes considered here can act simultaneously, clearly complicating the
20
21 elucidation of the parameters and requiring a previous design of a careful experimental
22
23 study specifically designed to obtain them.
24
25 Case 3. No effect of steric hindrance or catalyst deactivation.
26
27 This case corresponds to the simplest situation where the growth rate is maintained
28
29 along the reaction38,47. In this situation equation 15 is simplified to:
30
31
 1 − exp(−α ⋅ t ) 
32 (rC )t = jC0 ⋅   (26a)
33  1 + K S exp(−α ⋅ t ) 
34
35
36 The analytical integration of the above expression gives:
37
38  1+ KS   (1 + K S exp(−α ⋅ t ) ) 
39 mC (t ) = jC0 ⋅ t −   ⋅ ln  (26b)
40   K Sα   1+ KS 
41
42
43
This solution is represented in the curves corresponding to the evolution of mC in
44 Figure 2.
45
46
47
48
49 3. Application of the kinetic model to experimental data.
50
51 In Figures 4 to 7 we present different results of the application of the general
52
53 kinetic model developed in this contribution and some comparisons with other simpler
54
55 models used in the literature. Thus, Figures 4a and 4b show two examples taken from a
56
kinetic study of vertically aligned single wall carbon nanotube (VA-SWCNT) growth,
57
58 synthesized by catalytic chemical vapour deposition of ethanol18. In that paper, the
59
60 authors study the effect of the partial pressure of ethanol and of the synthesis
temperature over the evolution over time of the VA-SWCNT thickness. The kinetic

13
ACS Paragon Plus Environment
Submitted to The Journal of Physical Chemistry Page 14 of 31

1
2
3 analysis of the results obtained was made using a kinetic model that corresponds to
4
5 equation 21, here called the Exponential Model. As has been discussed before, the
6
7 Exponential Model is a particular case of the General Model described by equation 15.
8
9 The implicit simplifications contained in equation 21 do not allow, for example, an
10
explanation of the presence of a maximum in the growth rate, as is shown in Figures 4a
11
12 and 4b. However, this maximum is usually observed in this type of experiment18,42-
13
44,52,58
14 . Furthermore, as has also been discussed before, another important phenomenon
15
16 usually observed is the presence of an initial induction period that the Exponential
17
18
Model is unable to predict.
19
20 Table 1 shows the values of the kinetic parameters obtained with both models. In order
21
22 to achieve more statistical consistency, the parameter estimations have been made
23 fitting simultaneously both (1) the VA-SWCNT thickness data and G-band (i.e. the
24
25 measurement proportional to the amount of CNT accumulated on the catalyst in each
26
27 case); and (2) the growth rate data. Thus, the objective functions minimized are defined
28
29 as follows:
30
31 SSRTotal = SSR1 + SSR2 (27a)
32
33
SSR = ∑
34 (y exp − ycalc ) 2
35 (27b)
yexp
36
37
38 In the case of SSR1 the values of the variable y are the VA-SWCNT thickness or the G-
39
40 band intensity, while for SSR2 y corresponds to the growth rate data.
41
42 As can be seen in Figures 4a and 4b, the fittings obtained with the General Model are
43
44 more realistic than those obtained with the Exponential Model because the former is
45
46 able to explain the presence of a maximum value of the reaction rate. According to the
47 General Model, initially the growth rate is zero and, as the nucleation step progresses,
48
49 the reaction rate increases quickly reaching a maximum. Then the rate decreases as a
50
51 consequence of the catalyst deactivation. The different behaviour observed in figures 4a
52
53 and 4b is a consequence of the different operating conditions used in both cases18. Thus
54
55
in the case of figure 4b the carburization step is very fast, ψS = 890, and the maximum
56 rate appears very soon, in accordance with the tendency presented in Figure 3. In this
57
58 case, equation 18 can be simplified to equation 21 and as a consequence the values of
59
60 the kinetic parameters obtained with both models are very similar (see Table 1). The
number of parameters used with the General Model, in this case 3, is higher that with

14
ACS Paragon Plus Environment
Page 15 of 31 Submitted to The Journal of Physical Chemistry

1
2
3
the Exponential Model, 2. However, the inclusion of this additional parameter, ψS, is
4
5 totally justified from a physical point of view.
6
7
8
The data in Figure 5 corresponds to a study of SWNT synthesis by CCVD
9 followed by in situ RAMAN spectroscopy 56 also carried out by Maruyama’s group. As
10
11 can be seen, the evolution of the G-band signal intensity, proportional to the amount of
12
13 SWNT formed, shows a very long induction period due to the very low value of the
14
15 partial pressure of ethanol (0.1 Torr) used during the experiment. In these operating
16 conditions the rate of carburization of the metallic particles is low, and therefore the
17
18 time needed to complete the nucleation is greater causing an enlargement of the
19
20 induction period. Obviously, from figure 5 and the values presented in Table 1 it is clear
21
22 that the fitting obtained with the Exponential Model is clearly worse than that obtained
23
24
with the General Model. With respect to the values of the parameters, the fitting of the
25 long induction period is solved by the General Model giving a low value for the ψS
26
27 parameter and a high value for the KS parameter, in agreement with the trends presented
28
29 in Figures 1 and 3. Figure 5 also shows the evolution of the growth rate of the SWNT
30
31 calculated with the values estimated using the General Model. This curve shows clearly
32
33
that in the first instants of the reaction (in this case the first minute) the reaction rate is
34 almost zero, just the contrary of the result obtained with the Exponential Model that
35
36 initially predicts the higher reaction rate.
37
38 Finally, the results presented in Figures 6a and 6b correspond to a study of CCVD of
39
40 CH4 on a Ni-Al catalyst, carried out by our group42. These figures correspond to the
41
42 results obtained analyzing the influence of the methane concentration on the CNT
43
44 production and on the reaction rate. Figure 6a shows that the length of the induction
45 period is strongly dependent on the carbon source concentration, increasing as the
46
47 partial pressure of CH4 diminishes. This result confirms the conclusions obtained from
48
49 Figure 5. Table 2 shows the values of the parameters of the General Model, after
50
51 simultaneous fitting of the experimental data in Figures 6a and 6b. In this case, the
52
53
values of the variable y on equation 27a are carbon concentration, while for SSR2 y
54 corresponds to the growth rate data (equation 27b).
55
56
57 The evolution of the parameters with respect to the methane concentration
58
indicates that as the CH4 partial pressure increases, the values of jC0 and ψd also
59
60
increase. These results explain that at the end of the experiment the total amount of

15
ACS Paragon Plus Environment
Submitted to The Journal of Physical Chemistry Page 16 of 31

1
2
3 CNT formed is similar because the higher reaction rate is compensated by a higher
4
5 catalyst deactivation58.
6
7
8
As regards the parameters related to the carburization step, the value of ψS also
9 increases with the CH4 concentration and the parameter KS remains practically constant.
10
11 This result indicates that the model can explain the variation of the length of the
12
13 induction period only modifying the parameter ψS.
14
15 The values in Table 2 can be used to estimate the kinetic orders of these
16
17 parameters with respect to CH4. Thus, the orders for jC0 , ψd and ψS are respectively 0.8,
18
19 0.9 and 1.3. The kinetic orders for jC0 and ψd indicate that both stages, the CNT growth
20
21
22 and the deactivation, involve an adsorption step of the methane molecules on the
23 metallic particles surface, according to the mechanism proposed above. In addition, the
24
25 value of 1.3 for ψS can be explained assuming a second order process according to an
26
27 autocatalytic phenomenon considered for the carburization step.
28
29 In summary, for a given catalyst composition and carbon source, the developed
30
31 model allows the selection of the operating conditions that maximize the CNT
32
33 production avoiding, or minimizing, catalyst deactivation and the formation of
34
35 undesirable materials.
36
37
38
39 4. Conclusions.
40
41
42
We have developed a General Kinetic Model that includes all the relevant steps
43 involved in CNT growth. As an advance on previous models, in the present paper we
44
45 have stressed the importance of an adequate description of initial carburization-
46
47 nucleation and growth termination steps.
48
49 The extent of the initial induction period observed during the growth of CNTs can be
50
51 modulated modifying the operational conditions, especially the concentration of the
52
53 carbon source. The length of this period can be quantified through the values of
54
55
parameters ψS and KS.
56
57 The results presented in Figures 4 to 7 and Tables 1 and 2 indicate that the General
58
59 Model is a useful instrument that enables the importance of each individual step to be
60 discriminated: carburization, nucleation, growth and deactivation or steric hindrance. In

16
ACS Paragon Plus Environment
Page 17 of 31 Submitted to The Journal of Physical Chemistry

1
2
3 all cases the values obtained of the kinetic parameters have realistic physical meaning in
4
5 good agreement with the mechanism of CNT formation.
6
7
8
9
10 Acknowledgements.
11
12 The authors acknowledge financial support from MICINN (Spain)-FEDER, Project
13
14 CTQ 2007-62545/PPQ, and the Regional Government of Aragón, Departamento de
15
16
17 Ciencia, Tecnología y Universidad, Project CTP P02/08. Also, the authors thank Prof.
18
19 S. Maruyama for sending the experimental data shown in Figures 4 and 5.
20
21
22
23 References.
24
25 (1) Saito, R.; Dresselhaus, G.; Dresselhaus, M. S. Physical Properties of Carbon
26
27 Nanotubes, Imperial College Press, London, 1998.
28
29
(2) Carbon Nanotubes: Synthesis, Structure, Properties and Applications, vol. 80.
30 Dresselhaus, M. S.; Dresselhaus, G.; Avouris Ph. (Eds.), Springer, Berlin, 2001.
31
32 (3) Delgado, J.L.; Herranz, M. A.; Martín, N. J. Mater. Chem. 2008, 18,1417.
33
34 (4) Jiao, L.; Zhang, L.; Wang, X.; Diankov, G.; Dai, H. Nature 2009, 458, 877;
35
36 (5) Terrones, M. Nature 2009, 458, 845.
37
(6) Joselevich, E.; Dai, H.; Liu, J.; Hata, K.; Windle, A.H., in Carbon Nanotubes:
38
39 Advanced Topics in the Synthesis, Structure, Properties and Applications, Jorio, A.;
40
41 Dresselhaus, G.; Dresselhaus M.S. (Eds.), Topics in Applied Physics, vol. 111,
42
43 Springer-Verlag, Berlin, 2008, p.101.
44
45
(7) Fu, Q.; Liu, J. in Carbon Nanotechnology, Dai L. (Ed.), Elsevier B.V., 2006, p. 81.
46 (8) Kitiyanan, B.; Alvarez, W. E.; Harwell, J. H.; Resasco, D. E. Chem. Phys. Lett.
47
48 2000, 317, 497.
49
50 (9) Herrera, J. E.; Balzano, L.; Borgna, A.; Alvarez, W. E.; Resasco, D.E. J. Catal.
51
52 2001, 204, 129.
53 (10) Bachilo, S. M.; Balzano, L.; Herrera, J. E.; Pompeo, F.; Resasco, D. E.; Weisman,
54
55 R. B. J. Amer. Chem. Soc. 2003, 125, 11186.
56
57 (11) Hata, K.; Futaba, D. N.; Mizuno, K.; Namai, T.; Yumura, M.; Iijima, S. Science
58
59 2004, 306, 1362.
60

17
ACS Paragon Plus Environment
Submitted to The Journal of Physical Chemistry Page 18 of 31

1
2
3 (12) Murakami, Y.; Chiashi, S.; Miyauchi, Y.; Hu, M.; Ogura M.; Okubo, T.;
4
5 Maruyama, S. Chem. Phys. Lett. 2004, 385, 298.
6
7 (13) Noda, S.; Sugime, H.; Osawa, T.; Yoshiko, T.; Chiashi, S.; Murakami, Y.;
8
9 Maruyama, S. Carbon 2006, 44, 1414.
10
(14) Zhang, L.; Tan, Y.; Resasco, D.E. Chem. Phys. Lett. 2006, 422, 198.
11
12 (15) Murakami, Y.; Chiashi, S.; Miyauchi, Y.; Maruyama, S. Chem. Phys. Lett. 2003,
13
14 377, 49.
15
16 (16) Maruyama, S.; Einarsson, E.; Murakami, Y.; Edamura, T. Chem. Phys. Lett. 2005,
17
18
403, 320.
19 (17) J. Nanosci. Nanotech. 2008, 8, 1.
20
21 (18) Einarsson, E.; Murakami Y.; Kadowaki, M.; Maruyama, S. Carbon 2008, 46, 923.
22
23 (19) Futaba, D. N.; Hata, K.; Yamada, T.; Mizuno, K.; Yumura, M.; Iijima, S. Phys.
24
25 Rev. Lett. 2005, 95, 056104.
26
(20) Zhang, G.; Mann, D.; Zhang, L.; Javey, A.; Li, Y. M.; Yenilmez, E.; Wang, Q.;
27
28 McVittie, J. P.; Nishi, Y.; Gibbons, J.; Dai, H. Proc. Nat. Acad. Sci. 2005, 102,
29
30 16141.
31
32 (21) Zhong, G. F.; Iwasaki, T.; Honda, K.; Furukawa, Y.; Ohdomari, I.; Kawarada, H.
33
34
Chem. Vap. Dep. 2005, 11, 127.
35 (22) Eres, G.; Kinkhabwala, A. A.; Cui, H.; Geohegan, D. B.; Puretzky, A. A.;
36
37 Lowndes, D. H. J. Phys. Chem. B 2005, 109, 16684.
38
39 (23) Xu, Y. Q.; Flor, E.; Kim, M. J.; Hamadani, B.; Schmidt, H.; Smalley, R. E.; Hauge,
40
41 R. H. J. Amer. Chem. Soc. 2006, 128, 6560.
42 (24) Baker, R. T. K.; Harris, P. S.; Thomas, R. B.; Waite, R. J. J. Catal. 1972, 26, 51.
43
44 (25) Rostrup-Nielsen, J. R. J. Catal. 1972, 27, 343.
45
46 (26) Rostrup-Nielsen, J. R.; Trimm, D. L. J. Catal. 1977, 48, 155.
47
48 (27) Trimm, D. L. Catal. Rev.-Sci. Eng. 1977, 16, 155.
49
50
(28) Boellaard, E.; De Bokx, P. K.; Kock, A. J. H. M.; Geus, J. W. J. Catal. 1985, 96,
51 468.
52
53 (29) Alstrup, I. J. Catal. 1988, 109, 241.
54
55 (30) Baker, R. T. K. Carbon 1989, 27, 315.
56
57 (31) Alstrup, I.; Tavares, M. T. J. Catal. 1993, 139, 513.
58 (32) Snoeck, J. -W.; Froment, G. F.; Fowles, M. J. Catal. 1997, 169, 240.
59
60 (33) Snoeck, J. -W.; Froment, G. F.; Fowles, M. J. Catal. 1997, 169, 250.
(34) De Jong, K. P.; Geus, J. W. Catal. Rev.-Sci and Eng. 2000, 42, 481.

18
ACS Paragon Plus Environment
Page 19 of 31 Submitted to The Journal of Physical Chemistry

1
2
3 (35) Rodríguez-Manzo, J. A.; Terrones, M.; Terrones, H.; Kroto, H. W.; Sun, L.;
4
5 Banhart, F. Nature Nanotechnology 2007, 2, 307.
6
7 (36) Helveg, S.; Lopez-Cortes, C.; Sehested, J.; Hansen, P. L.; Clausen, B. S.; Rostrup-
8
9 Nielsen, J. R.; Abild-Pedersen, F.; Nørskov, J. K. Nature 2004, 427, 426.
10
(37) Lin, M.; Tan, J. P. Y.; Boothroyd, Ch.; Loh, K. P.; Tok, E. S.; Foo, Y.-L. Nano
11
12 Lett. 2006, 6, 449.
13
14 (38) Yamada, T.; Maigne, A.; Yudasaka, M.; Mizuno, K.; Futaba, D. N.; Yumura, M.;
15
16 Iijima, S.; Hata, K. Nano Lett. 2008, 8, 4288.
17
18
(39) Li, Q.; Zhang, X.; De Paula, R. F.; Zheng, L.; Zhao, Y.; Stan, L.; Holesinger, T. G.;
19 Arendt, P. N.; Peterson, D. E.; Zhu, Y. T. Adv. Mater. 2006, 18, 3160.
20
21 (40) Demichelli, M. C.; Ponzi, E. N.; Ferreti, O. A.; Yeramian, A. A. Chem. Eng. J.
22
23 1991, 46, 129.
24
25 (41) Toebes, M. L.; Bitter, J. H.; Jos van Dillen, A.; De Jong, K. P. Catal. Today 2002,
26
76, 33.
27
28 (42) Villacampa, J. I.; Royo, C.; Romeo, E.; Montoya, J. A.; Del Angel, P.; Monzón, A.
29
30 Appl. Catal. A 2003, 252, 363.
31
32 (43) Pérez-Cabero, M.; Romeo, E.; Royo, C.; Monzón, A.; Guerrero-Ruiz, A.;
33
34
Rodríguez-Ramos, I. J. Catal. 2004, 224, 197.
35 (44) Monzon, A.; Lolli, G.; Cosma, S.; Sayed-Ali, M.; Resasco, D. E. J. Nanosci.
36
37 Nanotech. 2008, 8, 6141.
38
39 (45) Vinten, P.; J. Lefebvre, J.; Finnie, P. Chem. Phys. Lett. 2009, 469, 293.
40
41 (46) Agorreta, E. L.; Peña, J. A.; Santamaría, J.; Monzón, A. Ind. Eng. Chem. Res.
42 1991, 30, 111.
43
44 (47) Louchev, O. A.; Laude, T.; Sato, Y.; Kanda, H. The J. Chem. Phys. 2003, 118,
45
46 7622.
47
48 (48) Tavares, M. T.; Alstrup, I.; Bernardo, C. A.; Rostrup-Nielsen, J. R. J. Catal. 1996,
49
50
158, 402.
51 (49) Kushinov, G. G.; Mogilnykh, Yu. I.; Kushinov, G. G. Catal. Today 1998, 42, 357.
52
53 (50) Zhang, Y.; Smith, K. J. J. Catal. 2005, 231, 354.
54
55 (51) Švrček, V.; Kleps, I.; Cracioniou, F.J.; Paillaud, L.; Dintzer, T.; Louis, B.; Begin,
56
57 D.; Pham-Huu, C.; Ledoux, M.-J.; Le Normand, F. The J. Chem. Phys. 2006, 124,
58 184705.
59
60 (52) Puretzky, A. A.; Geohegan, D. B.; Jesse, S.; Ivanov, I.N.; Eres, G. Appl. Phys. A
2005, 81, 223.

19
ACS Paragon Plus Environment
Submitted to The Journal of Physical Chemistry Page 20 of 31

1
2
3 (53) Sharma, R.; Iqbal, Z. Appl. Phys. Lett. 84 2004, 84, 990.
4
5 (54) Sharma, R.; Rez, P.; Treacy, M. M. J.; Stuart, S. J. J. Electron Microscopy 2005,
6
7 54, 231.
8
9 (55) Sharma, R.; Rez, P.; Brown, M.; Du, G.; Treacy, M. M. J. Nanotechnology 2007,
10
18, 125602.
11
12 (56) Chiashi, S.; ,Kohno, M.; Takata, Y.; Maruyama, S. J. Phys.: Conf. Ser. (Eighth
13
14 Intern. Conf. on Laser Ablation) 2007, 59, 155.
15
16 (57) Kim, D. -H.; Jang, H. -S.; Kim, Ch. -D.; Cho, D. -S.; Yang, H. -S.; Kang, H. -D.;
17
18
Min, B. -K.; Lee, H. -R.; Nano Lett. 2003, 3, 863.
19 (58) Mora, E.; Harutyunyan, A. R. J. Phys. Chem. C 2008, 112, 4805.
20
21 (59) De Chen; Christensen, K. O.; Ochoa-Fernández, E. ; Yu, Z.; Tøtdal, B.; Latorre,
22
23 N.; Monzón, A.; Holmen, A. J. Catal. 2005, 229, 82.
24
25 (60) Dussault, L.; Dupin, J. C.; Guimon, C.; Monthioux, M.; Latorre, N.; Ubieto, T.;
26
Romeo, E.; Royo C.; Monzón, A. J. Catal. 2007, 251, 223.
27
28 (61) Zhang, C.; Pisana, S.; Wirth, C. T.; Parvez, A.; Ducati, C.; Hofmann, S.;
29
30 Robertson, J. Diamond & Relat. Mater. 2008, 17, 1447.
31
32 (62) Zhang, L.; Li, Z.; Tan, Y.; Lolli, G.; Sakulchaicharoen, N. Requejo, F. G.; Mun, B.
33
34
S.; Resasco, D. E. Chem. Mater. 2006, 18, 5624.
35 (63) Benito, P.; Herrero, M.; Labajos, F.M.; Rives, V. Royo, C.; Latorre, N.; Monzon,
36
37 A. Chem. Eng. J. 2009, 149, 455.
38
39 (64) Corella, J.; Asua, J. M.; Ind. Eng. Chem. Proc. Des. Dev. 1982, 21, 55.
40
41 (65) Zhao, J.; Martinez-Limia, A.; Balbuena, P.B. Nanotechnology 2005, 16, S575.
42 (66) Ding, F.; Rosén, A.; Bolton, K. Comp. Mater. Sci. 2006, 35, 243.
43
44 (67) Soman, Ch.; Giorgio, T. Nano Research 2009, 2, 78.
45
46 (68) Szépe, S.; Levenspiel, O. Catalyst Deactivation, in: Proc. 4th Europ. Symp. on
47
48 Chem. React. Eng., Brussels 1968, Pergamon Press, London, 1971. p. 265.
49
50
(69) Bartholomew, C.H. Appl. Catal. A: General 2001, 212, 17.
51 (70) A. Monzón, E. Romeo, A. Borgna, Chem. Eng. J. 94 (2003) 19.
52
53 (71) Rodriguez, J. C.; Peña, J. A.; Monzón, A.; Hughes, R.; Li, K. Chem. Eng. J. 1995,
54
55 58, 7.
56
57 (72) Vannice, M.A. Kinetics of Catalytic Reactions, Springer Science+Business Media,
58 Inc., New York, USA, 2005, pp. 191-192.
59
60 (73) M.J. Lázaro, J.L. Pinilla, I. Suelves, R. Moliner, Int. J. Hydrogen Energy 33 (2008)
4104.

20
ACS Paragon Plus Environment
Page 21 of 31 Submitted to The Journal of Physical Chemistry

1
2
3 (74) D.P. Serrano, J.A. Botas, R. Guil-Lopez, Int. J. Hydrogen Energy 34 (2009) 4488.
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

21
ACS Paragon Plus Environment
Submitted to The Journal of Physical Chemistry Page 22 of 31

1
2
3
4 Table 1. Kinetic parameters of experimental data of VA-SWCNT growth. Comparison
5
6 of the models.
7
8 General Model, described by equation 15
9
10 Data in Figure 4a Data in Figure 4b Data in Figure 5
Parameter
11 (Ref. 18) (Ref. 18) (Ref. 56)
12
13 jC0 (*) 3.118 ± 0.041 5.131 ± 0.046 76.11 ± 3.71
14
15 ψ S (min-1) 7.333 ± 0.767 890.01 ± 12.45 0.00017 ± 7.5 E-6
16
17 K S (*) - - 30485.76 ± 2255.94
18
19 0.167 ±
20 ψ d (min-1) 0.311 ± 0.005 0.945 ± 0.043
21 0.003
22
23
ψ r (min-1) - - 0.057 ± 0.004
24
SSRTotal 0.38 1.18 70.65 (**)
25
26 Exponential Model, described by equation 21
27
28 β ∝ jC (∗)
0
2.453 ± 0.111 5.252 ± 0.133 3.359 ± 0.087
29
30 τ0=1/ψd (min) 4.111 ± 0.304 5.766 ± 0.300 0.130 ± 0.085
31
32 SSRTotal 7.86 13.19 2456.67 (**)
33
(*) Units corresponding to the data in the referred figures
34
35 (**) In this case the SSR2 is zero because there is no growth rate data.
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

22
ACS Paragon Plus Environment
Page 23 of 31 Submitted to The Journal of Physical Chemistry

1
2
3
4 Table 2. Application of General Model to Data in Figures 6a and 6b. Reference 42.
5
6 % CH4 2.5% 5.0% 7.5% 10.0%
7
8
jC0 5.91E-03 ± 1.11E-02 ± 1.53E-02 ± 1.81E-02 ±
9 (g C/g cat.min) 1.14E-04 7.27E-05 4.19E-04 4.47E-04
10 ψS 1.16E-03 ± 2.34E-03 ± 3.91E-03 ± 8.40E-03 ±
11 -1
12 (min ) 3.58E-05 6.85E-05 7.56E-06 1.64E-04
13 KS 502.372 ± 502.372 ± 502.371 ± 502.368 ±
14
15 (g cat./g C) 12.790 14.653 13.231 11.123
16 ψd 0.018 ± 0.032± 0.046± 0.062±
17 (min ) -1 4.453E-04 8.467E-05 2.366E-03 1.792E-03
18
19 ψr 5.448E-03±
0 0 0
20 (min-1) 1.88E-04
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

23
ACS Paragon Plus Environment
Submitted to The Journal of Physical Chemistry Page 24 of 31

1
2
3 Figure captions
4
5
6
7
8 Figure 1. Effect of the parameter KS on the reaction rate and CNT concentration. Case
9
10 with catalyst deactivation.
11
12 Figure 2. Effect of the parameter KS on the reaction rate and CNT concentration. Case
13
14 without catalyst deactivation.
15
16 Figure 3. Effect of the parameter ψS on the reaction rate and CNT concentration.
17
18
Figure 4. Fittings comparison of VA-SWCNT film thickness and growth rate data taken
19
20 from figures 2a and 2b of reference 18. (Figure 4a) data corresponding to γ0=2.7
21
22 µm/min and τ=3.6 min ; (Figure 4b) data corresponding to γ0=5.2 µm/min and τ=6 min.
23
24
25
Figure 5. General Model fitting of G-band intensity data taken from figure 4b of
26 reference 56.
27
28
29 Figure 6. Simultaneous fitting of CNT concentration (Figure 6a) and growth rate
30 (Figure 6b) data from reference 42.
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

24
ACS Paragon Plus Environment
Page 25 of 31 Submitted to The Journal of Physical Chemistry

1
2
3
4
5
6
7
8 12 0.8
ψd=0.08 min
-1
9 KS=1 KS=1
10 -1 0.7
11 10 ψr=0.005 min
12
10
2 0.6
13 2
14 8 10
mC (g C/g cat.)

0.5

rC (g C/g cat.min)
15
16
4
17 6 10
4 10 0.4
18
19 6
6
10 0.3
20 4 10
21 8 8
22 10 10 0.2
23 2
24 [KS]≡[g cat./g C] 0.1
25
26 0 0.0
27 0 5 10 15 20 25 30 35 40 45 50
28
29 Time (min)
30
31
32
33 Figure 1.
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

25
ACS Paragon Plus Environment
Submitted to The Journal of Physical Chemistry Page 26 of 31

1
2
3
4
5
6
7
8 50 1.2
9 KS=1
2 4 6 8
10 KS=1 10 10 10 10
2
11 10 1.0
40
12 4
13 10
14 6 0.8
10
mC (g C/g cat.)

rC (g C/g cat.min)
15 30
16 8
10
17 0.6
18
19 20
20 0.4
21
22
10
23
24
ψd=0, ψr=0 0.2

25 [KS]≡[g cat./g C]
26 0 0.0
27 0 5 10 15 20 25 30 35 40 45 50
28
29 Time (min)
30
31
32
33 Figure 2.
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

26
ACS Paragon Plus Environment
Page 27 of 31 Submitted to The Journal of Physical Chemistry

1
2
3
4
5
6
7
8 14 0.9
ψd=0.08 min
-1
9 ψS=1.953
10 ψS=1.953 0.8
12
ψr=0.005 min
-1
11 0.781
12 0.781 0.7
13 10
0.313
mC (g C/g cat.)

14 0.6

rC (g C/g cat.min)
15
16 8
0.313 0.5
17 0.125
18 6 0.4
19
20 0.3
0.125
21 4 0.05
22 0.2
23
2 0.05
24 -1
[ψS]≡[min ] 0.1
25
26 0 0.0
27 0 5 10 15 20 25 30 35 40 45 50
28
29 Time (min)
30
31
32
33 Figure 3.
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

27
ACS Paragon Plus Environment
Submitted to The Journal of Physical Chemistry Page 28 of 31

1
2
3
4 10 3.0
5 General model
6 Exponential model
7 VA-SWCNT film thickness (µm) 2.5
8 8
9
10

Growth rate (µm/min)


2.0
11 6
12
13 1.5
14
15 4
16 1.0
17
18
2
19 0.5
20
21
22 0 0.0
23 0 1 2 3 4 5 6 7 8 9 10
24
Growth time (min)
25
26
27
28 Figure 4a.
29
30
31 30 6
32
General model
33
Exponential model
34 25 5
VA-SWCNT film thickness (µm)

35
36
37
Growth rate (µm/min)
20 4
38
39
40 15 3
41
42
43 10 2
44
45
46 5 1
47
48
49 0 0
50 0 1 2 3 4 5 6 7 8 9 10
51
52
Growth time (min)
53
54
55 Figure 4b.
56
57
58
59
60

28
ACS Paragon Plus Environment
Page 29 of 31 Submitted to The Journal of Physical Chemistry

1
2
3
4
5
6
7
8 20 8
9
18
10 7
11 16
12 6
13 14
Intensity G band (a.u.)

14

Growth rate (min )


General Model 5
15 12
(G band signal)
16
10 Exponential Model 4
17
(G band signal)
18 General Model
8
19 (growth rate) 3
20 6 Exponential Model

-1
21 (growth rate) 2
22 4
23
1
24 2
25
26 0 0
0 2 4 6 8 10
27
28 Time (min)
29
30
31
32
33 Figure 5.
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

29
ACS Paragon Plus Environment
Submitted to The Journal of Physical Chemistry Page 30 of 31

1
2
3
4
5
6
0.30
7
8
9 CH4 conc.: 10 %
0.25
10
11
7.5%
CNT conc. (gC/g cat)

12 0.20
13
14 5%
15 0.15
16
17
18 0.10 2.5%
19
20
21 0.05
22
23
24 0.00
25 0 25 50 75 100 125 150
26 CVD time (min)
27
28
29
30 Figure 6a.
31
32
33 0.016
34
35
36
37 CH4 conc.: 10 %
Growth rate (gC/g cat.min)

38 0.012
39
40 7.5%
41
42 0.008
43
44 5%
45
46
47 0.004
48 2.5%
49
50
51
0.000
52
0 25 50 75 100 125 150
53
54 CVD Time (min)
55
56
57
58 Figure 6b.
59
60

30
ACS Paragon Plus Environment
Page 31 of 31 Submitted to The Journal of Physical Chemistry

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60

31
ACS Paragon Plus Environment

You might also like