Proceedings of HT2005 2005 ASME Summer Heat Transfer Conference July 17-22, 2005, San Francisco, California, USA HT2005-72172 Thermal Transport in Nanotube Composites for Large-Area Macroelectronics Satish Kumar Mohammad A. Alam Department of Mechanical Engineering Department of Electrical and Computer Engineering Purdue University Purdue University 585 Purdue Mall 465 Northwestern Avenue W. Lafayette IN 47907 W. Lafayette IN 47907 Jayathi Y. Murthy Department of Mechanical Engineering Purdue University 585 Purdue Mall W. Lafayette IN 47907 emerging as an alternative in many non-display applications. For high performance applications, however, the choices are limited: single crystal silicon or poly-silicon based TFTs [8,9] cannot be manufactured at low temperature (<200C) and are therefore not suitable for plastic substrates. As a result, researchers are exploring a new class of nano-composite TFTs based on bundles of silicon nanowires (Si-NWs) or carbon nanotubes (CNTs) [10-12]. Here, high-quality, nearlycrystalline NWs and CNTs are grown at high temperature on a temporary substrate, then released from the temporary substrate into a carrier fluid, and finally, the wire-saturated carrier fluid is spin-coated onto arbitrary (flexible) substrates at room temperature to form a thin film of randomly-oriented NWs or CNTs. Once the source/drain contacts are defined, this thin film of nearly crystalline nanowires or nanotubes constitutes the high performance channel of a TFT (see Fig. 1). The unique physical properties of carbon nanotube composites make them a promising novel material for use in TFTs. Individual nanotubes exhibit excellent electrical and thermal conductivities [13-15] and high Young’s modulus and strength-to-weight ratio [16]. However, it is yet unclear to what extent these properties will be manifested in polymer composites, where scattering at polymer/tube boundaries and at tube/tube contacts may substantially decrease electrical mobility and thermal conductivity. For silicon nanowires, a significant decrease in axial thermal conductivity has been noted in freestanding nanowires either due to phonon boundary scattering or due to phonon confinement effects [15]. Previous experiments and numerical simulations report that the interfacial contact conductance between nanotubes and substrate [18] may be extremely low. Similarly, the contact resistance between ABSTRACT Thermal transport in a new class of nanocomposites composed of isotropic 2D ensembles of nanotubes or nanowires in a substrate is considered for use as the channel region of thin film transistors. The random ensemble is generated numerically and simulated using a finite volume scheme. The effective thermal conductivity of a nanotube network embedded in a thin substrate is computed. Percolating conduction in the composite is studied as a function of wire/tube densities and channel lengths. The conductance exponents are validated against available experimental data for long channels devices. The effect of tube-tube contact conductance, tube-substrate contact conductance and substrate-tube conductivity ratio is analyzed for various channel lengths. It is found that beyond a certain limiting value, contact parameters do not result in any significant change in the effective thermal conductivity of the composite. It is also observed that the effective thermal conductivity of the composite saturates beyond a limiting channel-length/tube length ratio for the range of contact parameters under consideration. INTRODUCTION In recent years, there has been growing interest in lowcost large-area manufacture of thin film transistors (TFTs) on flexible substrates for use in applications such as displays, epaper, e-clothing, biological and chemical sensing, conformal radar, and others. TFTs based on amorphous silicon (a-Si) now dominate the market for large-area flat-panel displays [1-2]. When transistor performance is not critical, low-cost organic TFTs on flexible, lightweight, plastic substrates [3-7] are 1 Copyright © 2005 by ASME nanotubes has a major influence on the transport properties of the composite. There is little in the literature that provides a framework for understanding, controlling and designing nanowire and nanotube composites suitable for TFTs. The network of nanotubes used for the channel region is embedded in a substrate such as plastic, and is covered with glass on one side and polymer on the other. Source and drain contacts are deposited on this 2D matrix, and typically, the device is backgated. Thus the network layer is very thin, and may essentially be considered two-dimensional. Moreover, unlike bulk composites, source and drain may only be a few microns apart in typical macroelectronics, and the channel length may be comparable to the length of the tube or wire. Thus analyses employing classical percolation theory for periodic ensembles are not suitable. The finite length of the sample must be accounted for in the analysis. There are few experimental, theoretical and computational models which attempt to predict the thermal conductivity of nanotube composites. Biercuk et al. has reported 125 % increase in thermal conductivity of epoxySWNT composites for 1 wt % SWNT loading at room temperature [16]. Liu et al. used a silicon elastomer as the matrix and CNT as the filler in their experiments and reported 65 % enhancement in thermal conductivity with 3.8 wt % CNT loading [19]. Nan et al. employed an effective medium theory and predicted that the interfacial resistance can significantly degrade the thermal conductivity of the CNT composite [17,20]. Lusti and Gusev predicted the thermoelastic properties of nanotube-reinforced polymers using the finite element method [21]. Yang and Chen has used the phonon Boltzmann equation to study the phonon thermal conductivity of nanocomposites using a periodic two-dimensional model [22]. To our knowledge, there have been no published analyses of thermal transport in thin film composites based on nanotube/nanowire bundles for TFT applications. Our ultimate interest is in the coupled electro-thermal analysis of nanotube bundle transistors. Heat dissipation by the current carrying nanotubes can degrade the performance of the transistor. In this paper, the effective lateral thermal conductivity of finite planar nanotube/nanowire composites is investigated. The present work is a first step towards building a self-consistent electro-thermal model to compute the temperature distribution and I-V characteristics of network transistors. We analyze the geometry-dependent conductance properties of the composite for tube densities above the percolation threshold using a finite volume method. The general framework may be applied to both tubes and wires. The analysis is used to predict the conductance of the composite as a function of wire/tube density and channel length. Simulations are performed to study the effect of contact resistance between nanotubes as well as nanotube and substrate. The effect of substrate to nanotube conductivity ratio on transport properties is also analyzed. NOMENCLATURE A tube cross-sectional area Bic contact-conductance parameter between tubes Bis contact-conductance parameter between tube and substrate d diameter of tube h c, h s heat transfer coefficient characterizing tube-tube contact and tube-substrate contact H height of the channel keff effective lateral thermal conductivity kt ks thermal conductivity of tube LC Lt n P c, P s t channel length tube length conductance exponent contact perimeters substrate thickness temperature drain temperature non-dimensional temperature temperature difference across the channel displacement vector from tube segment centroid to substrate cell centroid conductance 0 scaling parameter for conductance th * v tube density tube density at percolation threshold T Td T thermal conductivity of substrate th contact geometry parameter ROD GENERATION In the present analysis, we consider a percolating random network of nanotubes or nanowires of length Lt and diameter d randomly dispersed in the mid-plane of a matrix of thickness t. Thus the nanotube network is essentially 2D, while the matrix containing it is 3D. The geometry in the mid-plane is shown in Fig. 1. The domain is assumed to be of height H, and the top and bottom boundaries are assumed periodic. The channel length is LC. For LC<Lt some tubes would span the source-to-drain distance LC, while others may not, depending on orientation. The source, drain and channel regions in Fig. 1 are divided into finite rectangular control volumes. A fixed probability p of a control volume originating a nanotube is chosen a priori. A random number is picked from a uniform distribution and compared with p. If it is less than p, a nanotube is originated from the control volume. The length of source and drain for tube generation is Lt, which ensures that any tube that could penetrate the channel region either from the left or the right is included in the simulations. The orientation of the tube is also chosen from a uniform random number generator. Since 2 Copyright © 2005 by ASME contact parameter v. In Eq. 1(a), the summation term denotes heat exchange between tube i and all tubes j which have an intersection with it; the term is non-zero only at the point of intersection. Similarly, in Eq.1(b), the summation terms denotes the volumetric source due to tubes intersecting the substrate. The dimensionless parameters are defined as: the tube length is fixed at Lt, all tubes may not span the channel region even for shorter channel lengths LC. Tubes crossing the top and bottom boundaries are treated assuming translational periodicity; that part of the tube crossing one of these boundaries reappears on the other side. Tube-tube intersections are computed from this numerically generated random network and stored for future use. The analysis is conducted only on the tubes that lie in the channel region. Bic Here, hc and hs are the heat transfer coefficients characterizing tube-to-tube and tube-to-substrate contact, Pc and Ps are the corresponding contact perimeters, kt is the thermal conductivity of the tube and A its cross-sectional area. The parameter v characterizes contact geometry and v is the contact area per unit volume of substrate. For low volume fractions, a starting estimate for Bis may be computed assuming cylindrical wires in an infinite medium [23]. Bic is more difficult to characterize and requires detailed contact modeling. Because of the uncertainty in these values, both are treated as parameters in the present study. Additional dimensionless parameters governing the problem are the conductivity ratio ks /kt, tube aspect ratio d/Lt, as well as the geometric ratios LC /Lt, t /Lt and H/ Lt.. The probability p of nanotube origination, described previously, is related to the tube density ; the corresponding dimensionless parameter is *, which is obtained by normalizing with the percolation threshold th. The percolation threshold for the network is estimated as the density at which the average distance between nanotubes equals the average length of the H S Lt D LC Lt Figure 1: Computational domain showing nanotubes, source (S), drain (D) and channel region. tubes, so that th ~ 1 LS GOVERNING EQUATION Though microscale heat transfer effects such as phonon ballistic transport and confinement may be important, in some regimes, boundary scattering effects are expected to dominate in long tubes, and Fourier conduction may be assumed as a starting point, albeit with a thermal conductivity that may differ significantly from bulk values. Using the dimensionless variable =(T-Td)/( T ) and nondimensionalizing all lengths by the tube diameter d, the governing equations in the tubes and substrate may be written as: d 2 i ds *2 Ak hc Pc d 2 h Pd2 ; Bis s s ; v v t kt A kt A Ps ks 2 . All computations done in this study are restricted to d/Lt = 1000.0 and t/d=2.0. The equations for charge transport in a resistor are analogous to the thermal transport equations in the Fourier conduction limit. In this limit, drift-diffusion theory based on Kirchoff’s law may be used to compute the potential distribution in the network, analogous to the temperature distribution. For electrical transport, the i term in Eq. 1 (a) now represents the dimensionless potential distribution along the tube. The quantity Bic is the dimensionless charge-transfer coefficient between tubes i and j at their intersection point and is specified a priori. Only those segments with intersections with other tubes include the Bic term in their discretization. The Bis term is zero for the charge transport since the substrate may be considered insulating. Boundary conditions s =1.0 (source) and s =0 (drain) are applied at the two ends of the channel region for the substrate. Similarly, for the tube tips embedded in the source, i =1.0 is imposed, while for the tube-tips embedded in the drain i =0 is imposed. All the tube tips terminating inside the substrate are assumed adiabatic. The top and bottom boundaries of channel are assumed as periodic boundaries for both substrate and tubes. The remaining two boundaries (parallel to the network plane) are assumed adiabatic in order to assess the lateral thermal conductivity of the composite. Bic j i Bis s i 0 intersecting tubes j 1(a) N tubes *2 s Bis v i s 0 i 1 1(b) Here, i (s ) is the non-dimensional temperature of the ith tube at the axial location s*, and s(x*,y*,z*) is the substrate temperature. Thermal contact between tubes i and j is characterized by the contact Biot number Bic. Heat exchange between each tube and the substrate is governed by substrate Biot number Bis and the * 3 Copyright © 2005 by ASME corresponding coupling terms in tube and substrate equations (Eq. 1(a) and 1(b)): NUMERICAL METHOD The finite volume method [24] is used to solve the temperature field in the tubes and the substrate. Each tube is divided into 1-D segments, and a control volume balance is performed on each tube segment. Possible contact with other tubes is checked for each segment. Bic is zero for the segment if there is no contact. Similarly, the substrate is divided into 3-D control volumes, and a control volume balance is performed. Segments of tubes located in each substrate control volume are identified and their heat exchange is ascribed to the control volume to ensure conservation. Second-order accurate central differencing is used for discretization. The discrete equations for all the tubes and the substrate are solved sequentially and iteratively until convergence. For the long channel lengths and large numbers of tubes, the sequential solver eventually stalls due to the high degree of coupling between tubes and also between the tubes and substrate. For these situations, a direct sparse solver developed by Kundert [25] is used to solve the resulting system of equations. Most of the results reported here are computed by taking an average over 100 random realizations of the network, though more realizations are used for low densities and short channel lengths. tube centroid substrate substrate centroid (2) where is the position vector of tube segment centroid relative to the cell centroid of substrate, as shown in Fig. 2. COMPUTATUTION OF EFFECTIVE LATERAL THERMAL CONDUCTIVITY The lateral thermal conductivity is computed using the expression: K t A d K A d ds substrate s s dx tubes K eff Ht LC where the first term in the numerator is heat flow through the tubes in the lateral direction, while the second term represents heat flow in the substrate in lateral direction. The heat flow in both tubes and substrate is computed at the source-channel junction, where the s =1.0 boundary condition is imposed. Here A , As and represents cross-sectional area of the tube, face-area of a control volume of substrate perpendicular to lateral (x) direction and the non-dimensional temperature drop across the channel. GRADIENT CALCULATION FOR COUPLING TERMS While solving Fourier conduction equations for the tubes and the substrate, the tube temperature (i) and substrate temperature (s) in the coupling terms of the equation are calculated at the centroid of the tube segments and the centroid RESULTS GENERAL BEHAVIOR The temperature distribution in the substrate and in the tube network is shown in Fig. 3 (a) and (b) for the case LC/Lt = 2.0, H/Lt = 2, Bic = 10.0, Bis = 1.0e-5, ks/kt = 0.001 and * = 18.0, corresponding to LC =4m, Lt = 2m, H = 4m, and = 4.5 m-2. Contours of constant temperature in the substrate would be one-dimensional in x for Bis = 0, but due to the interaction with the tubes, distortion in the contours is observed, consistent with the temperature plots in the tube in Fig. 3 (b). For ks/kt <<1, the tubes convey the boundary temperature much further into the interior that does the substrate, leading to a local energy source in the substrate. The distortion is related to the local density of the tubes. Tube centroid Cell centroid Fig. 2. A substrate control volume with two segments of a tube lying inside it. The displacement vector from substrate cell centroid to tube segment centroid is shown. of the substrate cell under consideration. The centroid of the substrate cell does not coincide with the centroid of tubesegments lying inside the cell, as shown in Fig. 2. This can introduce significant error in the temperature distribution since the substrate discretization is usually coarser than that of the tube. To account for this effect, the substrate temperature gradient in the plane of the nanotube network is computed and is used to interpolate the substrate temperature to the tube centroid location. This interpolated temperature is used in the 4 Copyright © 2005 by ASME using a mesh of size 100 segments per tube and a mesh of 10x20x1 cells in the substrate. 0.08 0.26 0.44 0.62 0.80 Y/L (a) (b) X/L Fig. 4. Increment in composite effective thermal conductivity (keff) for two different grid sizes in the substrate and rod. LC/Lt = 2.0 H/Lt = 2, Bic = 10.0, Bis = 1.0e-5, ks/kt = 0.001 and * = 10.0. NETWORK CONDUCTANCE In the absence of the substrate, the temperature and electrical potential distributions are analogous, and electrical conductance measurements [11] may be used to establish the validity of the network model. The conductance of the network is computed by taking an average over 200 random realizations of different orientations. The length of the tubes in [11] ranges from 1 to 3 m. The length distribution of nanotubes has not been reported in [11]. Therefore, we have taken a simple approach and assumed that all tubes are 2 m long in this simulation.. The percolation threshold for the network is estimated to be 0.25 m-2. Simulations are performed for densities in the range 1-10 m-2, for channel lengths varying from 1 to 25 m and with a width H of 90 m, corresponding to the dimensionless parameters LC/Lt ~ 0.5-12.5, H/Lt = 45 . The device dimensions and tube lengths are chosen to match those of the experiments of Snow et al. [11]. In Fig. 5 (a), the network conductance is shown as a function of LC/Lt for several tube densities above the percolation threshold for nearly perfect tube-tube contact (i.e. Bic ~ 5.0). For long channels (LC < Lt) there are no tubes directly bridging the source and drain and heat (current) can only flow because of the presence of the network. If the tube density is sufficiently high (greater than the percolation threshold), a continuous thermal (electrical) path exists from source to drain, and is seen to be non-zero even for LC/Lt >1. Fig. 5 (a) shows that the conductance exponent, n, is close to 1.0, which, in the electrical analogue, corresponds to ohmic conduction for the high densities (=10m-2; *=40). The exponent increases to -1.80 at lower densities (1.35m-2; *=5), Fig. 3. Non-dimensional temperature distribution in (a) substrate (b) tube network. LC/Lt = 2.0, H/Lt = 2, Bic = 10.0, Bis = 1.0e-5, ks/kt = 0.001 and * = 18.0. GRID INDEPENDENCE Grid independence tests were conducted for the case of LC/Lt = 2.0, H/Lt = 2, Bic = 10.0, Bis = 1.0e-5, ks/kt = 0.001 and * = 10.0 (LC =4m, H = 4m, = 2.5 m-2) for nondimensional channel length LC/Lt varying from 0.5-9 (LC = 1-18 m) Fig. 4 shows the percentage change in the composite thermal conductivity for two different grid sizes. For the first case there are 100 cells/unit tube length and 10 cells/unit tube length for the substrate i.e. a channel having a length equal to the tube length is divided into 10 cells. The second case corresponds to 200 segments per tube, with 20 substrate cells per unit length of tube. The grid spacing is the same in the y direction in the substrate. Only one cell is used in the direction normal to the network layer. The results are seen to differ by less than 0.5% between the two cases. The simulations presented in the rest of this paper were therefore performed 5 Copyright © 2005 by ASME indicating a non-linear dependence of conductance on channel length. Note that the asymptotic limit of the conductance Some evidence of this is visible in the scatter in the experimental data at low densities. The dependence of conductance exponent on channel length is explored in Fig. 5(b) for Bic ~ 5.0 and for densities in the range 1.35-10m-2, corresponding to * values of 5-40. For densities >3.0 m-2 (* >12), the exponent approaches -1.0 with increasing channel length (the ohmic limit). Larger exponents, corresponding to non-ohmic transport, are observed for the shorter channel lengths. This is consistent with experimental observations, where conductance is seen to scale more rapidly with channel length for small LC [11]. SUBSTRATE-TUBE CONTACT CONDUCTANCE EFFECT Recently, experimental studies conducted in [18] suggest that heat transport in nanotube composites may be limited by exceptionally small interfacial thermal conductance values. The value of the interfacial resistance between the carbon nanotube and the substrate was reported to be 8.3x10-8 m2 K/W [17,18] for carbon nanotubes in D2O. The nondimensional contact parameter Bis evaluated using this contact conductance is 10-5. Values of Bis in the range 10-1-10-5 were considered in this study. There are no reliable guidelines to determine the value of the tube-tube contact parameter Bic, and a range 10-5 – 1.0 is chosen. For the thermal conductivity ratio, again, a wide variation is possible. Free-standing MWCNTs have exhibited thermal conductivity values of 3000 W/mK or more [13,14], but these values are expected to fall substantially when the CNTs are encased in plastic. Free-standing Si-NWs show an order of magnitude reduction with respect to bulk thermal conductivity [15]. The thermal conductivity of typical plastic substrates is in the range of 0.1-1 W/mK. In this study, we explore a range of ks/kt of 10-1 – 10-3. Fig. 5 (a) Computed conductance dependence on channel length for different densities () in the strong coupling limit (Bic = 5) is compared with experimental results from [11]. 0 = 1.0 (simulation), 0 = 1.0 (experimental) for =10.0 m-2). 0 =1 (simulation) and 0 =1.40 (experimental) for =1.0 m . The -2 number after each curve corresponds to and the number in [ ] corresponds to the value used in the experiments from [11]. (b) Dependence of the conductance exponent n on channel length for different densities based on Fig. 6(a) (i.e. / 0 ~ LC ). Other parameters, chosen to reflect the n experimental conditions in [11], are: H=90m, Lt = 2 m and LC=1 - 25m. exponent for infinite samples with perfect tube/tube contact is 1.97 [26,27]. The linear scaling of conductance with LC for high densities agrees very well with experimental observations of Snow et al. [11]. The observed non-linear behavior for low density is expected because the density value is close to the percolation threshold. Snow et al. reported a conductanceexponent of -1.80 for a density of 1.0 m-2 and channel length > 5m. For the same device dimensions, this value of the exponent is close to that obtained from our simulations for a density of 1.35m-2. Small variations in experimental parameters such as tube diameter, nanotube contact strength, tube electronic properties as well as the presence of a distribution of tube lengths (1-3m), which is not included in the simulation, may explain the difference. The contact resistance between the nanotubes and the source and drain electrodes as well as insufficiently large samples for ensemble averaging in the experimental setup may also be responsible. Fig. 6. Effect of substrate-tube contact conductance on keff for varying channel length. LC/Lt = 2.0, H/Lt = 2, Bic = 10.0, ks/kt = 0.001 and * = 10.0. 6 Copyright © 2005 by ASME The percentage increase in keff of the composite is plotted against LC/Lt ratio for different Bis in Fig. 6. The tubetube contact parameter is held at Bic =10, denoting nearly perfect contact. The conductivity ratio is 10-3, denoting highly conducting tubes in a relatively insulating substrate. The overall shape of the curve reflects the percolation behavior of the network for this low value of ks/kt. A sharp increase in keff is observed for shorter channel lengths as a result of highly conducting tubes directly bridging source and drain. As the channel length increases, thermal conduction is dominated by the network conductance, which achieves invariance beyond LC/Lt >6 or so. The tube-substrate contact parameter Bis ceases to be limiting for Bis>10-3, and the keff variation with Lc/Lt becomes independent of Bis beyond this value. Since the tubes are much more conducting than the substrate, large increases in thermal conductivity are observed. For LC/Lt ~ 0.5, a 280 % increase in thermal conductivity is observed even for low values of Bis (~ 1.0e-5 ), while increasing Bis to 1.0e-1 leads to a 375% increase in the thermal conductivity. For larger channel lengths, increases of 175 % and 140 % in thermal conductivity are observed for Bis ~ 1.0e-3 and Bis ~1.0e-5. Since the ks/kt value used here is expected to be typical of many composites for TFT applications, the results in this section demonstrate that if the percolation properties of the network could be maintained by high tube-tube contact, the network itself could provide a pathway for heat removal. the overall shape of the curves is similar to that in Fig. 6. Decreasing Bic from 1.0 to 10-5 decreases keff by about 50% for long channels. Beyond Bic >1.0e-2, tube-tube contact is sufficiently perfect that it ceases to be limiting; consequently the keff curves become coincident in Fig. 8. Previous theoretical models for computing keff of nanotube composites ignore contact conductance between the nanotubes [17,20]. The present analysis reveals that for tube densities higher than the percolation threshold, Bic emerges as an important parameter controlling keff for low values of ks/kt. CONDUCTIVITY RATIO EFFECT We now consider the effect of thermal conductivity ratio on keff. Simulations are performed for ks/kt ratio varying from 10-1 to 10-3 keeping other parameters constant LC/Lt = 2.0, H/Lt = 2, Bic = 10.0, Bis = 1.0e-5 and * = 10.0 (LC =4m, H = 4m, = 2.5 m-2). For the high value of tube-tube contact considered here, Fig. 8 shows that percolating conduction in the network dominates for ks/kt=10-3, as expected; this is evidenced by the steep rise in keff for low Lc/Lt. As ks/kt is increased, some evidence of percolation may still be detected for ks/kt <5x10-3. For higher values, though, network conductance ceases to be a dominant contributor and the increase in keff over ks drops to zero, signifying that the substrate now dominates conduction through the composite. These computations point to the necessity of accurately characterizing the thermal conductivity of CNTs and Si NWs embedded in substrates. If the presence of the substrate substantially reduces kt, it is possible that ks/kt would be relatively high and the network would no longer provide a significant pathway for heat transfer even if effective tube-tube contact could be maintained. In this limit, the large surface area of contact between the tubes and the substrate allows heat to leak from the network to the substrate, and the primary mechanism for heat removal would be the substrate. However, TUBE-TUBE CONTACT CONDUCTANCE EFFECT In this section, the effect of tube-tube contact is investigated by varying Bic in the range 100-10-5. Fig. 7. Effect of tube-tube contact conductance on keff for varying channel length. LC/Lt = 2.0, H/Lt = 2, Bis = 1.0e-5, ks/kt = 0.001 and * = 10.0. Fig. 7 show the plots for different Bic, while other parameters are kept constant at LC/Lt = 2.0, H/Lt = 2,Bis = 1.0e-5, ks/kt = 0.001 and * = 10.0 ( corresponding to LC =4m, H = 4m, = 2.5 m-2). As expected for this low value of ks/kt, the overall behavior is dominated by the network conductance and Fig. 8. Effect of substrate-tube conductivity ratio on keff for varying channel length. LC/Lt = 2.0, H/Lt = 2, Bic = 10.0, Bis = 1.0e-5 and * = 10.0. 7 Copyright © 2005 by ASME to vary linearly with the density of the nanotubes. The range of tube-densities considered here is much higher than the percolation threshold (~ 0.25 m-2), but the volume fraction of the tubes is still relatively low, and may be considered to be in the dilute limit. This linear behavior suggests that the increase in thermal conductivity is directly proportional to the volume fraction of nanotubes, which is in agreement with the model proposed by Nan et al. [17] in the dilute limit. since both substrate and network conductance would be low in this limit, high channel temperatures and degradation of electrical performance would be expected. DENSITY EFFECT Nanotube density plays an important role in both the electrical and thermal performance of the composite. On the one hand, high densities tend to increase transport in the composite by increasing the number of available highconducting pathways between source and drain. On the other hand, high densities also imply many more contact locations between tubes, and a concomitant increase in the number of scattering locations for both electrons and phonons, leading to a drop in tube electrical and thermal conductivity. Furthermore, higher current in the network could lead to high local heat generation which could burn nanotubes and eventually cause transistor breakdown. The present analysis assumes kt to be known a priori, and does not consider coupled electro-thermal transport. However, the direct effect of tube density on keff may be deduced from our analysis. Fig. 9 shows the effect of tube density on thermal conductivity enhancement for LC/Lt = 2.0 (LC =4m), H/Lt = 2 (H = 4m), Bic = 10.0, Bis = 1.0e-5 and ks/kt=0.001. CONCLUSIONS A computational diffusive transport model has been developed to explore the dependence of thermal conductance and conductivity on the channel length for thin films made of random percolating nanotubes. In the limit of an insulating substrate, model predictions are compared with measurements of electrical conductance and found to be in good agreement. Transport in finite networks with tube-to-tube contact for medium channel lengths (LC/Lt ~1-10) is studied and provides insight into the dominant mechanisms for thermal transport in these composites. For low values of ks/kt, typical of highly conducting CNTs in plastic substrates, percolating conduction in the network is seen to dominate over a wide range of tube-tube and tube-substrate contact parameters. On the other hand, as ks/kt increases, the resistance to heat flow offered by the substrate is no longer limiting. The high surface area of contact between the substrate and the network implies that heat will leak from the network to the substrate, and be transported primarily by the substrate; in this limit, percolating conduction in the network ceases to be relevant. This switchover signals a significant decrease in effective thermal conductivity and is critical to characterize accurately. Accurate measurements of kt as well as tube-tube and tube-substrate contact resistance are necessary. The present analysis employs Fourier theory as a basis. It is likely that the large aspect ratio of tubes implies the predominance of boundary scattering, and diffusive transport is likely to dominate. However, the value of kt is expected to depart substantially from bulk, and would not be known a priori. Effects such as the influence of tube-substrate and tubetube scattering events in decreasing kt may be determined computationally by using molecular dynamics of tube and substrate. Ultimately, the self-consistent computation of device I-V performance requires a similar characterization of electrical transport in the network under drift-diffusion assumptions, and the coupling of electrical and thermal computations. Fig. 9. Effect of tube density on keff for varying channel length. LC/Lt = 2.0, H/Lt = 2, Bic = 10.0, Bis = 1.0e-5 and ks/kt = 0.001 Since all other parameters are held constant with respect to Figs. 6 and 7, and all tube densities are above the percolation threshold, the overall shape of the curves in Fig. 9 indicate the dominance of the network in determining keff.. As expected, higher tube densities result in an increase in keff as the number of available pathways increases; the present model is not capable of incorporating the effect on increased scattering in decreasing kt. Substantial increases in keff are possible for higher densities, with values of over 500% for long channels. For higher densities, the increase in thermal conductivity is seen ACKNOWLEDGMENTS Support of S. Kumar and J. Murthy under NSF Grant EEC-0228390 and the Purdue Research Foundation is gratefully acknowledged. REFERENCES [1] Kagan, C. R., and Andry, P., 2003, “Thin Film Transistors,” Marcel Dekker, Inc. New York. 8 Copyright © 2005 by ASME [2] Madelung, O., (ed.), “Technology and Applications of Amorphous Silicon,” Springer, Berlin, 2000. [15] Li, D., Wu, Y., Kim, P., Shi, L., Yang, P. and Majumdar, A., 2003, “Thermal conductivity of individual silicon nanowires,” App. Phys. Lett, 83, no.14, pp. 2934-2936. [3] Dimitrakopoulos, C., and Mascaro, D., 2001 “Organic thinfilm transistors A review of recent advances,” IBM J. Res. And Dev., 45, pp. 11-27. [16] Bieurck, M. J., Liaguno, M. C., Radosavljevic, M., Hyun, J. K., and Johnson, A. T., 2002, “Carbon nanotube for thermal management,” Appl. Phys. Lett., 80 (15), pp. 2767-276. [4] Curioni, A., and Andreoni, W., 2001 “Computer simulations for organic light-emitting diodes,” IBM J. Research and Dev., 45, pp. 101-113. [17] Nan, C. W., Liu, G., Lin, Y., and Li, M., 2004, “Interface effect on thermal conductivity of carbon nanotube composites,” Appl. Phys. Lett., 85 (16), pp. 3549-355. [5] Wisnieff, R., 1998 “Printing Screens,” Nature, 394, pp. 225. [18] Huxtable, S. T., Cahill, D. G., Shenogin, S., Xue, L., Oziski, R., Barone, P., Usrey, M., Strano M. S., Siddons, G., Shim, M., and Keblinski, P., 2003, “Interfacial heat flow in carbon nanotube suspensions,” Nat. Mats., 2, pp., 731-734. [6] Pope, M., and Swenberg, C. E., “Electronic Proceses in Organic Crystals and Polymers,” 2nd Ed., Oxford University Press, New York, pp. 337-340, 1999. [19] Liu, C. H., Hunag, H., Wu, Y., and Fan, S. S., 2004, “Thermal conductivity improvement of silicone elastomer with carbon nanotube loading,” Appl. Phys. Lett., 84 (21), pp. 42484250. [7] Peumans, P., 2003, “Small Molecular Weight Organic ThinFilm Photodetectors and Solar Cells,” J. App. Phy., 97(3), pp. 3693-3723. [20] Nan, C. W., Shi, Z., and Lin, 2003, “A simple model for thermal conductivity of carbon nanotube-based composites,” Chem. Phys. Lett., 375, pp. 666-669. [8] Tamura, T., Ogata, K., Takahashi, M., and Suzuki, K., Yamaguchi, H., and Todoroki, S. “Crsystal growth of Laser Annealed Polycrystalline Silicon as a function of Hydrogen Content of Precursors,” Proc. of Mat. Res. Soc. Symp, vol. 621, pp. Q9.5.1-Q9.5.6, 2000. [21] Lusti, H. R., and Gusev, A. A., 2004, “Finite element predictions for the thermoelastic properties of nanotube reinforced polymers,” Modeling Simul. Mater. Sci. Eng., 12, pp. 107-119. [9] Ucjikoga, S., 2002, “Low-temperature polycrystalline silicon thin-film transistor technologies for system-on-glass displays,” MRS Bull, 27, pp. 881-886. [22] Yang, R., and Chen, G., 2004, “Thermal Conductivity modeling of periodic two-dimensional nanocomposites,” Phy. Rev. B, 69, pp. 195316. [10] Duan, X., Niu, C., Sahi, V., Chen, J., Parce, J. W., Empedocles, S., and Goldman, J. L., 2003, “High-performance thin-film transistors using semiconductor nanowires and nanoribbons,” Nature, 425, pp. 274-278. [23] Morgan, V. T., 1975, “The overall convective heat transfer from smooth circular cylinders,” Advances in Heat Transfer, 11, pp. 199 – 264, 1975. [11] Snow, E. S., Novak, J. P., Campbell, P.M. and Park, D., 2003, “Random networks of carbon nanotubes as an electronic material,” Appl. Phys. Lett., 82(13), pp. 2145-2147. [24] Patankar, S.V., Numerical Heat Transfer and Fluid Flow, Hemisphere, New York, 1980. [12] Menard, E., Lee, K. J., Khang, D. Y., Nuzzo, R. G., and Rogers, J. A., 2004, “ A printable form of silicone for highperformance thin-film transistors on plastic substrates,” Appl. Phys. Lett., 84(26), pp. 5398-5400. [25] Kundert, K. S., 1988, “Sparse User's Guide”, Department of Electrical Engineering and Computer Sciences, University of California, Berkeley, CA, USA. [13] Kim, P., Shi, L., Majumdar, A., and McEuen, P. L., 2002, “Mesoscopic thermal transport and energy dissipation in carbon nanotubes,” Physica B, 323, pp. 67-70. [26] Lobb, C. J., and Frank, D. J., 1984, “Percolative conduction and the Alexander-Orbach conjecture in two dimensions,” Phy. Rev. B, 30 (7), pp. 4090-4092. [14] Kim, P., Shi, L., Majumdar, A., and McEuen, P.L , 2001, “Thermal transport measurements of individual multiwalled nanotubes,” Physical Review Letters, 87, no. 21. [27] Frank, D. J., and Lobb, C. J., 1988, “Highly efficient algorithm for percolative transport studies in two dimensions,” Phy. Rev. B, 37 (1), pp. 302-306. 9 Copyright © 2005 by ASME