JFETNANO

download JFETNANO

of 9

Transcript of JFETNANO

  • 8/7/2019 JFETNANO

    1/9

    IEEE TRANSACTIONS ON NANOTECHNOLOGY, VOL. 8, NO. 6, NOVEMBER 2009 749

    Junction Field Effect Transistors for NanoelectronicsJustin B. Jackson, Student Member, IEEE, Divesh Kapoor, Student Member, IEEE, and Mark S. Miller, Member, IEEE

    AbstractThe gate leakage currents of MOSFETs increase ex-ponentially with downward scaling, while the gate currents ofenhancement-mode JFETs for complementary logic decrease withscaling. In principle, a crossover point could exist below which theJFET may be the preferred device for some large-scale ICs. Thispaper first examines the crossover point with a simple scaling anal-ysis that suggests it lies in the neighborhood of a 25 nm gate length,depending on the supply voltages and the gate insulator used forthe MOSFET. Other JFETelectrical properties comparefavorablywith those of MOSFETs and exhibit similar scaling behaviors. Nu-merical simulations of a simple 25 nm gate length JFET showelectrical properties better than the conservative scaling analysisand comparable to reported 25 nm MOSFETs, with, for example,a gate current density of 12 A/cm2 at 0.7 V and a drain currentON-to-OFF ratio of 3.5 103 . A self-aligned polycrystalline silicon

    gate and some straightforward performance enhancements pro-posed for the 25 nm device may allow it essentially to stand in forthe geometrically similar 25 nm MOSFET in some circumstances.Additional device engineering outlined in this paper should furtherallow the silicon JFET to scale down to 10 nm gate lengths, wheresourcedrain tunneling becomes important. Ten-nanometer-scaleJFETs share many of the fabrication challenges of correspondingMOSFETs, and many of the well-developed concepts for MOS-FETs, such as double gatesand strain engineering, should be easilyadapted to JFETs. These devices also lend themselves to a subse-quent transition to IIIV semiconductorsfor heterostructures,performance increases, and scaling below 10 nmwithout the ox-ide interface problems MOSFETs would face. Overall, JFETs ap-pear to be of interest for the next one to two silicon technologynodes, and perhaps beyond.

    Index TermsJFETs, MOSFETs, nanotechnology.

    I. INTRODUCTION

    AS THE dimensions of MOSFETs have scaled downward

    to increase density and performance, gate leakage due to

    tunneling through thin gate oxides has become a major design

    constraint. The gate leakage current density increases approx-

    imately exponentially with decreasing oxide thickness. While

    thicker gate insulators with higher dielectric constants have mit-

    igated this problem, subsequent scaling still encounters this ex-

    ponential dependence. For complementary logic, JFETs need to

    be enhancement mode devices, turning ON with forward-biasedgate p-n junctions. Such currents, and the consequent restric-

    tions on gate voltages, have historically made these transistors

    noncompetitive for large-scale ICs. However, scaling a JFET to

    smaller dimensions can actually decrease the gate current some-

    what, because the scaled-up doping concentrations reduce the

    Manuscript received August 7, 2007; revised April 29, 2008. First publishedDecember 22, 2008; current version published November 11, 2009. The reviewof this paper was arranged by Associate Editor T. Hiramoto.

    The authors are withthe Department of Electrical and ComputerEngineering,University of Utah, Salt Lake City, UT 84112 USA (e-mail: [email protected]).

    Color versions of one or more of the figures in this paper are available onlineat http://ieeexplore.ieee.org.

    Digital Object Identifier 10.1109/TNANO.2008.2011383

    Fig. 1. Scaling models for JFET and MOSFET comparison. The dimensionsscale by > 1, and for the doping scales as 2 .

    junction saturation currents; and logic levels are decreasing into

    the range of p-n junction turn-on voltages. In principle, there

    could be a crossover point in scale below which complementary

    JFET circuitry has lower gate leakage and may become prefer-

    able. Whether the crossover point happens at a technologicallyrelevant size will depend on the details of where it occurs with

    respect to MOSFET scaling, and how well JFETs function at

    that scale and below.

    The largest tolerable transistor gate current density, which im-

    poses a minimum gate length L for MOSFETs, depends on theapplication the circuit is designed for [1]. For comparison pur-

    poses, the present paper considers gate lengths for single-gate,

    bulk devices as depicted in Fig. 1. For moderate-performance

    ICs, for example those used in desktop applications, a conser-

    vative limit on gate current density is 1 A/cm2 , though some

    authors have argued for and demonstrated a higher practical

    maximum of perhaps 100 A/cm2 [2][4]. For a MOSFET with

    a silicon dioxide gate insulator and a gate potential of 1 V,

    the leakage reaches 1 A/cm2 when the oxide thickness scales to

    about 1.7 nm [3], [5]. This silicon dioxide thickness corresponds

    to a transistor length ofL 70 nm [6]. Current production tech-nologies at the 65 nm technology node use silicon oxynitride

    gate insulators, with one example having a 35 nm gate length,

    though its 1.2 nm insulator was not scaled from the preceding

    90 nm technology node because of gate leakage [7]. An exam-

    ple of an emerging 45 nm technology takes advantage of the

    yet higher dielectric constant of a hafnia-based gate insulator,

    though its physical gate length remained at 35 nm [8].

    The scaling properties of enhancement mode JFETs and the

    consequences for complementary logic have received much less

    1536-125X/$26.00 2009 IEEE

    Authorized licensed use limited to: K.S. Institute of Technology. Downloaded on January 30, 2010 at 10:54 from IEEE Xplore. Restrictions apply.

  • 8/7/2019 JFETNANO

    2/9

    750 IEEE TRANSACTIONS ON NANOTECHNOLOGY, VOL. 8, NO. 6, NOVEMBER 2009

    attention in the literature. Electrically, JFETs behave similarly

    to corresponding MOSFETs, and they also benefit from lower

    input capacitance and lower noise. Because of these proper-

    ties, some applications have favored JFETs over MOSFETs, for

    example, as the low-noise front ends for charge sensing ampli-

    fiers. For ICs, complementary enhancement mode silicon JFETs

    have been investigated because of their inherently greater resis-tance to radiation damage [9], [10]. The scaling behaviors of the

    closely related MESFET have been studied and reported [11], as

    have complementary enhancement mode silicon MESFET tech-

    nologies [12]. However, the Schottky-contact gates have much

    larger saturation current densities than p-n junction gates, and

    an upper limit on channel doping to avoid gate tunneling cur-

    rents prevents these devices from scaling to smaller dimensions.

    For JFETs, a simple, one-sided diode model can give an initial

    gate current density estimate. Considering a 30-nm-thickn-type

    neutral layer doped to 5 1018 cm3 at a forward bias of 0.7 V

    yields a current density of the order of 1 A/cm2 . This suggests

    that, at the sub-one-volt supplies planned for the near future,

    JFET gate current densities may compete with MOSFET gatecurrent densities.

    Section II presents a scaling analysis of enhancement mode

    JFETs for complementary logic. A comparison to the scaling be-

    havior of MOSFETs shows that the gate leakage crossover point

    lies somewhere between gate lengths of 2035 nm, depending

    on the supply voltages and gate insulator. The numerical simula-

    tions reported in Section III examine the behavior of a simple 25

    nm gate length design and find electrical properties that validate

    the scaling analysis and compare favorably with reported MOS-

    FETs. Section IV discusses the implications and opportunities

    for JFET engineering, including many of the modifications de-

    veloped for MOSFETs that should be applicable to JFETs, andthe prospects for scaling JFETs to 10 nm and below. The final

    section summarizes the findings and conclusions of this paper.

    II. SCALING ANALYSIS

    The scaling analysis presented here correlates the JFET gate

    current with the gate length. This allows for comparison with

    corresponding MOSFET gate currents, as well as with other

    properties, under scaling. A good deal of MESFET scaling

    analysis applies to JFETs [11]. However, the literature does

    not appear to include MESFET analyses suitable for examin-

    ing a JFETMOSFET gate leakage crossover point. The simple

    JFET scaling discussed later varies the structural dimensions

    and dopings under an approximate constant voltage regime. A

    constant voltage scheme applies to these enhancement mode

    devices because forward diode currents limit supply voltages

    on the high end, and at the low end, OFF-state drain currents,

    Ioff, limit the minimum practical values of the threshold volt-age Vt . The voltages used in a given setting will depend on thefigures-of-merit needed for that application. This conservative

    scaling scheme has the largest, and most problematic, dopant

    and field increases. A constant voltage scaling for the MOSFET

    is also used for comparison. Other scaling regimes are possible

    for both transistors. For example, a two-parameter generalized

    scaling is common for MOSFETs [13], and the wider variety

    of scalings available for MESFETs, depending on the figure-

    of-merit of interest [11], also applies to JFETs. In addition to

    gate currents, the properties scaled and compared next include

    the drain current, transconductance, and subthreshold current,

    estimated under the gradual channel approximation, as well as

    the gate capacitance and band-to-band tunneling. Discussions

    of the principal 2-D and short-channel effects, which are largelysimilar for JFETs and MOSFETs, are given with the numerical

    simulation results of Section III.

    A. JFET Scaling

    Fig. 1 depicts the n-channel bulk JFET scaling model. All

    dimensions scale by > 1, including the gate length L/ andwidth W/. An oxide defines the bottom of the channel, whichhas a metallurgical thickness of a/, and h/ is the depletionthickness into the channel. To reduce 2-D and short-channel

    effects, the channel will be conservatively constrained to L =2a [14]. For numerical comparisons, the reference design has

    a gate length of L = 50 nm, channel doping of Nd = 1.6

    1018 cm3 , and gate doping of Na = 1.6 1019 cm3 , which

    has a threshold voltage of approximately Vt = 250 mV. Theseparameters were chosen, in part, to enable scaling to a viable

    10 nm gate length device with minor modifications, as discussed

    in Section IV.

    For scaling, threshold will be defined here to occur

    when the depletion edge coincides with the channel bottom,

    with the drainsource potential Vds = 0. Taking the gatechannel junction to be one-sided, this happens at h = a =

    2s (Vbi Vg )/qNd , where Vg is the gate potential, s thesemiconductor permittivity, and q the elementary charge. Thebuilt-in potential Vbi will be larger than the nonscaled appliedpotentials for the enhancement mode devices under consider-

    ation, and Vbi will vary only logarithmically with dopingor even more slowly for degenerately doped material. Conse-

    quently, the channel doping will need to scale as 2 to ac-commodate the scaling ofa and h. With these assumptions andscalings, the threshold voltage

    Vt = Vbi qa2 Nd

    2s(1)

    will also change slowly with scaling. Designing a particular Vtentails changing both the channel thickness and doping.

    The forward gate current density depends on the doping and

    thus the gate length. A 1-D ideal diode model illustrates this,giving a gate current density of

    Jg =1

    qn2i

    Dn

    Na Wp+

    DpNd Wn

    (eVg /Vth 1) (2)

    where ni is the intrinsic carrier concentration, Dn and Dpthe minority carrier diffusivities, Wn and Wp the neutral layerwidths, and Vthe thermal potential. Consequently, in this model,and neglecting that diffusivity decreases with doping, the for-

    ward gate current scales as Ig 1/3 . Such a 1-D model should

    overestimate the gate current density along the channel because

    not all of the gatechannel interface sees the full gate volt-

    age, analogous with current crowding seen at the baseemitter

    Authorized licensed use limited to: K.S. Institute of Technology. Downloaded on January 30, 2010 at 10:54 from IEEE Xplore. Restrictions apply.

  • 8/7/2019 JFETNANO

    3/9

    JACKSON et al.: JUNCTION FIELD EFFECT TRANSISTORS FOR NANOELECTRONICS 751

    Fig. 2. Gate current density scaling of MOSFETs and JFETs versus gatelength. The silicon JFET curves (solid black lines), plotted for 0.6, 0.7, and0.8 V gate biases, and the GaAs JFET curves (solid blue lines) for 0.9 and1 V gate biases, come from 1-D models with bandgap narrowing. The 1 Vtunneling data [18] for MOSFET gate current densities through pure SiO 2 (red

    open diamonds) and pure HfO2 (red open triangles) are correlated with gatelengths using the international technology roadmap for semiconductors (ITRS)road map [6] and fit with exponential dependences (red dash-dotted lines). Thesilicon JFET band-to-band tunneling current density scaling estimate (greendashed line) is for an abrupt junction and a 0.7 V drain-to-gate bias. The pointsplotted from the 25 nm JFET simulation include the gate current density (blacksolid squares) at 0.6 and 0.7 V gate biases and band-to-band tunneling (greensolid circle) at 0.7 V.

    junction in bipolar transistors. The largest gate currents should

    flow for small drainsource biases. Under the present constant

    potential scaling, with the diffusion current densities in the

    quasi-neutral p and n regions determining the forward gate cur-

    rents, such 2-D effects should have the same scaling behavior

    as the 1-D model.

    However,at thesmaller dimensions of interest here, the higher

    doping levels will make bandgap narrowing important, reducing

    the gatechannel barrier and increasing the saturation current

    density. Fig. 2 plots a 1-D diode model for the gate current den-

    sity versus gate length for the scaled reference design, including

    bandgap narrowing [15], at gate potentials ofVg = 0.6, 0.7, and0.8 V. The neutral layer thicknesses for the reference design in

    this 1-D model are Wn = Wp = L/2, and the minority carrierdiffusivities are Dp = 4 cm

    2/s [16] and Dn = 6.2 cm2/s [17].

    The gate current densities trend upward with decreasing gate

    length due to bandgap narrowing, instead of following the be-havior of (2). Significantly, throughout the 1050 nm gate length

    range plotted, the gate current density for Vg = 0.6 V is on theorder of 1 A/cm2 , and the 0.7 V curve lies close to 100 A/cm2 ,

    which spans the range important for medium-performance ap-

    plications [1].

    The typical JFET drain current expressions obtained in the

    gradual channel approximation can be cast into a second-order

    MOSFET-like form for enhancement mode devices [19]. In sat-

    uration, the drain current scales as

    Idsa t =

    1

    s n

    2a

    W

    L

    Vg

    Vt2

    (3)

    where n is the channel electron mobility. For the short-channelJFETs under consideration here, electron velocity saturation at

    vsa t will generally be important, for which the saturation draincurrent approximately becomes

    Idsa t s vsa t W

    a(Vg Vt ). (4)

    With velocity saturation, the transconductance in saturation,gm sa t s vsa t W/a, does not decrease with downward scaling.For the reference design in saturation with vsa t = 10

    7 cm/s, the

    transconductance per gate width is 420 mS/mm.

    The intrinsic gate capacitance can have contributions from

    the depletion layer capacitance and, when the gate is forward-

    biased, from diffusion capacitance. The diffusion capacitance

    per area due to both electron and hole injection scales is

    Cdiff.g =1

    3qn2i2Vth

    WnNd

    +WpNa

    eVg /Vt h . (5)

    The junction capacitance per area scales as

    Cdep.g = s

    h s

    a. (6)

    Estimating their contributions in an L = 25 nm device givesCdep.g = 8.4 10

    7 F/cm2 and a much smaller Cdiff.g = 1.2

    108 F/cm2 , using a gate bias of Vg = 0.8 V and includingbandgap narrowing for the diffusion capacitance. These values

    and the respective scaling behaviors show that the diffusion

    capacitance should be negligible with respect to the depletion

    capacitance over the scales and biases of interest.

    The subthreshold drain current given by a 1-D channel barrier

    model is

    Ids =

    1

    2 n CDW

    L V2

    th e(Vg sVt )/Vt h

    1 eVd s /Vt h

    (7)

    where CD is the extrinsic Debye capacitance for the n-typechannel doping [20]. Neglecting changes in mobility, in this

    model the subthreshold current will vary with CD , which con-tributes the scaling because of the changing doping. Thisexpression gives an inverse subthreshold slope of

    S = log10 (e)Vth 60 mV/decade. (8)

    For most practical devices, though, the inverse subthreshold

    slope will be larger due to 2-D effects and to some potential

    drop across an underlying oxide.

    At the smallest dimensions, and thus largest doping con-

    centrations, band-to-band tunneling in the reverse-biased gate

    drain junction can contribute significantly to leakage. The

    tunnelingis largest when thedevice is turnedoff with thesource

    drain potential equal to the voltage supply. The tunneling cur-

    rent density depends sensitively on the local electric field. Fig. 2

    includes an estimate of the band-to-band tunneling current den-

    sity scaling based on a phenomenological fit [1] of experimental

    data from Fair and Wivell [21] and from Stork and Isaac [22],

    corresponding to the JFET turned off and a 0.7 V potential on

    the drain. The maximum electric field used for plotting comes

    from the conservative assumption of an abrupt p-n junction.

    With this estimate, the band-to-band tunneling current becomes

    larger than the forward gate current density for L

  • 8/7/2019 JFETNANO

    4/9

    752 IEEE TRANSACTIONS ON NANOTECHNOLOGY, VOL. 8, NO. 6, NOVEMBER 2009

    A more realistic, graded junction significantly reduces the max-

    imum electric field and the tunneling current, though making

    such an estimate requires more device design assumptions than

    those of Fig. 1. Solomon et al. [23] reported on extensive mea-

    surements of the tunneling currents in heavily doped p-n junc-

    tions and found an empirical fit based on the effective tunneling

    distance between bands. While that empirical fit is not wellsuited to the present scaling analysis, because of its effective

    tunneling distance dependence on doping profiles, data from

    particular measured devices can give some guidance for actual

    junctions with graded compositions. For example, a p+ n diode

    with n-doping of 6.2 1018 cm3 (wafer PN-BF2 , Q2 [23]),

    which here corresponds to the 25 nm JFET, had a tunneling

    current of 1.5 103 A/cm2 at a reverse bias of 0.7 V, which

    lies a factor of 100 below our estimate in Fig. 2. Some simple

    design changes and tunneling estimates discussed in Section IV

    indicate that prohibitively large band-to-band tunneling leakage

    can be avoided down to 10 nm gate lengths in silicon devices.

    B. MOSFET Scaling Comparisons

    The sharpest distinction between JFETs and MOSFETs un-

    der scaling occurs for their gate currents. The MOSFET gate

    current leakage due to tunneling through the oxide varies ap-

    proximatelyas et o x / , where is a material-dependent poten-tial barrier parameter. Fig. 2 plots estimates of MOSFET gate

    current densities versus gate length for two limiting cases of

    gate oxide materials, assuming either SiO2 or pure, high dielec-

    tric constant HfO2 . Experimental current densities for tunnel-

    ing through SiO2 layers from Yeo et al. [18] at a gate potential

    of 1 V are correlated with gate lengths using the 2005 ITRS

    roadmap [6] and fit with an exponential dependence in Fig. 2.

    Similarly, tunneling data at 1 V for HfO2 layers [18] are plotted

    using their equivalent SiO2 thickness and fit with an exponen-

    tial dependence. The curves in Fig. 2 indicate that the crossover

    point from MOSFET to JFET gate current densities occurs at

    roughly a 20 nm gate length for pure hafnia and 35 nm for sili-

    con dioxide, assuming a 0.7 V JFET supply voltage. Generally,

    the oxide tunneling will lie in between the fitted curves for SiO 2and HfO2 , either because silicon oxynitride is used or because

    pure hafnia will not be used. Additionally, oxide tunneling data

    for 0.7 V would shift the MOSFET curves leftward, but still

    retain the exponential dependences.

    A disadvantage of the JFET with respect to a similarly sized

    MOSFET is that the drain current can be somewhat lower. Com-paring the JFET drain current in saturation of (3) with an anal-

    ogous MOSFET square law model that includes the body effect

    coefficient m [15] gives

    IJFETdsIMOSds

    = mneff

    sox

    toxa

    . (9)

    Here, the MOSFET inversion layer effective electron mobility

    is eff, and the gate oxide permittivity and thickness are ox andtox . For m = 1.4, tox /a = 1/10, s /ox = 3, and equal mobili-ties, the idealized JFET drain current is approximately 2/5 the

    magnitude of a similarly sized MOSFET. This can be improved

    by making a smaller, while increasing Nd to maintain a given

    TABLE ISCALING OF JFET AND MOSFET PARAMETERS AND PROPERTIES

    Fig. 3. Modelfor 25 nm JFET simulation andPADREpotential solutions with

    Vds = Vgs = 0 V and 0.8 V.

    threshold voltage according to (1), though at the expense of

    increasing band-to-band tunneling in the smallest devices.

    Conversely, the JFET has an advantage of lower gate capaci-

    tance. The ratio of JFET to MOSFET intrinsic gate capacitances

    is

    CgCox

    =s

    ox

    toxa

    . (10)

    Using the same parameter estimates as for the drain currents

    gives a JFET capacitance that is approximately 1/3 that of a

    similar MOSFET.Table I compiles the JFET scaling results and their counter-

    parts from a constant potential scaling of MOSFETs. Apart from

    the dramatic difference in gate current scalings, JFETs behave

    quite similarly to MOSFETs under scaling.

    III. 25 nm JFET Simulation

    The properties of a simple n-channel, enhancement mode

    JFET model with a gate length ofL = 25 nm were investigatedusing the PADRE semiconductor device simulation tool [24].

    Fig. 3 depicts the model geometry, where the 12.5-nm-deep

    channel is doped to Nd = 5.0 1018 cm3 , and the 25-nm-

    long gate that extends 10 nm into the substrate is doped to

    p+ = 1.0 1020 cm3 . Thesource and drain contacts aredopedto n+ = 1.0 1020 cm3 andhave symmetric setbacks from thegate of 10 nm. The simulated device width was W = 100 nm.The back side contact to the 77.5 nm silicon dioxide was de-

    fined as p-type silicon and its potential was held at zero. Thesimulations presented here included field- and concentration-

    dependent mobilities, bandgap narrowing from heavy doping,

    band-to-band tunneling, ShockleyReedHall recombination,

    and FermiDirac statistics.

    Adjusting the channel doping to the value given earlier gave

    a threshold voltage of approximately 150 mV and the output

    characteristics of Fig. 4. The electric potential solutions given

    Authorized licensed use limited to: K.S. Institute of Technology. Downloaded on January 30, 2010 at 10:54 from IEEE Xplore. Restrictions apply.

  • 8/7/2019 JFETNANO

    5/9

    JACKSON et al.: JUNCTION FIELD EFFECT TRANSISTORS FOR NANOELECTRONICS 753

    Fig. 4. Simulated forward drain characteristics for a 25 nm JFET. The gatepotential increases in steps of 0.1 V from 0 to 0.8 V. At Vg = Vds = 0.7 V,gm = 690 mS/mm, and gd s = 84 mS/mm.

    in Fig. 3 correspond to the equilibrium solution and to the solu-

    tion for Vg = Vds = 0.8 V in the characteristics of Fig. 4. Thedrain current at Vg = Vds = 0.7 V, which will be taken hereas ION for comparison purposes, is around 200 A/m, at aforward gate current density of 12 A/cm2 . This ON current com-

    pares well with the drain current of approximately 300 A/mat 0.7 V measured by Krivopavic et al. [25] for an experimental

    MOSFET with L = 25 nmthough that MOSFET drive cur-rent increased to approximately 1 mA/m at 1.5 V; and thegate current density through its nitridated silicon dioxide gate

    dielectric was 25 A/cm2 at 0.9 V. Doris et al. [26] reported a a

    drive current of approximately 300 A/m for an experimentalMOSFET with L = 21nmat0.7V,whichincreasedto1mA/mat 1.5 V; however, the corresponding gate current density was

    not reported. Bai et al. [7] reported a drive current of approxi-

    mately 520 A/m for a production MOSFET with L = 35 nmat 0.7 V, which increased to 1 mA/m at 1.2 V. The aforemen-tioned JFET Ion is slightly better than, but consistent with, whatwould be expected from the JFET/MOSFET ratios of saturation

    currents given by the ratio in (9). Optimizing the simple 25 nm

    JFET design with some of the techniques proposed in Section IV

    could well narrow this performance gap. The transconductance

    at Vg = Vds = 0.7 V is approximately gm = 690 mS/mm, andthe output conductance gds = 84 mS/mm. These values give aratio ofgm /gds = 8.2, which would be suitable for many digitallogic applications.

    Fig. 5 plots the gate current versus the gate potential for

    two values of the drainsource potential, Vds = 0.0 and 0.7 V.Allocating the resulting gate currents to the gate area, W L,gives the two points plotted in Fig. 2 for Vg = 0.6 and 0.7 Vwith Vds = 0.0 V. These two values lie below the predictions ofthe scaling model of Section II. For comparison, Solomon et al.

    reported a forward diode current density of 50 A/cm2 at a bias of

    0.7 V for a p+ n diode with n-doping of 6.2 1018 cm3 (wafer

    PN-BF2 , Q2), corresponding to the 25 nm JFET gate-channel

    junction [23]. The discrepancy may be due to a combination of

    the effective area being much less than the gate area, the effective

    Fig. 5. Simulated gate current versus gate voltage for a 25 nm JFET at Vd s =0.0 and 0.7 V. For Vd s = 0.7 V, model runs with and without band-to-bandtunneling turned on show the relative contribution from tunneling.

    Fig. 6. Simulated band-to-band tunneling in the reverse-biased gatedrainjunction. Tunneling currents were estimated by taking the difference betweenmodel runs with and without band-to-band tunneling turned ON.

    neutral lengths in the 2-D simulation being larger than the 1-D

    model lengths, and possibly the differing model minority carrier

    diffusivities. The two curves for the simulated gate current with

    Vds = 0.7 V correspond to two model runs, with and withoutband-to-band tunneling turned ON. These runs allow the band-

    to-band tunneling at Vg = 0 V to be also plotted on the scalingaxes of Fig. 2, again using the gate area to estimate the current

    density. The simulated tunneling current density is close to the

    analytic estimate. Taking the difference of the curves with and

    without band-to-band tunneling gives the curve of Fig. 6. The

    point at 0.7 V gives a tunneling current of 60 pA per micrometer

    of gate length.

    Fig. 7 gives the transfer characteristics for the 25 nm JFET

    model. At Vds = 0.7 V, Ioff = 60 nA/m. This OFF current ismuch larger than the band-to-band tunneling current, indicating

    that the OFF current is a subthreshold diffusion current. Impor-

    tantly, because the band-to-band tunneling current in this model

    device is a factor of 103 less than the diffusion current, a larger

    channel doping concentration for an increased drive current

    Authorized licensed use limited to: K.S. Institute of Technology. Downloaded on January 30, 2010 at 10:54 from IEEE Xplore. Restrictions apply.

  • 8/7/2019 JFETNANO

    6/9

    754 IEEE TRANSACTIONS ON NANOTECHNOLOGY, VOL. 8, NO. 6, NOVEMBER 2009

    Fig. 7. Simulated transfer characteristics for a 25 nm JFET, on both semiloga-rithmic and linear axes. The upper curves on a logarithmic current scale exhibitinverse subthreshold slopes of 90100 mV/decade and DIBL of approximately115 mV/V. The lower curves on a linear current scale show threshold currentsofVt = 175 mV and 100 mV for Vds = 50 mV and 0.7 V, respectively.

    could be used while tolerating its larger band-to-band tunneling

    leakage. The simulated 25 nm JFET ION/IOFF ratio of 3.5 103

    at 0.7 V compares well with the measured MOSFET ratios at

    0.7 V for the devices cited earlier with ION/IOFF = 5.2 103 at

    L = 35 nm [7], 2.1 103 at L = 25 nm [25], and 2.1 103 atL = 21 nm [26]. Those MOSFET ratios increased with supplypotential to, respectively, 1.46 104 at 1.2 V [7], 7.2 103 at

    1.5 V [25], and 7.2 103 at 1.5 V [26]. The simulated JFET

    shows a 90100 mV/decade inverse subthreshold slope. This

    slope is slightly less than than the 115 mV/decade simulated for

    a 25 nm MOSFET by Taur et al. [27] and comparable to the

    95 mV/decade measurement reported by Krivokapic et al. [25].The simulated transfer characteristics of Fig. 7 exhibit the short-

    channel effect of drain-induced barrier lowering (DIBL). Using

    the threshold voltage shift between 0.05 V and 0.7 V rail volt-

    ages of approximately 0.075 V, gives a DIBL of approximately

    115 mV/V. This value compares favorably with a measured

    DIBL of 146 mV/V reported by Krivokapic et al. [25] but more

    than double the simulated value of 50 mV/V reported by Frank

    et al. [1] for a 25 nm MOSFET.

    While the abrupt junction model used here greatly simplifies

    the doping profiles expected from fabrication, its concentra-

    tions and depletion width are representative of reported junc-

    tions and should reasonably capture most behaviors of an actual

    JFET without introducing additional arbitrary model structure

    parameters to describe doping implant profiles. For the 25 nm

    JFET, relevantexperimental p-n junction examples include those

    demonstrated by Parket al. [28] andthe implant profiles reported

    by Walther et al. [29]. Orlowski [30] reviewed some research

    underway to develop just such shallow, high-concentration junc-

    tions that the ITRS roadmap [6] calls for by the end of the decade

    for MOSFETs. The abrupt junction JFET model should give

    quantitatively similar results to a graded junction, with graded

    junctions improving the transconductance linearity [14]. Irre-

    spective of profile, the threshold voltage and channel current

    are determined by the designed channel charge and depth, and

    thus quantitative comparisons with the abrupt junction model

    should be meaningful. However, the abrupt junction will give

    a noticeably higher maximum electric field, which will lead

    to an overestimate of the reverse-bias band-to-band tunneling

    leakage.

    IV. DISCUSSION: NANOELECTRONIC JFET ENGINEERING

    The scaling analysis and simulations of the preceding sec-tions strongly suggest that complementary silicon JFETs may

    compete with MOSFETs, at gate lengths below approximately

    25 nm. The first section that follows discusses optimizing and

    scaling the simulated 25 nm device. Optimizing the design may

    enable it essentially to stand in for a 25 nm MOSFET that has

    a hafnia-based dielectric at lower supply voltages. Addition-

    ally, the great geometric similarity of the two devices layouts,

    and the materials and fabrication process compatibilities, should

    confer an advantage to the JFETs over some other devices under

    consideration for MOSFET replacement, as the impacts on the

    system architecture would likely be modest. Scaling the opti-

    mized 25 nm design to near 10 nm would benefit in part fromsome techniques already developed for MOSFETs. In this scale

    range, the primary device engineering challenge will likely be

    the leakage from band-to-band tunneling in the reverse-biased

    gatedrain junction. The second section that follows discusses

    the prospects for replacing silicon with IIIV semiconductors

    in JFETs to increase channel mobilities, for example, and to

    expand the range of possible heterostructures. Similar mate-

    rial changes under consideration for MOSFETs would require

    identifying suitable gate insulators [31]. The third section that

    follows discusses how JFETs might effectively scale below the

    10 nm barrier, imposed by leakage from source-to-drain tun-

    neling [1]. Such scaling may require wraparound gates, to con-

    trol short-channel effects, as well as source and drain materialchanges, to control sourcedrain tunneling.

    A. Optimizing and Scaling Silicon-Based JFETs

    Some straightforward improvements to the simple 25 nm de-

    sign could increase the ON drain current, reduce the leakage

    currents, and reduce the footprint. Referring to (3) for guidance,

    decreasing the channel thickness a increases the ON current,and the threshold voltage can remain the same provided Nd in-creases to keep the product a2 Nd constant. Fig. 5 shows thatthe greatest gate current flows for large gate bias and small

    sourcedrain potential, as would be found in an inverter with

    a high input. About half of this current flows through the end

    faces of the p-gate region simulated in Section III. Forming

    the gate instead by depositing a heavily doped polycrystalline

    silicon gate directly on the channel, using techniques similar

    to those used for bipolar junction transistor emitters, would re-

    duce the junction area and current. Such a raised gate would

    similarly reduce band-to-band tunneling at the drain end of

    the gate and reduce gatesource and gatedrain capacitances.

    Importantly, the polycrystalline-silicon-raised gate would facil-

    itate a self-aligned gate process directly analogous to that used

    for MOSFETs. Reducing the source and drain contact setbacks

    can reduce the device footprint, until band-to-band tunneling

    at the gatedrain junction becomes prohibitive. Engineering the

    Authorized licensed use limited to: K.S. Institute of Technology. Downloaded on January 30, 2010 at 10:54 from IEEE Xplore. Restrictions apply.

  • 8/7/2019 JFETNANO

    7/9

    JACKSON et al.: JUNCTION FIELD EFFECT TRANSISTORS FOR NANOELECTRONICS 755

    lateral doping gradient between the gate and drain can reduce

    the maximum electric field, and consequently, the tunneling, al-

    lowing a shorter setback, and the setbacks can be asymmetric,

    with a smaller sourcegate setback.

    The basic enhancement mode JFET can benefit from tech-

    nologies developed to scale MOSFET performance without

    changing the gate length [32]. The raised gate can serve asan ion implantation mask for a halo implant to reduce short-

    channel effects. Uniaxial strain can improve mobility [33], [34].

    Fabricating a p-channel transistor on (1 1 0) silicon can decrease

    the effective mass, and thus, increase the hole current. The JFET

    also lends itself to using higher mobility silicon germanium al-

    loys in the channel, though material with smaller bandgaps can

    greatly increase the band-to-band tunneling leakage.

    Band-to-band tunneling in the reverse-biased gatedrainjunc-

    tion presents a design challenge in scaling the 25 nm device to

    10 nm. This problem is directly analogous, and of a similar mag-

    nitude, to the well-studied drain-to-body tunneling problem in

    MOSFETs [1]. For the JFET, the constant potential scaling of

    Section II indicates that the channel thickness should decreasefrom 12.5 nm to 5 nm, and the channel doping should increase to

    Nd = 4.0 1019 cm3 . Using the 1-D, one-sided abrupt junc-

    tion model plotted in Fig. 2, with a 0.7 V reverse bias, band-to-

    band tunneling would be of the order of 100 kA/cm2 . If, instead,

    a linearly graded junction model is used, then the estimate is

    of the order of 1 kA/cm2 [21]. The tunneling measurements

    from Solomon et al. include data from two p+ n junctions with

    n dopings of Nd = 2.21 1019 and 9.05 1019 cm3 (wafer

    PN-BF2 Q5 and Q8) that gave tunneling currents of 3 A/cm2 and

    70 A/cm2 , respectively. Those data show both that the tunnel-

    ing estimates for the JFET are conservative and that the tun-

    neling leakage at 10 nm may well be manageable. Generally,a too large leakage could in part be reduced by using a less

    conservative L/a 1.5, enabled by a halo implant to controlshort-channel effects, and allowing for a lower channel doping.

    Further reducing the doping gradient at the metallurgical junc-

    tion also helps. For example, in the idealized limit of an undoped

    10 nm layer between the gate and channel dopings, giving an

    electric field of approximately 1.7 MV/cm, the tunneling current

    would beon the order of1 A/cm2 [1]. This dopant gradient could

    be approximated by having the peak of the channel doping occur

    at the bottom of the channel, by not fully scaling the gatedrain

    setback, and by having the gatechannel metallurgical junction

    occur slightly up in the gate.

    B. Compound Semiconductors and Heterostructures

    Changing the JFET semiconductor from silicon to a IIIV

    compound semiconductor can improve some device properties

    and allow for a wider range of heterostructures for device en-

    gineering. The semiconductors used for JFETs include Si, SiC,

    GaAs, InP, GaP, and InGaAs/InP [35][38]. Enhancement mode

    GaAs JFETs have been studied for ICs [39]. A similar semicon-

    ductor change for MOSFETs entails developing suitable di-

    electrics, greatly complicated by the ubiquitous interface states

    at the IIIV semiconductoroxide interface [40]. The perfor-

    mance benefits can include larger mobilities andlarger bandgaps

    to decrease gate leakage. For example, Fig. 2 includes gate cur-

    rent density estimates for enhancement mode GaAs JFETs using

    a 1-D diode model. The curves in Fig. 2 use constant values for

    the product of minority carrier diffusivity and effective intrin-

    sic concentration, with nie D of 1013 cm4/s for electrons and

    2 1015 cm4/s for holes, because for GaAs at very high dop-

    ing levels, bandgap narrowing is counterbalanced by a wideningdue to degeneracy effects [41]. At a given forward bias, these

    calculations give several orders of magnitude decease in gate

    current density compared to silicon devices. This suggests that

    GaAs-based JFETs could be the preferred devices for low-power

    applications, which otherwise impose a minimum gate length of

    approximately 35 nm for silicon MOSFETs [1]. Additionally,

    the larger bandgap of GaAs would greatly reduce leakage from

    band-to-band tunneling. The mobility of the channel can be

    increased by defining the channel bottom with a heterobarrier

    and using modulation or remote doping. Using a smaller

    bandgap gate could significantly reduce gate injection into the

    JFET channel, though a larger bandgap in the gate depletion

    field would still be desirable to avoid band-to-band tunneling.Further modifications can include using high-mobility quantum

    wells to define the channel, which have been shown to lead to

    large transconductances [42].

    C. Nanowire JFETs and the 10 nm Barrier

    Double gate and wraparound gate geometries have been in-

    vestigated for MOSFETs to control short-channel effects [32],

    [43]. JFETs lend themselves to similar modifications, which

    may be necessary for scaling below the 10 nm barrier. A cylin-

    drical geometry has the best scaling properties for controlling

    short-channel effects [44]. Nanowires grown with a catalyst nat-

    urally have this geometry. They allow much greater compositionmodulation than planar structures because the strain constraint

    is relaxed, and thus a greater range of electronic structure en-

    gineering [45][48]. Well-defined heterostructures in both the

    radial and axial direction have been reported [48], [49]. Radial

    heterostructures can, in principle, be used to engineer the gate

    leakage, and axial heterostructures to control the sourcedrain

    tunneling. Epitaxial IIIV nanowires have been grown directly

    on silicon [50]. Nanowire MOSFETs have been made both in

    silicon and IIIV semiconductors [51][54]. Schottky contacts

    for the source and drain can be utilized to reduce the source

    to drain leakage current, with nickel disilicide contacts to sili-

    con nanowire reported [55]. However, practical nanowire JFETspresent great challenges in growth and doping control.

    V. CONCLUSION

    This paper has examined the implications of a gate leakage

    crossover point from MOSFETs to JFETs under scaling. A sim-

    ple scaling analysis suggests that a practical crossover point

    lies in the vicinity of 25 nm, depending on the supply voltage

    and the MOSFET gate insulator being considered. Otherwise,

    the analysis indicates that while the JFET drain current can be

    somewhat smaller than a similar MOSFET, the gate capacitance

    is also smaller, and other properties scale similarly to those of

    MOSFETs. Numerical simulations of a 25 nm JFET showed

    Authorized licensed use limited to: K.S. Institute of Technology. Downloaded on January 30, 2010 at 10:54 from IEEE Xplore. Restrictions apply.

  • 8/7/2019 JFETNANO

    8/9

    756 IEEE TRANSACTIONS ON NANOTECHNOLOGY, VOL. 8, NO. 6, NOVEMBER 2009

    behavior and performance comparable to reported 25 nm MOS-

    FETs at the same lower supply voltage and consistent with

    the JFET scaling analysis. A number of modifications to the

    idealized 25 nm model to increase its performance, including

    a self-aligned gate proposal and the application of techniques

    already developed for MOSFETs, were outlined that may en-

    able it to stand in for 25 nm MOSFETs in some circumstances,and subsequently, be scaled to the 10 nm range. If IIIV semi-

    conductors replace silicon in future 10-nm-scale technologies,

    JFETs have a natural advantage over MOSFETs because of the

    great challenge of developing a high-quality insulator for IIIV

    semiconductors. Because of the geometrical and electrical sim-

    ilarities of the JFET and the MOSFET, the consequences for the

    system architecture may well be modest, especially compared to

    the architectural changes that would be needed for other devices

    under development for MOSFET replacements. Taken together,

    the results presented here suggest that the crossover point oc-

    curs at a technologically relevant point between 20 and 30 nm

    for silicon devices, and that JFETs may well be of interest for

    upcoming silicon technology nodes, and perhaps beyond.

    REFERENCES

    [1] D. J. Frank, Power-constrained CMOS scaling limits, IBM J. Res.Develop., vol. 46, no. 2/3, pp. 235244, Mar./May 2002.

    [2] D. J. Frank, R. H. Dennard, E. Nowak, Y. T. Paul, M. Solomon, andH.-S. P. Wong, Device scaling limits of Si MOSFETs and their appli-cation dependencies, Proc. IEEE, vol. 89, no. 3, pp. 259288, Mar.2001.

    [3] J. D. Plummer and P. B. Griffin, Material and process limits in siliconVLSI technology, Proc. IEEE, vol. 89, no. 3, pp. 240258, Mar. 2001.

    [4] T. Ghani, K. Mistry, P. Packan, S. Thompson, M. Stealer, S. Tyagi, andM. Bohr, Scaling challenges and device design requirements for highperformance sub-50 nm gate length planar CMOS transistors, in Proc.2000 Symp. VLSI Technol. Dig. Tech. Papers, 2000, pp. 174175.

    [5] S. H. Lo, D. A. Buchanan, Y. Taur, and W. Wang, Quantum-mechanicalmodeling of electron tunneling current from the inversion layer of ultra-thin-oxide nMOSFETs, IEEE Electron Device Lett., vol. 18, no. 5,pp. 209211, May 1997.

    [6] SIA, International technology roadmap for semiconductors, Semicon-ductor Industry Association (SIA), Semiconductor Industry Association,San Jose, CA, 2005.

    [7] P. Bai, C. Auth, S. Balakrishnan, M. Bost, R. Brain, R. H. V. Chikarmane,M. Hussein, J. Hwang, D. Ingerly, R. James, C. K. J. Jeong, E. Lee,S.-H. Lee, N. Lindert, M. Liu, Z. Ma, T. Marieb, A. Murthy, R. Nagisetty,S.Natarajan, J.Neirynck, A. Ott, J.S. C. Parker, R.Shaheed,S. Sivakumar,C. Weber, B. Woolery, A. Yeoh, K. Zhang, and M. Bohr, A 65 nm logictechnology featuring 35 nm gate lengths, enhanced channel strain, 8 Cuinterconnect layers, low-k ILD and 0.57 um2 SRAM cell, in Proc. 2004

    Int. Electron Devices Meeting, pp. 657660.[8] K. Mistry, C. Allen, C. Auth, B. Beattie, D. Bergstrom, M. Bost,

    M. Brazier, M. Buehler, A. Cappellani, R. Chau, C.-H. Choi, G. Ding,K. Fischer, T. Ghani, R. Grover, W. Han, D. Hanken, M. Hattendorf,J. He, J. Hicks, R. Heussner, D. Ingerly, P. Jain, R. James, L. Jong,S. Joshi, C. Kenyon, K. Kuhn, K. Lee, H. Liu, J. Maiz, B. McIntyre,P. Moon, J. Neirynck, S. Pae, C. Parker, D. Parsons, C. Prasad, L. Pipes,M. Prince, P. Ranade, T. Reynolds, J. Sandford, L. Shifren, J. Sebastian,J. Seiple, D. Simon, S. Sivakumar, P. Smith, C. Thomas, T. Troeger,P. Vandervoorn, S. Williams, and K. Zawadzki, A 45 nm logic technol-ogy with high-k+metal gate transistors, strained silicon, 9 Cu interconnectlayers, 193 nm dry patterning, and 100% Pb-free packaging, in Proc.

    Electron Devices Meeting 2007, IEDM Tech. Dig., pp. 247250.[9] B.-Y. Tsaur and H. K. Choi, Radiation-hardened silicon-on-insulator

    complementary junction field-effect transistors, IEEE Electron DeviceLett., vol. EDL-7, no. 5, pp. 324326, May 1986.

    [10] L. F. Halle and J. F. Knudsen, JFET SOS devices-Part I: Transistorcharacteristics and modeling results, IEEE Trans. Electron Devices,vol. 35, no. 3, pp. 353358, Mar. 1988.

    [11] G. V. Ram and M. I. Elmasry, On the scaling of Si-MESFETs, IEEEElectron Device Lett., vol. EDL-1, no. 12, pp. 259262, Dec. 1980.

    [12] K. P. MacWilliams and J. D. Plummer, Device physics and tech-nology of complementary silicon MESFETas for VLSI applications,

    IEEE Trans. Electron Devices., vol. 38, no. 12, pp. 26192631, Dec.1991.

    [13] G. Baccarani, M. R. Wordeman, and R. H. Dennard, Generalized scalingtheory and its application to a micrometerMOSFET design, IEEE Trans.

    Electron Devices., vol. ED-31, no. 4, pp. 452462, Apr. 1984.[14] S. M. Sze, Physics of Semiconductor Devices. New York: Wiley, 1981.[15] Y. Taur and T. H. Ning, Fundamentals of Modern VLSI Devices. Cam-

    bridge, U.K.: Cambridge Univ. Press, 2002.[16] C. H. Wang, K. Misiakos, and A. Neugroschel, Minority-carrier transport

    parametersin n-type silicon, IEEE Trans.ElectronDevices,vol.37,no.5,pp. 13141322, May 1990.

    [17] S. E. Swirhun, Y.-H. Kwark, and R. M. Swanson, Measurement of elec-tron lifetime, electron mobility and band-gap narrowing in heavily dopedp-type silicon, in Proc. IEDM, 1986, vol. 86, pp. 2427.

    [18] Y.-C. Yeo, T.-J. King, and C. Hu, Direct tunneling leakage current andscalability of alternative gatedielectrics, Appl.Phys. Lett.,vol.81,no.11,pp. 20912093, Sep. 2002.

    [19] D. A. Neaman, in Semiconductor Physics and Devices, K. Butcher, Ed.New York: McGraw-Hill, 2003.

    [20] R. J. Brewer, The barrier mode behaviour of a junction FET at lowdrain currents, Solid State Electron., vol. 18, pp. 10131017, 1975.

    [21] R. B. Fair and H. W. Wivell, Zener and avalanche breakdown in As-implanted low-voltage Si n-p junctions, IEEE Trans. Electron Devices,vol. ED-23, no. 5, pp. 512518, May 1976.

    [22] J. M. C. Stork and R. D. Isaac, Tunneling in base-emitter junctions,IEEE Trans. Electron Devices, vol. ED-30, no. 11, pp. 15271534, Nov.1983.

    [23] P. M. Solomon, J. Jopling, D. J. Frank, C. DEmic, O. Dokumaci,P. Ronsheim, and W. E. Haensch, Universal tunneling behavior in tech-nologically relevant P/N junction diodes, J. Appl. Phys., vol. 95, no. 10,pp. 58005812, 2004.

    [24] [Online]. Available: Simulations were performed by PADRE onhttp://nanohub.org

    [25] Z. Krivokapic, W. Maszara, F. Arasnia, E. Paton, Y. Kim, L. W. Marathe,and M.-R. Lin, High performance 25 nm FDSOI devices with extremelythinsiliconchannel, in Proc. 2003 Symp.VLSl Technol. Dig.Tech. Papers,pp. 131132.

    [26] B. Doris, M. Ieong, T. Kanarsky, Y. Zhang, R. Roy, O. Dokumaci, Z. Ren,F.-F. Jamin, L. Shi, W. Natzle, H.-J. Huang, J. Mezzapelle, A. Mocuta,S. Womack, M. Gribelyuk, E. Jones, R. Miller, H.-S. Wong, andW. Haensch, Extreme scaling with ultra-thin Si channel MOSFETs, inProc. Electron Devices Meeting, 2002, IEDM 02. Digest. Int. , pp. 267270.

    [27] Y. Taur, C. H. Wann, and D. J. Frank, 25 nm CMOS design considera-tions, in Proc. IEDM, 1998, vol. 29, pp. 789792.

    [28] H.-S.Park,K. Jeong, H.-W. Suh, H.-J.Jung, andW.-K. Choi, 5.6-nm p+/n junction formation for sub-0.05-m PMOSFETs by using low-energyB10 H14 ion implantation, J. Korean Phys. Soc., vol. 44, no. 6, pp. 15941597, 2004.

    [29] S. Walther, L. Godet,T. Buyuklimanli, and J. Weeman, Plasma implantedultra shallow junction boron depth profiles: Effect of plasma chemistryand sheath conditions, J. Vac. Sci. Techol. B, Microelectron., vol. 24,no. 1, pp. 489493, 2006.

    [30] M. Orlowski, CMOS challenges of keeping up with Moores law, in

    Proc. Adv. Thermal Process. Semicond. 2005, 13th IEEE Int. Conf., p. 19.[31] R. Chau,S. Datta,M. Doczy,B. Doyle, B.Jin, J. Kavalieros, A. Majumdar,

    M. Metz, andM. Radosavljevic, Benchmarkingnanotechnologyfor high-performance and low-power logic transistor applications, IEEE Trans.

    Nanotechnol., vol. 4, no. 2, pp. 153157, Mar. 2005.[32] H. S. P. Wong,Beyond theconventional transistor, IBMJ. Res. Develop.,

    vol. 46, no. 2/3, pp. 133168, Mar./May 2002.[33] C. S. Smith, Piezoresistance effect in germanium and silicon, Phys.

    Rev., vol. 94, pp. 4249, 1954.[34] B. M. Haugerud, L. A. Bosworth, and R. E. Belford, Mechanically in-

    duced strain enhancement of metaloxidesemiconductor field effect tran-sistors, J. Appl. Phys., vol. 94, no. 6, pp. 41024107, Sep. 2003.

    [35] J. McGarrity, F. McLean, W. DeLancey, J. Palmour, C. Carter, J. Edmond,and R. Oakley, Silicon carbide JFET radiation response, IEEE Trans.

    Nucl. Sci., vol. 39, no. 6, pp. 19741981, Dec. 1992.[36] C. A. Liechti, Microwave field-effect transistors-1976, IEEE Trans.

    Microw. Theory, vol. MTT-24, no. 6, pp. 279300, Jun. 1976.

    Authorized licensed use limited to: K.S. Institute of Technology. Downloaded on January 30, 2010 at 10:54 from IEEE Xplore. Restrictions apply.

  • 8/7/2019 JFETNANO

    9/9