Você está na página 1de 6

IEEE TRANSACTIONS ON ELECTRON DEVICES, VOL. 48, NO.

2, FEBRUARY 2001

279

Quantum Device-Simulation with the Density-Gradient Model on Unstructured Grids


Andreas Wettstein, Andreas Schenk, and Wolfgang Fichtner, Fellow, IEEE

AbstractWe describe an implementation of the density-gradient device equations which is simple and works in any dimension without imposing additional requirements on the mesh compared to classical simulations. It is therefore applicable to real-world device simulation with complex geometries. We use our implementation to determine the quantum mechanical effects for a MOS-diode, a MOSFET and a double-gated SOI MOSFET. The results are compared to those obtained by a 1D-SchrdingerPoisson solver. We also investigate a simplified variant of the density-gradient term and show that, while it can reproduce terminal characteristics, it does not give the correct density distribution inside the device. Index TermsBox method, density gradient method, multidimensional device simulation, quantum corrections, semiconductordevice modeling.

I. INTRODUCTION HE LENGTH scales in todays semiconductor devices are so small that quantum effects become important. Apart from tunneling through the gate insulator, the most important effect is the modification of the gate capacity as a function of the gate voltage (capacitancevoltage, or CV, characteristics), where confinement leads to a threshold voltage shift and an apparent increase of the oxide thickness. Various methods have been suggested to model these quantum effects. Among the approaches that are compatible with classical device simulators based on the drift-diffusion (or hydrodynamic) approach, the physically most accurate method is to include the Schrdinger equation into the self-consistent computation of the device characteristics [1]. As for current devices the quantization is only relevant for the direction perpendicular to the oxidesilicon interface, one restricts to a one-dimensional (1-D) treatment of the Schrdinger equation. Two-dimensional (2-D) devices are treated by making a couple of 1-D slices along the channel. Finite temperature is taken into account by computing multiple subbands and populating them according to their energy assuming MaxwellBoltzmann or Fermi distribution functions. Solving the Schrdinger equation is not only time-consuming in itself, it also leads to direct coupling of all grid points within a slice. This destroys the sparsity
Manuscript received March 27, 2000; revised July 3, 2000. The review of this paper was arranged by Editior A. H. Marshak. The authors are with the Institut fr Integrierte Systeme, Swiss Federal Institute of Technology, Zrich, CH-8092 Switzerland (e-mail: wettstein@iis.ee.ethz.ch). Publisher Item Identifier S 0018-9383(01)00767-5.

structure in the Jacobian needed for the Newton-iteration used to solve the semiconductor equations. Therefore, these nonlocal couplings are typically neglected. Another serious drawback of this method is that it is one dimensional in nature and thus realistic devices with nonplanar oxidesilicon interfaces can not be handled satisfactory. Using a 2-D Schrdinger equation could help here, but at the price of much increased numerical cost to determine the eigensolutions and of yet higher nonlocality. An additional physical problem is the choice of appropriate boundary conditions. Various simpler methods have been suggested. The probably most familiar is that by van Dort et al. [2], which expresses the quantum effect by an apparent band edge shift that is a simple function of the electric field. The model is based on the expression for the lowest eigenenergy of a particle in a triangular potential and reproduces the characteristics obtained with the Schrdinger equation quite well, and thus is widely used in practice. This model does not, however, give the correct charge distribution in the device. While in principle it can be used for 2and 3-D devices of nonrectangular geometry, it is unclear how well the model is still justified in such cases. A third alternative are density gradient (or, generalized to the level of hydrodynamics, quantum hydrodynamic models) [3][7]. In these models, an additional term is introduced in the continuity equations. This term contains higher-order derivatives of the potential or of the charge density. It thereby introduces a limited degree of additional nonlocality into the equations which accounts for the effect of quantum mechanics. The model is simple and its ability to describe CV characteristics [8] and even tunneling currents [9] is well established. To implement it, the Jacobian of the system has to be modified, but it remains sparse. Furthermore, the model is multidimensional by construction. In this paper, we describe how the density gradient model can be implemented in a multidimensional device simulator using the box method on unstructured grids. While these grids are essential to keep the number of grid points in realistic devices small, the discretization of higher order derivatives is not straightforward. The strategy we have chosen is to introduce a new variable that encapsulates the higher order derivatives, thereby making two second-order differential equations out of one fourth-order equation. We will also motivate and investigate a simplified density-gradient method, which turns out to give correct terminal characteristics, but fails to predict the correct charge distribution in the interior of the device.

00189383/01$10.00 2001 IEEE

280

IEEE TRANSACTIONS ON ELECTRON DEVICES, VOL. 48, NO. 2, FEBRUARY 2001

II. DRIFTDIFFUSION EQUATIONS WITH DENSITY GRADIENT CORRECTION In [10], it is shown that in lowest-order Born approximation, the current equation reads (1) where we already have left out time derivatives and the convective term (terms quadratic in the particle current densities ). is the particle density, the temperature (in contrast to [10], we do not consider second order momentum terms and assume to be constant), is the Boltzmann constant, the elementary the effective mass and the mobility. is the pocharge, and electrostatic tential, including both the band edges , with the carrier charge denoted potential , is given by [10] by .

Fig. 1. Potential 8, correction according to (10).

Q according to (4), their smooth sum, and Q

(2) where is the dimensionality of the mesh and a 1-D problem, we find . For

For illustration, we consider a simple example. Assume that ), and the the system is infinite, 1-D and in equilibrium ( potential has the form for for Under these assumptions, (1) simplifies to (3) where evaluates to (4) and . is a continuous function in , that is, looking only at the second term in (3), we see that the quantum correction smoothes out the jump in on a length scale ; see Fig. 1. Equation (3) is an ordinary first order differential equation can be integrated to give which for .

An obvious remedy would be to neglect the last term in (3). , which is This leads in our example to well behaved. Using this form of , the last term in (3) is by smaller than the second term. As the derivation a factor is small, neglecting the last term thus of (1) assumes that is also justified in the limit where (3) is strictly valid. Another approach to improve (3) has been described in [12], [13]. However, computing (2) is still numerically too expensive to be done in actual device simulation. Using a fast Fourier transform [11] reduces numerical complexity, but is impractical for general, unstructured grids. We therefore follow [10] and to vary slowly on a length scale of , and simassume . Furthermore, in equilibrium, plify (2) to . Hence, the last term in (1) becomes

(5) on the right hand side gives only a The explicit contribution to the continuity equation and can be omitted in lowest order quantum correction. Thus, (1) simplifies to (6) where is given by (7) The quantum effect thus shows up as an additional force term derived from of a potential-like quantity that is comprised of higher order derivatives of the classical potential. As might have abrupt jumps, the quantity above might be undefined in certain points. We therefore replace (7) by a slightly modified partial differential equation for , (8)

and are the integration constants for and , respectively. If [which violates the assumptions used to derive (1), but is common in practice], the and denominator of this equation becomes zero for a the density in the barrier diverges. Therefore, for general structures, this model can not be used unaltered (This problem does not arise for the thin barriers in [10], [11], because throughout the barrier).

where

WETTSTEIN et al.: QUANTUM DEVICE-SIMULATION WITH DENSITY-GRADIENT MODEL

281

As is , this modification, like the approximation done on (5), does not contribute in lowest-order quantum correction. For a semiconductor with multiple conduction band minima and anisotropic effective mass it is not clear which value to take for in (8). We take as the density of states mass and handle the problem by introducing the fit factor above. In order to take account of position dependent effective masses, we add a mass term contribution to the potential in the same manner it appears in the classical drift-diffusion equations

The gradient square term in (8) corresponds to the last term in (3), which we found to lead to problems under certain circumstances. It is therefore tempting to replace by the simplified form (9) For our 1-D-potential step example discussed earlier (using ) this leads to (10) is close to what we have obtained before in which for small (4), as can be seen from Fig. 1. As we will show later, while gives reasonable results for the CV characteristics, it using does not give the correct charge distributions in the channel. , therefore (8) can also In equilibrium, be written as

grid data structures designed for use with the box method, the information needed to compute these derivatives might not be available at all. Therefore, this approach is only useful on simple grid structures like tensor grids. Our approach is to introduce as a new variable, that is, the number of unknowns of the nonlinear system is increased. On the other hand, as (8) is only of second order, it can conveniently be discretized combining function values only on nearest neighbor grid points. While substituting into (6) leads to additional off-diagonal elements in the Jacobian and changes the matrix structure, our approach adds an additional block to the Jacobian, but the sparsity structure of each block still reflects the nearest-neighbor relationship of the underlying mesh. For (6), the ScharfetterGummel discretization [18] is used. Equation(8) is discretized as

(11) which are the forms often used in the literature. We will continue to use the form (8), however. This is advantageous for cases where we are not interested in transport through barriers but only in properties like CV characteristics. For these cases, (8) allows to treat the barrier as insulating and to exclude it from the solution of the continuity equation while still being able to correctly compute . III. NUMERICS While implementations of the quantum hydrodynamic model using finite element methods have been reported [14], [15], the method most popular for implementation of the device equations are the box method and f inite d ifference methods (for an introduction, see [16]). In our case, the choice of the box method was predetermined by the target simulation environment [17]. The usual approach to discretize (6) is to substitute according to (11) into this equation. The order of this differential equation is thereby increased by two. This means that to discretize it in a given grid point, not only nearest neighbor grid points need be examined, but also grid points further away. On

The indices and are used to designate grid points. The -sum is the volume runs over the nearest neighbors of grid point . of the box for grid point , is the distance of grid points and , and is the area of the box face between these two points. As boundary conditions for . (8), we assume that at the outer boundary of the device. This would corresponds to if the density-based formula (11) for was used. While this boundary condition is particularly easy to implement, it is probably not ideal at contacts where the electric field can be large, for example at metallic gates. To solve the nonlinear system of equations, we perform a Newton iteration on the entire coupled system of equations [current equations (6), (8), and Poisson equation]. For all examples discussed below, the linear systems that arise in each Newton step were solved by a direct method. To our surprise, the convergence the Newton iterations of the quantum hydrodynamic models for the examples we consider in the following section was worse than for the Schrdinger equation, even though a full Jacobian was available. Also, due to the increased number of unknowns in the nonlinear system, the time needed to obtain the eigensolutions of the 1-D-Schrdinger equation was counterbalanced. IV. EXAMPLES A. MOS diode Our first example is a one-dimensional MOS diode. We compare five different approaches: 1-D-SchrdingerPoisson solver [19], van Dort model [2], the density gradient method based on (8), the simplified density gradient method using (9), and the classical model. In all the simulations, the oxide has been treated as a true insulator and no current flow has been allowed. The upper graph in Fig. 2 shows the CV-characteristics for a MOS-diode with 4 nm gate oxide thickness, a channel doping cm and a metal gate, computed at 300 K. The of and . The graph fit parameters used were shows that the two density gradient models and the van Dort model describe the curve well, that is, the deviation from the

282

IEEE TRANSACTIONS ON ELECTRON DEVICES, VOL. 48, NO. 2, FEBRUARY 2001

Fig. 3. CV curves for MOS-diodes with 3 nm gate oxide and 2:5 1 10 cm doping (left) and 5 nm oxide thickness and 10 cm doping (right). The upper graphs are for 300 K, the lower graphs for 200 K.

Fig. 2. Top: CV characteristics for a MOS-diode with 4 nm gate oxide and 5 1 10 cm channel doping. Bottom: Electron density in the same MOS-diode at gate-back voltage V = 4 V. z is the direction perpendicular to the silicon-oxide interface which is located at z = 0. The temperature was 300 K for both graphs.

Schrdinger result is negligible compared to the difference from the classical result. For the same device, we have plotted the channel density profile at a gate-back contact voltage of 4 V in the bottom of Fig. 2. While the CV-curves are quite accurate, the density profile for the van Dort model and the simplified Density Gradient model (9) are far off the Schrdinger result. In particular, the density for the latter decreases much too fast when approaching the oxide interface. On the other hand, the density gradient model (8) reproduces the Schrdinger density well. We varied the oxide thickness, the channel doping, and the temperature to check if the CV curves can be reproduced with the same choice of and . Fig. 3 shows the results. It can be seen, that all the models work well for different device parameters without need to readjust the fit constants. B. MOSFET In Fig. 4 we show a comparison of the five models for the current-voltage characteristics of a hypothetical MOSFET with 4 nm gate oxide thickness, 300 nm gate length, substrate doping cm , an effective channel length of 200 nm and a channel width of 1 m. As channel mobility models make use of the field distribution [20] and would need model-specific recalibration of their fit parameters, we restricted to a nonrealistic, constant mobility to make interpretation of the results more straightforand ward. We used the parameters from Section IV-A,

Fig. 4. Drain current I versus gate-source voltage V for the MOSFET described in the text at 300 K and drain-source voltage V = 0:1 V. In the upper graph the subthreshold regime is plotted, the lower graph shows the linear regime. The insets show zooms of the main graphs.

. The graphs show all quantization models agree well for this example, too.

WETTSTEIN et al.: QUANTUM DEVICE-SIMULATION WITH DENSITY-GRADIENT MODEL

283

which shows a cut perpendicular to the interfaces in the middle of the channel at a gate voltage V. V. SUMMARY AND CONCLUSION We have described a multidimensional implementation of the density-gradient method which does not depend on a tensor grid to work. The main aspect of this approach is to introduce the additional variable to the nonlinear system of unknowns. The equation for can easily be discretized even on unstructured grids using the box method. We compared the results obtained to those of a full Schrdinger equation based quantization model, the van Dort model and a simplified density-gradient method. While the simplified model and the van Dort model are able to reproduce CV characteristics well, only the full density gradient method also gives the correct charge distribution in the channel. In our numerical examples we found no advantage of our implementation of the density gradient method over the 1-D Schrdinger equation approach in terms of stability and computation time for the examples where the latter is applicable. The higher flexibility regarding device geometry and mesh are therefore the primary advantages of the density gradient method over the Schrdinger method. The density-gradient method could be also advantageous for applications where the knowledge of the full Jacobian of the nonlinear system is mandatory, like, for example, small-signal analysis. ACKNOWLEDGMENT The authors thank M. G. Ancona and C. L. Gardner for useful discussion.
Fig. 6. Density profile in the middle of the channel of the SOI MOSFET at a gate voltage of 1 V. The classical density and the density for the van Dort model agree.

Fig. 5. Drain current density vs. gate voltage for the SOI MOSFET described 0:01 V, the temperature was 300 K. The in the text. The drain bias was V inset shows a zoom of the main graph.

REFERENCES
[1] F. Stern, Self-consistent results for n-type Si inversion layers, Phys. Rev. B, vol. 5, pp. 48914899, June 1972. [2] M. J. van Dort, P. H. Woerlee, and A. J. Walker, A simple model for quantization effects in heavily-doped silicon MOSFETs at inversion conditions, Solid-State Electron., vol. 37, no. 3, pp. 411414, 1994. [3] M. G. Ancona and H. F. Tiersten, Macroscopic physics of the silicon inversion layer, Phys. Rev. B, vol. 35, pp. 79597965, May 1987. [4] M. G. Ancona and G. J. Iafrate, Quantum correction to the equation of state of an electron gas in a semiconductor, Phys. Rev. B, vol. 39, no. 13, pp. 95369540, May 1989. [5] J.-R. Zhou and D. K. Ferry, Ballistic phenomena in GaAs MESFETs: Modeling with quantum moment equations, Semicond. Sci. Technol., vol. 7, pp. B546B548, Mar. 1992. [6] H. L. Grubin, T. R. Govindan, J. P. Kreskovsky, and M. A. Stroscio, Transport via the Liouville equation and moments of quantum distribution functions, Solid-State Electron., vol. 36, pp. 16971709, Dec. 1993. [7] C. L. Gardner, The quantum hydrodynamic model for semiconductor devices, SIAM J. Appl. Math., vol. 54, pp. 409427, Apr. 1994. [8] M. G. Ancona, Z. Yu, W.-C. Lee, R. W. Dutton, and P. V. Voorde, Simulation of quantum confinement effects in ultra-thin-oxide MOS structures, J. Technol. Comput. Aided Design, May 1999. [9] M. G. Ancona et al., Density-gradient analysis of tunneling in MOS structures with ultra-thin oxides, in SISPAD 99 Tech. Dig.. Kyoto, Japan, Sept. 1999, pp. 235238. [10] C. L. Gardner and C. A. Ringhofer, Smooth quantum potential for the hydrodynamic model, Phys. Rev. E, vol. 53, pp. 157167, Jan. 1996. [11] , Smooth quantum hydrodynamic model simulation of the resonant tunneling diode, VLSI Design, vol. 8, pp. 143146, 1998. , Numerical simulation of the smooth quantum hydrodynamic [12] model for semiconductor devices, Comput. Meth. Appl. Mechan. Eng., vol. 181, pp. 393401, Jan. 2000.

C. Double-Gated SOI MOSFET To investigate the limits of the models, we modeled a hypothetical double-gated SOI MOSFET with a very thin silicon layer of 5 nm and a nearly undoped channel of 50 nm length. While this is not a realistic device yet, it is useful to investigate how well the density gradient method and the van Dort model agree to the results obtained by the Schrdinger approach for a structure much different from an ordinary MOSFET. In Fig. 5 we show the drain current as a function of gate voltage (both gates are ramped together) of the device in the subthreshold regime. The van Dort model is found to agree with the classical curve here. This can be expected, because it was not intended to be applied in this case where potential is not triangular, but basically box shaped (note that at higher gate voltages the agreement is good again because the potential near the upper and lower interface becomes triangular). The density gradient and the simplified density gradient model give the same result and both underestimate the effect of quantization. We checked that the discrepancy remains if the channel length is increased to 500 nm. Thus this is not a 2-D effect. The behavior of the drain current is reflected by the density profile given in Fig. 6,

284

IEEE TRANSACTIONS ON ELECTRON DEVICES, VOL. 48, NO. 2, FEBRUARY 2001

[13] [14] [15] [16] [17] [18] [19]

[20]

, Resonant tunneling in the smooth quantum hydrodynamic model for semiconductor devices, Transport Theory Stat. Phys., vol. 29, pp. 563570, 2000. Z. Chen, A finite element method for the quantum hydrodynamic model for semiconductor devices, Comput. Math. Appl., vol. 31, pp. 1726, 1996. , Quantum hydrodynamic simulation of hysteresis in the resonant tunneling diode, J. Comput. Phys., vol. 117, pp. 274280, Mar. 1995. S. Selberherr, Analysis and Simulation of Semiconductor Devices. Berlin, Germany: Springer-Verlag, 1984. ISE Integrated Systems Engineering AG, DESSIS Ref. Manual: , 1998. D. L. Scharfetter and H. K. Gummel, Large-signal analysis of a silicon read diode oscillator, IEEE Trans. Electron Devices, vol. ED-16, pp. 6477, Jan. 1969. A. Wettstein et al., Charge carrier quantization effects in double-gated SOI MOSFETs, in ULSI Sci. Technol. 1997, H. Z. Massoud, H. Iwai, C. Claeys, and R. B. Fair, Eds. Pennington, NJ: The Electrochem. Soc., Inc., Mar. 1997, pp. 613621. M. N. Darwish et al., An improved electron and hole mobility model for general purpose device simulation, IEEE Trans. Electron Devices, vol. 44, pp. 15291537, Sept. 1997.

Andreas Schenk was born in Berlin, Germany, in 1957. He received the Dipl. in physics and the Ph.D. in theoretical physics from Humboldt University (HUB), Berlin, Germany, in 1981 and 1987, respectively. In 1987 he became a Research Assistant at the Department of Semiconductor Theory, HUB, and in 1988 he joined the R&D division of WF Berlin. From 1987 till 1991 he was working on various aspects of the physics and simulation of optoelectronic devices, especially infrared detector arrays, and the development and implementation of physical models for modeling infrared sensors. He is now with the Swiss Federal Institute of Technology, Zurich, Switzerland, as a Senior Lecturer at the Integrated Systems Laboratory. His main activities are in the development of physics-based models for simulation of submicron silicon devices. Dr. Schenk is a member of the German Physical Society..

Andreas Wettstein received the Dipl. in physics from the Universitt Karlsruhe, Germany, in 1995. Since then he has been with the Institut fr Integrierte Systeme, Swiss Federal Institute of Technology, Zrich, Switzerland.

Wolfgang Fichtner (F90) received the Dipl.Ing. degree in physics and the Ph.D. degree in electrical engineering from the Technical University of Vienna, Austria, in 1974 and 1978, respectively. From 1975 to 1978, he was an Assistant Professor in the Department of Electrical Engineering, Technical University of Vienna. From 1979 through 1985, he was with AT&T Bell Laboratories, Murray Hill, NJ. Since 1985, he has been a Professor and Head of the Integrated Systems Laboratory at the Swiss Federal Institute of Technology, Zurich, Switzerland. In 1993, he founded ISE Integrated Systems Engineering AG, a company in the field of technology CAD. Dr. Fichtner is a member of the Swiss National Academy of Engineering.

Você também pode gostar