JOURNAL OF APPLIED PHYSICS 102, 073720 共2007兲

Self-consistent ensemble Monte Carlo simulations show terahertz oscillations in single-walled carbon nanotubes Akin Akturk,a兲 Neil Goldsman, and Gary Pennington Electrical and Computer Engineering, University of Maryland, College Park, Maryland 20742, USA

共Received 1 February 2007; accepted 17 August 2007; published online 15 October 2007兲 We investigate electrical transient and stationary transport properties of semiconducting single-walled zigzag carbon nanotubes 共CNTs兲, using a transient ensemble Monte Carlo 共MC兲 simulator that self-consistently solves the semiclassical transport and Poisson equations. We developed the ensemble MC simulator to obtain time and space dependencies of the CNT electron concentration, velocity, and current profiles self-consistently with electrical potential distribution on the tube. Our calculated MC results indicate that self-induced terahertz CNT current oscillations on the tube and at the contacts emerge under several direct current biases. We associate these terahertz CNT oscillations with intersubband scatterings that cause the transfer of electrons from the first subband to the second, intrasubband scatterings and the nonlinear dispersion curves of each subband. The slow-moving electrons in the second subband bunch together locally on the tube, whereas the fast-moving first subband electrons move beyond the bunch and leave a relatively positive charged region behind. Also, intrasubband scatterings and subband curves give rise to low and high density electron regions by creating dispersion. These relatively low and high density electron regions create a charge dipole that then perturbs the electrical potential, resulting in a propagating domain, and thus current oscillations at tens of terahertz. After we investigate the physics of these calculated terahertz oscillations, we propose methods to modulate and shift the main oscillation frequency by varying the applied bias, tube length, or the diameter. © 2007 American Institute of Physics. 关DOI: 10.1063/1.2794690兴 I. INTRODUCTION

The electrical properties resulting from carbon nanotube 共CNT兲 atomic structure have inspired numerous theoretical and experimental investigations of quasi-one-dimensional transport in CNTs. These investigations include studies of low- and high-field mobilities, as well as phonon scatterings.1–15 As part of these efforts, we built on our previous work1–4 and developed a Monte Carlo 共MC兲 methodology to investigate space and time dependent electron transport self-consistently with built-in electric fields. The selfconsistent solution of the semiclassical transport and Poisson equations on the tube provide potential, electron current, velocity, and concentration profiles spatially along the tube at different time instances. Using these profiles, we then investigate the stationary and transient electron transport characteristics in a CNT, and explore the bias, length and diameter dependencies of these characteristics. Of particular interest is that the ensemble MC calculations predict the possibility of current oscillations at tens of terahertz and nanometers in time and space under several applied direct current 共dc兲 step biases, inside and at the contacts of the CNT. The temporal and spatial current oscillations are a result of propagating charge dipoles on the tube. A dipole initially forms by spatial charge bunching due to intersubband scatterings, intrasubband scatterings, and dispersion relations of the subbands. It then drifts along the tube, superposed onto the external field. Further, intersubband interactions cause multisubband scatterings, especially between the first two subbands for the biases we applied. As an a兲

Electronic mail: [email protected]

0021-8979/2007/102共7兲/073720/7/$23.00

electron scatters from the first to the second subband, it slows down due to two reasons: Effective mass in the second subband is higher, and it is more likely to scatter to the energy minimum of the second subband due to large densities of states around the subband minimum. Also, intrasubband scatterings may cause charge bunching due to favored scatterings to lower energies, especially to near subband minimum where densities of states are large. In addition, the dispersion resulting from the nonparabolic subband structure may contribute to spatial variations in electron density. To investigate electron transport in CNTs of several lengths and diameters, and under different biases, we first determine CNT electrical parameters to be used in the ensemble MC: CNT electron and phonon dispersion curves. Next, we show details of obtaining CNT electron-phonon scatterings using Fermi’s Golden Rule, the deformation potential approximation, and the CNT dispersion curves. We then present our ensemble transient MC methodology, followed by time and space dependent constant field ensemble MC results where built-in fields are taken as equal to the applied field. We later solve the Poisson equation to update local fields, and present temporal and spatial transient MC calculations, especially terahertz current and velocity oscillations. We also examine the length, bias, and diameter dependencies of these oscillations. II. CNT MONTE CARLO METHODS A. Details of obtaining CNT electron–phonon scattering rates

To calculate the electron–phonon scattering rate for any electron energy–momentum combination, we employ Fer102, 073720-1

© 2007 American Institute of Physics

Downloaded 13 Jun 2009 to 128.220.8.25. Redistribution subject to AIP license or copyright; see http://jap.aip.org/jap/copyright.jsp

073720-2

J. Appl. Phys. 102, 073720 共2007兲

Akturk, Goldsman, and Pennington

mi’s Golden Rule and the deformation potential approach. Using Fermi’s Golden Rule, electron scattering rate from momentum k to k⬘ can be written as follows: 2␲ 2 兩H兩 ␦关E共k⬘兲 − E共k兲 ± Ep共q兲兴. ⌫共k,k⬘兲 = ប

共1兲

Here 兩H兩2, which is calculated using the deformation potential, estimates the coupling between the states k and k⬘. For a semiconducting single-walled zigzag CNT with fundamental indices of 共n , m = 0兲, this approach gives rise to the following expression for the CNT electron-phonon scattering rate from k in subband ␤ 共k = k , ␤兲 to k⬘ in subband ␤⬘ 共k⬘ = k⬘ , ␤⬘兲 via a phonon 共q = q , ␩兲 with momentum q 共q = k⬘ − k兲 in subband ␩ 共␩ = ␤⬘ − ␤ ± 2n or ␤⬘ − ␤兲1,2,14,15:



冉 冏 冏冊

E p共q, ␩兲 = E p0共␩兲 + 共បvs␪兲 兩q兩 − ␭

2 ¯⌫共k兲 = L 兺 ប␲D 2␲ ␤ L␳





=␥



冉冑 冊 冉 冊

1 + 4 cos

共4兲

.







␲/T

−␲/T

Q2 E p共k⬘ − k, ␤⬘ − ␤兲

1 1 + exp关E p共k⬘ − k, ␤⬘ − ␤兲/Eth兴 − 1 0



⫻␦共E共k⬘, ␤⬘兲 − E共k, ␤兲 ± E p共k⬘ − k, ␤⬘ − ␤兲兲dk⬘ . 共2兲

The newly introduced variables are the deformation potential D 关9 eV 共Refs. 1 and 2兲兴, coefficient Q 关2␲ / a for optical phonons 共a = 2.46 Å兲, and 冑q2 + 共2␩ / d兲2 for acoustic phonons 共d is the tube diameter equal to na / ␲兲兴, tube length L, room temperature thermal energy Eth 共25.8 meV兲, and the CNT linear mass density ␳ 共␳ = 4nM / AT, where M = 12 g is the molar weight of carbon, T = 4.26 Å is the length of CNT’s translational vector, and A is the Avogadro’s number兲. In addition, the term in square brackets is the Bose–Einstein phonon occupation number at equilibrium, where the addition of zero or one corresponds to absorption or emission. To obtain the CNT electron dispersion relations to be used in the previous delta function, we apply zone folding methods to the tight binding graphene dispersion curve,9 written as E共kx,ky兲

d

Here E p0共␩兲 is the phonon energy at zero momentum; ␪ and ␭ are step functions; and vs is the longitudinal sound velocity of graphene, taken as 200 Å / ps. To find the total scattering rate from k in subband ␤, with corresponding energy E共k兲 = E共k , ␤兲, we integrate the scattering rate formula in Eq. 共2兲 over k⬘ for each subband ␤ ⬘:

1 1 ␲បD2Q2 + ␦关E共k⬘兲 ⌫共k,k⬘兲 = L␳E p共q兲 exp共E p共q兲/Eth兲 − 1 0 − E共k兲 ± E p共q兲兴.



冉 冊

3akx aky aky cos + 4 cos2 , 2 2 2 共3兲

where ␥ is the nearest-neighbor ␲-hopping integral taken as 3 eV,1,2 and kx and ky are the electron wave vectors. As zigzag tubes are wrapped such that kx is along the tube axis, we substitute kx by k or kz, and ky by 2␤ / d, taking the confinement around the circumference into consideration such that ky␲d = ␤2␲. This gives the zigzag CNTs’ dispersion relations as a function of wave vector along the tube axis k, and wave vector index around the circumference ␤. We also note that 冑3a = 4.26 Å is the length of the zigzag tube’s translational vector T. In addition, we include the lowest three subbands for this work. As cosine is an even function of ␤ and k, we have two valleys composed of three subbands with their minima at k = 0. The phonon dispersion curves in Eq. 共2兲 are determined using zone folded graphene phonon spectra, around the ⌫ and K points using piecewise linear lines,2 as a function of phonon wave vector index around the tube circumference ␩, and the phonon wave vector along the tube axis q:

共5兲 We then make a change of variables from k⬘ to E共k⬘ , ␤⬘兲 to evaluate the delta function: 2 ¯⌫共k兲 = បD 兺 2␳ ␤









Emax

Emin

Q2 E p共k⬘ − k, ␤⬘ − ␤兲

1 1 + exp关E p共k⬘ − k, ␤⬘ − ␤兲/Eth兴 − 1 0



⫻␦共E共k⬘, ␤⬘兲 − E共k, ␤兲 ± E p共k⬘ − k, ␤⬘ − ␤兲兲 ⫻





E共k⬘, ␤⬘兲 dE共k⬘, ␤⬘兲. ␥2T sin共Tk⬘/2兲cos共␲␤⬘/n兲

共6兲

Here the term between the delta function and dE共k⬘ , ␤⬘兲 is the density of states ⳵k⬘ / ⳵E共k⬘ , ␤⬘兲, Emin is the energy at subband minimum, and Emax is the subband minimum of the fourth subband, because we only include the first three subbands in our calculations. As in the quasi-one-dimensional CNT tube calculations we include two valleys with three subbands, at most 24 共k⬘ , ␤⬘兲’s or E共k⬘兲’s satisfy the delta function, and therefore contribute to the integral. Specifically, these 24 potential scattering mechanisms correspond to inter- and intravalley 共⫻2兲, acoustic and optical 共⫻2兲, emissions and absorptions 共⫻2兲 for the lowest three subbands 共⫻3兲. To find the possible scattering events, we solve for k⬘ or q 共q = k⬘ − k兲 that satisfy the energy and momentum conservations, for each subband ␤⬘: E共k⬘, ␤⬘兲 − E共k, ␤兲 ± E p共q,n兲 = 0, or







1 + 4 cos



冊 冉 冊 冉 冏 冏冊冊

共7兲

冉 冊

T共k + q兲 ␲␤⬘ ␲␤⬘ cos + 4 cos2 2 n n

± E p0共␩兲 + បvs␪ 兩q兩 − ␭

␩ d

= E共k, ␤兲.

共8兲

Previously, we know the initial electron energy E共k , ␤兲, momentum k, and subband ␤. For each of the three subbands 共␤⬘兲 in two valleys assuming emission or absorption 共± in

Downloaded 13 Jun 2009 to 128.220.8.25. Redistribution subject to AIP license or copyright; see http://jap.aip.org/jap/copyright.jsp

073720-3

J. Appl. Phys. 102, 073720 共2007兲

Akturk, Goldsman, and Pennington

front of the phonon energy function兲, we solve this equation for q using the phonon energy dispersion curves for different subbands ␩ 关␩ = ␤⬘ − ␤共±2n兲兴. If a phonon momentum q gives the same energy as the right-hand side of the previous equation, we have a solution. Then the scattering rate to k + q , ␤⬘ is evaluated using the following formula if the phonon branch is acoustic: ⌫共k,k + q兲 =

បD2关q2 + 4共␤⬘ − ␤兲2/d2兴 2␳E p共q, ␤⬘ − ␤兲 ⫻ ⫻

冋 冏

1 1 + exp关E p共q, ␤⬘ − ␤兲/Eth兴 − 1 0





E共k + q, ␤⬘兲 , ␥2T sin关T共k + q兲/2兴cos共␲␤⬘/n兲

共9兲

or the formula below if it is optical: ⌫共k,k + q兲 =

2ប␲2D2 a2␳E p共q, ␤⬘ − ␤兲 ⫻ ⫻

冋 冏

1 1 + exp关E p共q, ␤⬘ − ␤兲/Eth兴 − 1 0





E共k + q, ␤⬘兲 . ␥ T sin关T共k + q兲/2兴cos共␲␤⬘/n兲 2

共10兲

For example, the intrasubband scattering rate corresponding to an acoustic phonon that has a linear dispersion relation starting from zero, Ep共q , 0兲 = បvs兩q兩, is ⌫共k,k + q兲 =



1 D2兩q兩 1 + 2␳vs exp共បvs兩q兩/Eth兲 − 1 0 ⫻







E共k, ␤兲 ⫿ បvs兩q兩 . ␥ T sin关T共k + q兲/2兴cos共␲␤/n兲 2

共11兲

We note that for all other momenta that do not satisfy Eq. 共8兲, the associated scattering rate is zero due to the delta function in Eq. 共5兲. Using Eqs. 共9兲 and 共10兲, and the phonon momenta that satisfy the energy and momentum conservation rules, the total scattering rate from E共k , ␤兲 can be written as follows: ¯⌫共k兲 =



q=q,␩

⌫共k,k + q兲.

共12兲

Figure 1 shows the electron-phonon scattering rates from the lowest three n = 10 CNT subbands to its lowest three subbands up to the energy minimum of the fourth subband. Here the scattering rates are calculated for positive momenta. We then obtained the associated energy for those momenta, and plot the energy versus scattering rate curves. We have scattering peaks due to the large densities of states at subband minima. This results in high scattering rates for phonon emissions or absorptions that enable electron transfers to subband minima.

FIG. 1. 共Color online兲 Electron–phonon scattering rates from the lowest three subbands to the lowest three subbands of the n = 10 CNT up to the energy minimum of the fourth subband.

acteristics using the MC simulator, we start by configuring the steady-state distribution for initial conditions which we take as the tube under zero applied bias. We also must determine magnitude of the discrete time steps to resolve the numerical solution. To resolve the flight times between possible scattering events for electrons with any energy, we correlate the time step to the inverse of the maximum scattering rate. Such a time step considering the scattering rate can be at most a few femtoseconds, as can be seen in Fig. 1 where the peak scattering rate approaches 1015 s−1. We also need to continually update the electric field because the charge distribution inside the tube changes. The time step for updating the electric field should be comparable to the average response time of the electrostatic potential to a change in concentration. To determine the relaxation time for the local electric field, we first calculate the dielectric relaxation length for the CNT using the Poisson equation, as follows:

LD =



Eth⑀CNT , e 2␳ c

共13兲

where e is the electronic charge; ⑀CNT is the dielectric constant14,16 of the n = 10 CNT 共9.7⑀0兲; and ␳c is the net charge density. Here the linear charge density is roughly 10−4 Å−1. 共Under intrinsic conditions, linear charge density is about 10−5 Å−1.兲 Next, to find the potential relaxation time, we take the average electron velocity 共vave兲 to be approximately 107 cm/ s, based on our earlier steady-state calculations. Finally, this gives an approximate relaxation time of a few femtoseconds using t = LD / vave. Therefore, we set the time step ⌬t for solving the Poisson equation to 2 fs. Once we decide on the time steps, we discretize the momentum space using 2000 points from −␲ / T to ␲ / T for each of the three subbands 关⌬k = 共␲ / T兲 ⫻ 10−3兴. We calculate the density of each electron associated with a momentum k and subband ␤, using the following formula:

B. Ensemble transient Monte Carlo methodology

To characterize electron transport and potential on a CNT under different bias conditions, we developed a transient ensemble MC simulator. To obtain CNT transport char-

N共E共k, ␤兲兲 = f共E共k, ␤兲兲g共E共k, ␤兲兲⌬E共k, ␤兲,

共14兲

where f共E共k , ␤兲兲 is the Fermi–Dirac function,

Downloaded 13 Jun 2009 to 128.220.8.25. Redistribution subject to AIP license or copyright; see http://jap.aip.org/jap/copyright.jsp

073720-4

J. Appl. Phys. 102, 073720 共2007兲

Akturk, Goldsman, and Pennington

f共E共k, ␤兲兲 =

1 , 1 + exp关共E共k, ␤兲 − EF兲/Eth兴

共15兲

g共E共k , ␤兲兲 is the density of states introduced in Eq. 共6兲, ⌬E共k , ␤兲 is the corresponding energy step, and EF is the Fermi energy that is zero for an intrinsic tube. Furthermore, we here use 2000 momentum steps to resolve the exponential electron distribution. However, we ignore the electrons whose contributions are insignificant to the overall electron concentration. Next, to obtain the initial distribution on the tube at time t = 0, starting from the cathode 共anode兲 we distribute each electron with a positive 共negative兲 momentum k along the tube axis z every l distance apart, l共E共k, ␤兲兲 =





␥2T sin关Tk/2兴cos共␲␤/n兲 ⌬t បE共k, ␤兲

共16兲

using the electron velocity v共k , ␤兲 written in parentheses in Eq. 共16兲 共v共k , ␤兲 = 共1 / ប兲关⳵E共k , ␤兲 / ⳵k兴兲, with a weight17 equal to: ¯ 共k, ␤,z = il,t = 0兲 = N共E共k, ␤兲兲l共E共k, ␤兲兲, N i = 0,1,2, . . . ,c





L , l共E共k, ␤兲兲

pared to those theoretical1,2,12,13,18,19 and experimental5–7,10,11 data show agreement. We also have similar average velocities to those published previously.1,2,12,13,18,19 A. Time and space dependent constant field ensemble Monte Carlo

We next developed an ensemble MC simulator that resolves the semiclassical transport under a constant externally applied field. In the constant field ensemble MC simulator, we take the local electric fields as equal to the applied external field. 共Later in the following subsection, we selfconsistently solve for the built-in fields as well.兲 Specifically, we instantaneously change the applied field from 0 to 0.9 V/200 nm, on a 200 nm long single-walled zigzag n = 10 CNT, and then set the built-in fields equal to 0.9 V/200 nm for t ⬎ 0. In addition, we take the CNT as N doped with a constant Fermi level at 35% of the band gap 共EG for n = 10 is 1.06 eV兲 above the middle of the band gap. We calculate the total CNT electron concentration as a function of time t and space z evaluating first the double sum: z+⌬l

共17兲

where c共L / l共E共k , ␤兲兲兲 is L / l共E共k , ␤兲兲 rounded to the closest integer from below. We here note that the summation over a ¯ 共k , ␤ , z , t兲 for a distance l, around a point z0 of the weights N specific momentum k and subband ␤ divided by the distance l, gives the electron concentration associated with k and ␤ around the vicinity of that point zo. After we generate the zero bias profile, we apply a voltage, and calculate the new electrical potential profile on the tube using the electron distribution. To obtain the selfconsistent electrical potential on a uniform space mesh, we employ a tridiagonal matrix solver. We then let CNT electrons drift for ⌬t seconds in conjunction with the calculated built-in fields. At the end of their flights, we check if any of them probabilistically scatters with acoustic or optical phonons. For scattered electrons, we calculate new energies and momenta using the energy and momentum conservation laws. Next, we update the boundary electrons for new injections. We then recalculate the potential using the Poisson solver, and resolve the transport for the next ⌬t seconds. We repeat this for a length of time that is long enough for the important physical transport phenomena to emerge. In addition, this process is repeated for each electron on the tube simultaneously until it exits the tube from one end. Furthermore, we here start ramping the voltage at one side from zero to VA in 1 ps using uniform steps. We then keep the potential of this side fixed at VA after 1 ps.

III. ENSEMBLE MONTE CARLO RESULTS

We first determined the low-field mobility for infinitely long intrinsic semiconducting single-walled zigzag CNTs of varying diameters. Our calculated low field mobilities com-

兺 兺 N¯共k, ␤,z0,t兲,

共18兲

zo=z−⌬l k,␤

¯ 共k , ␤ , z , t兲 is the weight of an electron at location z where N 0 0 and time t, with the associated momentum k and subband ␤. We next divide the total number of electrons in 共z − ⌬l , z + ⌬l兲 by the length of this space interval 2⌬l to find the CNT electron concentration per length. Then we calculate the CNT current as a function of time and space using the following: z+⌬l

e I共z,t兲 =

兺 兺 N¯共k, ␤,z0,t兲v共k, ␤兲

z0=z−⌬l k,␤

2⌬l

,

共19兲

where e is the electronic charge, and v共k , ␤兲 is the average electron velocity around k in subband ␤. Using the CNT current and electron concentration, we later obtain average CNT velocity as a function of time and space after dividing the CNT current by the product of the CNT electron concentration and the electronic charge. Figure 2 共inset兲 shows the average CNT electron velocity 共concentration兲 as a function of tube location for an applied field of 45 kV/cm on a 200 nm long n = 10 CNT. To find the time averaged characteristics, we integrated the CNT velocity and concentration from 1 ps 共long after the steadystate is reached兲 to 10 ps by summing up the sampled values every 2⌬t seconds 共4 fs兲. We then normalized these by the number of samples. Further, we have here space dependent oscillations, with a period equal to roughly 40 nm. We attribute these space dependent oscillations to electron-phonon scatterings in the first subband. More specifically, an electron would gain 200 or 160 meV after a collision-free flight of approximately 44 or 36 nm under an applied field of 0.9 V/200 nm. These energies above the subband minima are enough to enable phonon emissions back to the minima. The electrons entering the tube with initial energy-momentum

Downloaded 13 Jun 2009 to 128.220.8.25. Redistribution subject to AIP license or copyright; see http://jap.aip.org/jap/copyright.jsp

073720-5

Akturk, Goldsman, and Pennington

J. Appl. Phys. 102, 073720 共2007兲

FIG. 2. 共Color online兲 The average CNT electron velocity on a 200 nm long CNT under a constant field of 45 kV/cm. 共Inset兲 Corresponding CNT electron concentration profile.

values near the first subband minimum 共initial energymomentum values of the majority of electrons entering the tube satisfy this due to the Fermi–Dirac statistics and the density of states兲 tend to scatter back to the subband minimum by these phonon emissions, after collision free flights of about 36 or 44 nm. These high scattering rates are related to the peaking of the density of states at the subband minimum. As a significant portion of the electrons scatter back to the first subband minimum after about 40 nm, and the associated velocities are low near the minimum, the electron concentration relatively increases around this location. This first velocity oscillation in space is followed by similar oscillations with smaller amplitudes, because the likelihood of escaping these emissions rises as electrons travel the length of the tube. Further, to investigate whether these oscillations will cause current oscillations in time, we developed a time and space dependent ensemble MC code described next. B. Time and space dependent transient ensemble Monte Carlo

To investigate whether the aforementioned spacedependent oscillations will give rise to oscillations in time, we included the Poisson equation in our MC simulator. The solution of the Poisson equation provides details of the local electric field in accordance with the semiclassical transport. We simulated a 200 nm long single-walled zigzag tube that is doped as described previously. We started with 0 applied bias condition, and then ramped the voltage on one side from 0 to 0.9 V linearly in 1 ps. The terminal voltages of 0 and 0.9 V were then kept unchanged for t ⬎ 1 ps. Figure 3共a兲 shows our calculated CNT current oscillations in time and space. The temporal progress of the current oscillations is also plotted in Fig. 3共b兲, which are slices taken from Fig. 3共a兲 at four time instances: 9.5, 9.508, 9.516, and 9.524 ps. Figure 3共b兲 shows that the minimum of the current travels from near z = 139 nm at t = 9.5 ps to z = 131 nm at t = 9.508 ps, then to z = 123 nm at t = 9.516 ps, and next to z = 112 nm at t = 9.524 ps. This indicates an approximate av-

FIG. 3. 共Color online兲 共a兲 CNT electron current as a function of time 共between 9 and 10 ps兲 and space 共tube location兲. 共b兲 CNT electron current as a function of space at four time instances: 9.5, 9.508, 9.516, and 9.524 ps. 共c兲 CNT electron current power spectrum, for tube points that are about 2 nm apart. 共CNT is 200 nm long; applied bias is 0.9 V.兲

erage envelope velocity of 108 cm/ s. As the average wavelength of the current is approximately 80 nm as shown in Fig. 3共b兲, the corresponding oscillation frequency is roughly 12.5 THz. Next, to get a better estimate of the oscillation frequency, we take the power spectrum of the current in time at tube points about 2 nm apart. The peak oscillation frequency of these power spectrum curves plotted in Fig. 3共c兲 is about 15.0 THz. 共The first peak corresponds to the dc component of the current, which is approximately 3.72 ⫻ 10−8 A.兲 We also have weaker oscillations at the harmonics; two and three times the main frequency. We show the CNT electron concentration and average velocity in Figs. 4共a兲 and 4共b兲. We note that the CNT electron concentration and average velocity are out of phase,

Downloaded 13 Jun 2009 to 128.220.8.25. Redistribution subject to AIP license or copyright; see http://jap.aip.org/jap/copyright.jsp

073720-6

Akturk, Goldsman, and Pennington

J. Appl. Phys. 102, 073720 共2007兲

first subband as in the constant field case described earlier, nonparabolicity of the subbands, or intersubband scatterings.20 More specifically, as electrons enter the tube from the ground side, they gain high velocities leaving behind a region with relatively low number of electrons. This gives rise to a potential increase on this side, enabling electrons to gain enough energies to scatter from the first subband to the second. 共The energy difference between the lowest two subbands is 0.63 eV for an n = 10 CNT.兲 Once they scatter to the second subband, they have higher effective masses and lower velocities. In addition, due to the high density of states at the subbands’ minima, they tend to scatter to the second subband’s minimum, where the average velocities are low. Thus, as electrons scatter from the first subband to the second, they pile up locally as can be seen in Figs. 5共a兲 and 5共b兲. These electrons later scatter back to the first subband, attaining higher velocities instantaneously. 共Fewer electrons populate the third subband as shown in Fig. 5共c兲, diminishing its effect on the overall electron transport.兲 Also, intrasubband scatterings and subband structures create spatially dependent electron velocities. 共However, subband structures alone do not give rise to oscillations, when electron-phonon scatterings are disabled.兲 Due to the large differences in electron velocities in adjacent regions, a charge dipole forms. This dipole then modifies the local electrostatic potential, and propagates along the tube giving rise to current oscillations.

C. CNT frequency modulation

FIG. 4. 共Color online兲 CNT electron 共a兲 concentration, 共b兲 velocity and 共c兲 potential as a function of time 共between 9 and 10 ps兲 and space 共tube location兲. 共CNT is 200 nm long; applied bias is 0.9 V.兲

thus creating current oscillations. Particularly, relatively low and high density electron regions due to fast and slow moving electrons give rise to charge dipoles. The charge dipole then modifies the local electric fields, as shown in Fig. 4共c兲. This results in dipole propagation along the length of the tube, and thus current oscillations. CNT propagating dipoles may emerge due to three different mechanisms: Electron-phonon scatterings within the

We next investigate bias, length, and index 共or diameter兲 dependencies of the CNT main oscillation frequency. We first increased the dc bias on the 100 nm long n = 10 CNT. As the dc bias is raised from 0.9 to 1.9 V, the oscillation frequency changed between 17.1 and 14.0 THz. We also simulated the CNT with higher voltages of 2.0 and 2.1 V. In these cases, electrons gained energies higher than the minimum of the fourth subband, and therefore some were deleted from the simulation domain. The remaining electrons went through a second intersubband scattering between the first and the second subbands as they moved along the tube, lowering the overall oscillation frequency by approximately half. We then examined the effects of the CNT length on the oscillation frequency. We changed the length from 50 to 200 nm, while fixing the applied dc bias at 0.9 V. As the tube length L increased from 50 to 200 nm, current oscillations remained. However, the oscillation frequency varied periodically between 17.9 THz 共L = 50 nm兲 and 15.0 THz 共L = 200 nm兲. Last, we investigated how the oscillation frequency was affected using different diameter, or fundamental index, tubes. We simulated three 100 nm long tubes, with fundamental indices of 10, 13, and 16, and diameters of roughly 0.8, 10.3, and 12.7 nm, respectively. The oscillation frequency, under a 0.9 V dc bias, changed approximately linearly as we increased the tube index 共or the tube diameter兲 from about 17.1 THz for n = 10, to 23.2 THz for n = 13, and then to 27.3 THz for n = 16. This indicates that using differ-

Downloaded 13 Jun 2009 to 128.220.8.25. Redistribution subject to AIP license or copyright; see http://jap.aip.org/jap/copyright.jsp

073720-7

J. Appl. Phys. 102, 073720 共2007兲

Akturk, Goldsman, and Pennington

IV. CONCLUSION

We report current oscillations in single-walled zigzag semiconducting CNTs under several dc biases. These current oscillations are as high as tens of terahertz that cover an important band, but generally not fully employed due to physical and material-dependent constraints, for radio frequency communication. Further, we attribute these current oscillations to inter- and intrasubband electron–phonon scatterings, and nonparabolicity of the CNT subband structure. Specifically, electrons first scatter from the first to the second subband. This slows them down due to higher effective masses and lower velocities in the second subband. These electrons then scatter back to the first subband, instantaneously gaining higher velocities. This, and electron bunching due to intrasubband scatterings and electron dispersion curves, in turn create a propagating charge dipole, and current oscillations. Additionally, we show that the main frequency of these oscillations may be modulated by the dc bias, tube length or the tube diameter. This may open paradigms in oscillator designs for terahertz radio frequency circuits.

ACKNOWLEDGMENTS

The authors would like to thank Professor Michael Fuhrer from the University of Maryland College Park for invaluable discussions, and Alma Wickenden from the Army Research Laboratory for support. G. Pennington and N. Goldsman, Phys. Rev. B 68, 045426 共2003兲. G. Pennington and N. Goldsman IEICE Trans. Electron. E86, 372 共2003兲. 3 A. Akturk, G. Pennington, and N. Goldsman, IEEE Trans. Electron Devices 52, 577 共2005兲. 4 G. Pennington, N. Goldsman, A. Akturk, and A. Wickenden, Appl. Phys. Lett. 90, 062110 共2007兲. 5 T. Durkop, B. M. Kim, and M. S. Fuhrer, J. Phys.: Condens. Matter 16, R553 共2004兲. 6 T. Durkop, S. A. Getty, E. Cobas, and M. S. Fuhrer, Nano Lett. 4, 35 共2004兲. 7 Y.-F. Chen and M. S. Fuhrer, Phys. Rev. Lett. 95, 236803 共2005兲. 8 M. S. Dresselhaus, G. Dresselhaus, and A. Jorio, Annu. Rev. Mater. Res. 34, 247 共2004兲. 9 R. Saito, M. S. Dresselhaus, and G. Dresselhaus, Physical Properties of Carbon Nanotubes 共Imperial College, London, 1998兲. 10 X. Zhou, J.-Y. Park, S. Huang, J. Liu, and P. L. McEuen, Phys. Rev. Lett. 95, 146805 共2005兲. 11 S. Li, Z. Yu, C. Rutherglen, and P. J. Burke, Nano Lett. 4, 2003 共2004兲. 12 V. Perebeinos, J. Tersoff, and P. Avouris, Nano Lett. 6, 205 共2006兲. 13 V. Perebeinos, J. Tersoff, and P. Avouris, Phys. Rev. Lett. 94, 086802 共2005兲. 14 G. Pennington and N. Goldsman, Phys. Rev. B 71, 205318 共2005兲. 15 A. Akturk, G. Pennington, N. Goldsman, and A. Wickenden, Phys. Rev. Lett. 98, 166803 共2007兲. 16 L. X. Benedict, S. G. Louie, and M. L. Cohen, Phys. Rev. B 52, 8541 共1995兲. 17 T. Wang, K. Hess, and G. J. Iafrate, J. Appl. Phys. 58, 857 共1985兲. 18 A. Verma, M. Z. Kauser, and P. P. Ruden, J. Appl. Phys. 97, 114319 共2005兲. 19 A. Verma, M. Z. Kauser, and P. P. Ruden, Appl. Phys. Lett. 87, 123101 共2005兲. 20 P. A. Lebwohl and P. J. Price, Appl. Phys. Lett. 19, 530 共1971兲. 1 2

FIG. 5. 共Color online兲 CNT electron concentration in the 共a兲 first subband, 共b兲 second subband, and 共c兲 third subband as a function of time 共between 9 and 10 ps兲 and space 共tube location兲.

ent diameter tubes, we may be able to cover a large frequency band at THz frequencies, which is important in radio frequency communication.

Downloaded 13 Jun 2009 to 128.220.8.25. Redistribution subject to AIP license or copyright; see http://jap.aip.org/jap/copyright.jsp

Self-consistent ensemble Monte Carlo simulations show terahertz ...

propagating domain, and thus current oscillations at tens of terahertz. After we ..... FIG. 1. Color online Electron–phonon scattering rates from the lowest.

819KB Sizes 0 Downloads 158 Views

Recommend Documents

Sonification of Markov chain Monte Carlo simulations
This paper illustrates the use of sonification as a tool for monitor- ... tional visualization methods to understand the important features of Ф. When. , however ...

Monte Carlo simulations of interacting anyon chains - Semantic Scholar
Apr 8, 2010 - ... Field Laboratory, Florida State University, Tallahassee, FL 32310, USA ..... Mod. Phys. 80 (2008) 1083. [6] G. Rumer, Gцttingen Nachr. Tech.

Monte Carlo simulations of interacting anyon chains - Semantic Scholar
Apr 8, 2010 - ... Field Laboratory, Florida State University, Tallahassee, FL 32310, USA ..... Mod. Phys. 80 (2008) 1083. [6] G. Rumer, Gцttingen Nachr. Tech.

Monte Carlo simulations for a model of amphiphiles ...
Available online at www.sciencedirect.com. Physica A 319 (2003) .... many sites. The amphiphiles possess internal degrees of freedom with n different states.

Monte-Carlo Simulations of Magnetic Tunnel Junctions
its predicts correctly the general shape of the PDF extracted from the simulations results and tends to an exponential law in the weak current regime as the MTJ behavior becomes a. Poisson's process (theoretical derivation in [2]). Figure 3 shows tha

Monte Carlo Simulation
You are going to use simulation elsewhere in the .... If we use Monte Carlo simulation to price a. European ...... Do not put all of your “business logic” in your GUI.

a monte carlo study
Mar 22, 2005 - We confirm this result using simulated data for a wide range of specifications by ...... Federal Reserve Bank of Kansas City and University of Missouri. ... Clements M.P., Krolzig H.$M. (1998), lA Comparison of the Forecast ...

Sequential Monte Carlo multiple testing
Oct 13, 2011 - can be reproduced through a Galaxy Pages document at: ... Then, in Section 3, we show on both simulated and real data that this method can ...

Introduction to Monte Carlo Simulation
Crystal Ball Global Business Unit ... Simulation is the application of models to predict future outcomes ... As an experimenter increases the number of cases to.

Sequential Monte Carlo multiple testing
Oct 13, 2011 - An example of such a local analysis is the study of how the relation ... and then perform a statistical test of a null hypothesis H0 versus. ∗To whom ... resampling risk (Gandy, 2009), and prediction of P-values using. Random ...

Hamiltonian Monte Carlo for Hierarchical Models
Dec 3, 2013 - eigenvalues, which encode the direction and magnitudes of the local deviation from isotropy. data, latent mean µ set to zero, and a log-normal ...

Introduction to Monte Carlo Simulation - PDFKUL.COM
Monte Carlo Simulation Steps. • Following are the important steps for Monte Carlo simulation: 1. Deterministic model generation. 2. Input distribution identification. 3. Random number generation. 4. Analysis and decision making ..... perform output

Statistical Modeling for Monte Carlo Simulation using Hspice - CiteSeerX
To enable Monte Carlo methods, a statistical model is needed. This is a model ..... However, it is difficult to determine the correlation without a lot of statistical data. The best case .... [3] HSPICE Simulation and Analysis User Guide. March 2005.

Bayes and Big Data: The Consensus Monte Carlo ... - Semantic Scholar
Oct 31, 2013 - posterior distribution based on very large data sets. When the ... and Jordan (2011) extend the bootstrap to distributed data with the “bag of little ...

Using the Direct Simulation Monte Carlo Approach for ...
The viability of using the Direct Simulation Monte Carlo (DSMC) approach to study the blast-impact ... by computing load definition for two model geometries - a box and an 'I' shaped beam. ... On the other hand, particle methods do not make the conti

Bayes and Big Data: The Consensus Monte Carlo ... - Rob McCulloch
Oct 31, 2013 - The number of distinct configurations of xij in each domain is small. ...... within around 11,000 advertisers using a particular Google advertising.

A Non-Resampling Sequential Monte Carlo Detector for ... - IEEE Xplore
SMC methods are traditionally built on the techniques of sequential importance sampling (SIS) and resampling. In this paper, we apply the SMC methodology.

Wigner-Boltzmann Monte Carlo approach to ... - Springer Link
Aug 19, 2009 - Quantum and semiclassical approaches are compared for transistor simulation. ... The simplest way to model the statistics of a quantum ..... Heisenberg inequalities, such excitations, that we will call ..... pean Solid State Device Res

accelerated monte carlo for kullback-leibler divergence ...
When using ˜Dts a (x) with antithetical variates, errors in the odd-order terms cancel, significantly improving efficiency. 9. VARIATIONAL IMPORTANCE ...