36
  2.1 Poisson's Equati 2 ' phenomena, fem I only (2.1-1). As the next step i  potential i/J. For remembering th:  B=rot  Note that we as compared to all We substitute (2 rot ( £ lf "rot  z = O" hol can be expresse ex pressed as: Some Fundamen tal Pr operties To accurately analyze an arbitrary semiconductor strucll1£$ which is intended as a self contained de vice under various op erating conditions, a mathematical model has t o be given. The equations which form this mathematical model are commonly called the basic semiconductor equations. They can be derived from Maxwell's equations (2-1), (2-2), (2-3) and (2-4), several relations obtained from solid-state physics knowledge about semiconductors and various - sometimes overly simplistic - assumptions.  E=- rotH=l+- (2-1) at aB = -- (2-2) at div 15=  p (2-3) div B=O (2-4)  Now we substit 15= - The first term homogenous. 1 Poisson's equ div (s The space char] elementary che the negatively which will be !  E and i5 are the electric field and displacement vector; Hand Bare the magnetic field and induction vector, respectively.] denotes the conduction current density, and  pis the electric charge density. The next sections will be devoted entirely to an outline of the procedures which have t o be carrie d out in order to derive the basic semicond uctor equations. 2.1 Poisson's Equation Poisson's eguation is essentially the third Maxwell equation  (2-3). However, to make this equation directly applicable to semiconductor problems, some manipulations have to be undertaken. We first introduce a relation for the electric displacement vector i5 and the electric field vector  E (2.1-1).  p=q From a purel without intro!  broug ht about Sections 2.3 a1 The permitti qua ntity. In p~ materials cum of the permi amorphous str (2.1-1) e denotes the ~rmittivity tensor. This relation is valid for all materials which have a time independent permittivity. Furthermore, polarization by mechanical forces is neglected [2.44]. Both assumptions hold relatively well considering the usual applications of semiconductor device s. However, an investiga tion of piezoelectric

Material ISM Examenc

Embed Size (px)

DESCRIPTION

x

Citation preview

  • 2.1 Poisson's Equati

    2 ' phenomena, fem I only (2.1-1).

    As the next step i

    potential i /J. For remembering th:

    B=rot

    Note that we as

    compared to all

    We substitute (2

    rot (

    lf "rot z = O" hol can be expresse

    expressed as:

    Some Fundamental Properties

    To accurately analyze an arbitrary semiconductor strucll1$ which is intended as a self

    contained device under various operating conditions, a mathematical model has to be given. The

    equations which form this mathematical model are commonly called the basic semiconductor

    equations. They can be derived from Maxwell's equations (2-1), (2-2), (2-3) and (2-4), several

    relations obtained from solid-state physics knowledge about semiconductors and various -

    sometimes overly simplistic - assumptions. E=-

    at5 rotH=l+- (2-1) at aB

    rot E= -- (2-2) at

    div 15= p (2-3)

    div B=O (2-4)

    Now we substit

    15= -

    The first term

    homogenous. 1

    Poisson's equ

    div (s The space char]

    elementary che

    the negatively

    which will be !

    E and i5 are the electric field and displacement vector; Hand Bare the magnetic field and induction vector, respectively.] denotes the conduction current density, and pis the electric

    charge density.

    The next sections will be devoted entirely to an outline of the procedures which have to be

    carried out in order to derive the basic semiconductor equations.

    2.1 Poisson's Equation

    Poisson's eguation is essentially the third Maxwell equation (2-3). However, to make this

    equation directly applicable to semiconductor problems, some manipulations have to be

    undertaken. We first introduce a relation for the electric displacement vector i5 and the electric field vector E (2.1-1).

    p=q

    From a purel

    without intro!

    brought about

    Sections 2.3 a1

    The permitti

    quantity. In p~

    materials cum

    of the permi

    amorphous str

    (2.1-1)

    e denotes the ~rmittivity tensor. This relation is valid for all materials which have a time

    independent permittivity . Furthermore, polarization by mechanical forces is neglected [2.44].

    Both assumptions hold relatively well considering the usual applications of semiconductor

    devices. However, an investigation of piezoelectric

    I

    I

  • 2.1 Poisson's Equation 9

    ended as a

    model has

    commonly

    Maxwell's

    solid-state

    ies overly

    phenomena, ferroelectric phenomena and nonlinear optics is impossible when using } only

    (2.1-1).

    As the next step it is desirable to relate the electric field vector E to the electrostatic potential If;. For that purpose we solve (2-4) by introducing a vector field A and remembering that "div rot" applied to any vector is always zero.

    B=rotA,j!Y-A-=0-~'.~- -:ffr / rol- s: = -:> (2.1-2) Note that we assume for the applicability of (2.1-2) that the speed of light is large compared

    to all velocities which are relevant for a device [2.44].

    We substitute (2.1-2) into (2-2) and we obtain readily (2.1-3).

    - a.4)

    rot E+at =0

    (2.1-3)

    If "rot z = O" holds for a vector field z we know from basic differential calculus that z can be expressed as a gradient field. Therefore, the electric field vector E can be expressed as:

    - a.4 E= ---grad!f; at

    (2.l-4)

    (2-1

    )

    Now we substitute (2.l-4) into (2.1-1) and then the result into (2-3).

    - a.4 D = - e . - - c; . grad 1/; i3 t

    div(:-~~)+ div (e. grad If;)= -p

    (2.1-5)

    (2-2)

    (2-3) (2-4)

    gnetic

    field ty,

    and pis

    (2.1-6)

    vhich have

    The first term in (2.1-6) is zero if the permittivity s can' be considered to be homogenous.

    Thus, we finally end up with (2.1-7) which is the well known form of Poisson's equation.

    div (e. grad If;)= -p (2.1-7) The space charge density p can be further broken apart (2.1-8) into the product of the

    elementary charge q times the sum of the positively charged hole concentration p, the

    negatively charged electron concentration n and an additional concentration C which will

    be subject of later investigations.

    owever, to

    ne mani-

    he electric

    p = q . (p - n + C) (2.1-8)

    (2.1-1

    )

    From a purely mathematical point of view (2.1-8) represents a substitution only, without

    introducing any assumptions. However, additional assumptions are brought about by

    modeling the quantities n, p etc. as will become clearly apparent in Sections 2.3 and 2.4.

    The permittivity s will be treated here in all further investigations as a scalar quantity. In

    principle it has to be represented as a tensor of rank two. However, the materials currently

    in use for device fabrication do not show a significant anistropy of the permittivity owing

    to their special composition, e.g. cubic lattice or amorphous structure. Inhomogeneity

    effects of the permittivity have been neglected

    rich have a

    al forces is

    the usual

    ezoelectric

    (

  • IO 2 Some Fundamental Properties 2.3 Carrier Transport Equations

    material e, [] --

    Si 11.7

    Si02 3.9

    Si3N4 7.2 typical Ga As 12.5 Ge 16.l

    This result is interpreted fai

    total conduction current an

    charge. In order to obtain t

    carried out. We first define a

    making use of the d~o1

    - an divJ -q- -=q n ot

    in (2.1-7). There does not exist pronounced experimental evidence for inhomogeneity

    effects. For some materials the relative permittivity constants s, = e/e0 are summarized in Table 2.1-1.

    Table 2.1-1. Relative permittivity constants

    div grad t/J =CJ_. (n-p-C) e

    (2.1-9)

    . - op div J +q--=-

    P 8t

    It is obvious that we can not

    different ways (2.2-5), (2.2-6;

    equation more easily. The 4

    the net generation or reco

    recombination and negativ

    about the structure of R ex.

    carefully (cf. Section 4.2) us

    doctors. If we have a moc

    considered as two equations

    no necessity or even evidenc

    upon local quantities and

    recombination phenomena

    considering only the deriva

    In particular for Si3N4 the value of e, depends strongly on the individual processing

    conditions; it can vary quite significantly.

    Ifwe introduce (2.1-8) and the assumption of a homogeneous scalar permittivity into (2.1-

    7), we obtain the final form of Poisson's equation to be uSeal'or semiconductor device

    modeling.

    2.2 Continuity Equations

    The continuity equations can be derived in a straightforward manner from the first Maxwell

    equation (2-1). If we apply the operator "div" on this equation we obtain:

    - - op divrotH=divJ +-=0 (2.2-1) ot

    2.3 Carrier Transp

    (2.2-4)

    The derivation of current

    cumbersome task. It is not

    wide field of physics behim

    relations in detail. Therefo

    proof, but with reference 14

    Without loss of generality t

    the charge constant per par

    _{lrift velocity} of the

    particl density can be written

    as (

    JP=q-p-vP 1.=

    -q n v.

    The major problem is to fin

    to the electric field vector

    information about the drift

    means of a distribution fu

    coordinates x=(x,y,z)r, m

    ~en dimensional space.1

    Now we split the conduction current density J into a component JP caused by holes and a component Jn caused by electrons:

    (2.2-2)

    Furthermore, we assume that all charges in the semiconductor, except the mobile carriers

    electrons and holes, are time invariant. Thus we neglect the influence of charged defects,

    e.g. vacancies, dislocations, deep recombination traps, which may change their charge state

    in time.

    oc -=0 ot If we substitute (2.1-8) and (2.2-2) into (2.2-1) and if we make use of (2.2-3) we obtain:

    (2.2-3)

    )

    '

  • 'roperties 2.3 Carrier Transport Equations IL

    nhomo

    f./e0 are

    This result is interpreted fairly trivially. It just means that sources and sinks of the total

    conduction current are fully compensated by the time variation of the mobile

    charge. In order to obtain two continuity equations a few formal steps have to be carried

    out. We first define a quantity R in (2.2-5) and, secondly, we rewrite (2.2-4) by

    making use of the definition R.

    - an divJ -q -=q - R n at (2.2-5

    )

    (2.1-9

    ).

    . - ap div J + q . - = - q. R v at

    It is obvious that we can not gain informat ion by writing one equation (2.2 -4) in two

    different ways (2.2-5), (2.2-6). However, these formal steps enable us to interprete the

    equation more easily. The quantity R can be understood as a function describing the net

    generation or recombination of electrons and holes. Positive R means recombination

    and negative R means generation. So far we have no informat ion about the structure of

    R except equations (2.2-5) and (2.2-6). R has to be modeled carefully (cf. Section 4.2)

    using knowledge from the solid -state physics of semiconductors. If we have a model

    for R, equations (2.2-5) and (2.2-6) can really be considered as two equations. It seems

    worthwhile to note explicitly here that there is no necessity or even evidence that R can

    be expressed as a function depending only upon local quantities and not upon integral

    quantities; non-local generation or recombination phenomena may certainly take place

    in semiconductor devices considering only the derivation of the continuity equations.

    (2.2-6)

    icessmg

    /ity into

    nducto

    r

    the

    first

    obtain:

    (2.2-1) 10

    'Y holes

    (2.2-2)

    mobile

    ience of

    . ch

    may

    (2.2-3)

    2-3)

    we

    (2.2-4)

    2.3 Carrier Transport Equations

    The derivation o f current relations for the semiconductor equations is a very

    cumbersome task. It is not the intention of this book to cover the extraordinarily wide

    field of physics behind all the considerations necessary to derive the current relations

    in detail. Therefore, some of the required relations will be given without proof, but

    with reference to a text more specialized in that field.

    Without loss of generality the current density of charged particles is the product of the

    charge constant per particle, the part icle concentration and the average velocity ~drift

    velocit}'.} of the particles. So the hole current density and the electron current density

    can be written as (2.3-1) and (2.3-2), respectively .

    Jv=q-p-vp

    Jn= =tl : n. vn (2.3-1

    )

    (2.3-2

    )

    The major problem is to fin~ expressions which relate the average carrier velocit ies to

    the electric field vector E and to the carrier concentration. In order to obtain

    information about the drift velocity we have to describe the carrier concentration by means of a distribution function fv in phase space which is the space of spatial

    coordinates x = (x, y, z)", momentum coordinates k = (kx, kY , k,f and time t , thus a seven dimensional space. The distribution function determines the carrier concen-

    )

    I

  • 1

    2

    2 Some Fundamental Properties 2.3 Carrier Transport Eqi

    tration per unit volume of phase space. By integrating the distribution function over the

    entire momentum volume V k we obtain the carrier concentration v (x, t ). v stands for n or p,

    denoting electrons or holes.

    4.1n3 f fJx,k,t )-dk=v(x,t ) (2.3-3

    )

    (2.3-8) is termed~

    number of carriers sc

    time . .fv (:X, k, t) gives [1-fv(x,k',t )] gives

    unoccupied and can, t

    of the scattering even1

    equals the number of

    unit time. Thorough i

    found in, e.g., [2.21]

    The derivative of x. carriers.

    Vk

    This normalization (2.3-3) defines .fv as a probability. In the literature various different normalizations can be found, e.g. [2.42], [2.49]. The distribution function has the property that its derivative along a particle trajectory xv(t ),

    k.(t ) with respect to time vanishes in the entire phase space in compliance with the Liou

    ville theorem about the invariance of the phase volume for a system moving along the

    phase paths or an account of the conservation of the number of states [2.49]. d +

    xv dt=uv

    d dt I, (xv (t), kv (t), t) = O (2.3-4

    )

    We have now to subst

    the Boltzmann trans1

    of; ftve -+-.g at -n

    = - f (fv(x.

    Equation (2.3-4) is the Boltzmann transport equation in implicit form. By expanding the total derivative we obtain:

    (2.3-5)

    Vk'

    Here grad, denotes the gradient operator with respect to the momentum coordinates k; grad , is the gradient operator with respect to the spatial coordinates x. Equation (2.3-5tshows that the variation of the distribution fu. nction at each point of phase

    space (x, k) with time is caused by the motion of particles in normal space(x) and in mom en

    tum space t: The derivative of kv wi~h respect to time multiplied with Planck's constant h equals the sum of all forces F. These forces have to be devided into two classes (2.3-7).

    -fv(x,

    dkv

    dt

    t; h h=-

    h ' 2 ti (2.3-6)

    A fairly accurate appr

    carrier concentrations

    ficult task to accompl

    seven independent v a

    rather requires the use

    for numerical approac

    [2.42].

    An alternative approa

    the motion of one or n

    e.g. [2.43]. However,

    [2.64], [2.65] and thei

    application. One should be awarer

    implies already seven

    The scattering pro

    The duration of a c particle; collisions

    Carrier-carrier inte the right hand sid:

    External forces an dimensions of the

    F.=Fve+Fvi (2.3- 7)

    f ve comprises forces due to macroscopic external fields and ft vi denotes forces due to internal localized crystal attributes like impurity atoms or ions, vacancies, and thermaf)att1ce vibrations. It is quite impossible to calculate the effect of internal j forces F vi upon the distribution function from the laws of dynamics [2.49]. Statistical laws have to be invoked instead. By introducing the quantity S; (k, k'). d k' which is the probability per unit time that a carrier in the state k will be scattered into the momentum volume dk ',we can write the internal collision term as follows:

    t vi I ( _ ( _ ) __ grad, .fv J; = fv (.x, k, t) 1-.fv (x, k', t) S; (k, k')-

    Vk' (2.3-8) - .fv (x, k', t) (1-.fv (x, t; t)) S. (k', k)) d/('

    I I

    '

  • 2.3 Carrier Transport Equations 13

    al Properties

    (2.3-3)

    (2.3-8) is termed the collision integral. The first term in the integrand describes the number

    of carriers scattered from the state k into the volume element dk' per unit time. j~ (x, k, t) gives the probability that a carrier initially occupies the state k. [1-fv (x , k ', t)] gives the probability that the volume element d k' is initially unoccupied and can, therefore, accept a carrier. S; (k, k') gives an a priori probability of the scattering event. Correspondingly, the

    second term in the integrand of (2.3-8) equals the number of electrons scattered from

    volume element dk' into state k per unit time. Thorough investigations about the scattering probabilitx S; (k, k') can be found in, e.g., [2.21 ], [2.95]. The derivative of xv with respect to time represents the group velocity of the carriers.

    iction over

    t). v stands

    re vanous

    a particle e

    space in

    volume for

    tion of the r

    Xv

    dt=uv (2.3-9)

    (2.3-5)

    We have now to substitute the relations (2.3-6) to (2.3-9) into (2.3-5), and we obtain the Boltzmann transport eguation in exnlicit form.

    of, f ve -+- gradkfv+u, gradxfv = ot n

    = - J tJ: (x,i(, t) (1- I. (x, k', t)) . s. (k, k')- Vk'

    - I, (x, k', t) (1- I; (x, k, t)) S, (k', k)) dk'

    (2.3-10)

    (2.3-4)

    expanding

    itum coor-

    rdinates :X

    :. ch point

    of 11

    space(x)

    (2.3-6)

    A fairly accurate approach would be to directly solve (2.3-10) in order to calculate carrier

    concentrations and drift velocities. However, this is an extraordinarily difficult task to

    accomplish. (2.3-10) represents an integro-differential equation with seven independent

    variables. This equation does not admit a closed solution. It rather requires the use of

    iterative procedures which, moreover, are scarcely suitable for numerical approaches [2.10],

    or additionally, invoke very stringent assumptions [2.42].

    An alternative approach to solving the Boltzmann equation consists in simulating the

    motion of one or more carriers ~t microscopic level with Monte Carlo methods, e.g. [2.43].

    However, this category of simulations is very computationally intensive [2.64], [2.65] and

    therefore, with a few exceptions only, not suitable for engineering application.

    One should be aware of the fact that the validity of the Boltzmann equation (2.3-10) implies

    already several assumptions (cf. [2.10], [2.17]).

    The scattering probability is independent of external forces.

    The duration of a collision is much shorter than the average time of motion of a particle;

    collisions are instantaneous.

    Carrier-carrier interaction is negligible. This effect would change the integrand of the right hand side integral in (2.3-10) highly nonlinear inj, [2.4].

    External forces are almost constant over a length comparable to the physical dimensions of the wave packet describing the motion of a carrier.

    nt n equals es (2.3-7).

    (2.3- 7)

    ,rces due to .ncies, and of internal j ics [2.49].

    !

    quantity .t

    e k will be ion term as

    (2.3-8)

    '

  • 14

    2 Some Fundamental Properties 2.3 Carrier Transport Equations

    The band theory and the effective mass theorem apply to the semiconductor under consideration [2.76].

    However, it is my intention here to outline the derivation of the classical current relat ions

    and only to pinpoint the problems associated with much more basic and error-prone

    models.

    ' By assuming that all scattering processes are elastic and by neglecting all effects~

    of .egeneracy the scattering integral can be approximated and the Boltzmann equation is reduced to a pure differential equation [2.21], [2.42], [2.76].

    ofv t ; d f, ~ d I. J.-fvo - + - gra k + u gra = - ---

    o t 1i v v x v . v

    ~ ofvdk=4 u. at

    Vk

    Vk

    (2.3-11

    )

    f e. (u. grad, f.)

    Vk The physical mot ivation for the right hand side of (2.3-11) is as follows: Suppose that

    at some moment of t ime t =0 all external forces are switched off and fv is homogeneous.

    ft T grad, fv + u. grad, J~ = 0

    Vk

    (2.3-12

    )

    v denotes the carrier concent

    mass, T denotes the lattice te

    average collision time. The c

    Boltzmann constant. The ext

    field E if the magnetic inducti also for the validity

    of (2.3-

    It follows from (2.3-11) that the distribution function will change as a result of

    collisions only. (2.3-13) will reduce to:

    at r; (2.3-13

    )

    I, (x, t; t) =L, (x, k)+ (f,. (x,/Z, O)-f.0 (x, k)) v>

    (2.3-14

    )

    In (2.3-18) it has been

    assurm compared to the

    thermal e carriers (cf. [2.67]).

    We obtain ordinary differen

    holes using the above given

    The solution of this differential equation is quite simple.

    J.0 is the equilibrium distribution function, and the quantity r, shows the rate of return to the state of equilibrium from the disturbed state, therefore, it is termed the relaxation

    time. Under the very restrictive assumptions stated above the problem of I solving the Boltzmann equation can be eased drastically by modeling the relaxation time as only a

    function of energy [2.42].

    l In order to obtain the current relations from (2.3-11) we multiply this equation with I the group velocity ii., and then we integrate the equation over momentum space.

    a q -(n-v)+-n ot n m:

    a q -(p-u)--.p ot p m

    p

    These equations can be reg;

    "closed solution" of these ~

    obtain an approximate soluti

    Vk Vk Vk

    = -f Uv_Jv-fvo. dk T .

    v

    (2.3-15

    )

    Vk

    For the solution of(2.3-15) we make use of the following four solutions to integrals,

    the verificat ion and discussion of which is not necessarily trivial, but well established

    in the literature, e.g. [2.13], [2.21], [2.71].

    We rewrite (2.3-21) and

    (2.3-'. collision times r, and

    chargewe end up with:

    !

    f

    l

    J

    I

  • 2.3 Carrier Transport Equations 15

    amental Properties

    semiconductor - ofv - 3 a - u .-.dk=4n -(vv) v at at v (2.3-16)

    :lassical current

    more basic and

    Vk

    .ng all effects~ of the Boltzmann

    [2.76].

    . ( F ve ) - F ve J . - 3 - V Uv h.gradkfv dk=----r; Uvgradkfvdk=-4n Fvem:

    Vk Vk

    (2.3-17)

    (2.3-11)

    Vk

    (2.3-18)

    s: Suppose that

    i off and fv is

    (2.3-19)

    Vk Vk

    (2.3-12)

    v denotes the carrier concentration, D y is the 9.!:ill, velocity, m: represents the effective mass, Tdenotes the lattice temperature and Ty in the right hand sides of (2.3-19) is an average

    collision time. The constant k on the right hand side of(2.3-18) denotes the Boltzmann

    constant. The external forces F,e can be expressed in terms of the electric field E if the magnetic induction B (Lorentz force) is neglected which is a requirement also for the validity of (2.3-16).

    e as a result of

    (2.3-13)

    (2.3-20)

    (2.3-14)

    In (2.3-18) it has been assumed that the drift energy of the carriers is negligibly small

    compared to the thermal energy. Therefore, relation (2.3-18) is invalid for hot carriers (cf.

    [2.67]).

    We obtain ordinary differential equations for the drift velocities of electrons and holes using

    the above given integrals and the force relations (2.3-20).

    a q - 1 n. vn -(n vn)+-. n E+-. grad(n k T)= ---

    3t mn mn Tn

    (2.3-21)

    ows the rate of ,

    it is termed the e the problem of I g the relaxation a _ q _ 1 P. vp

    - (p vp)- -. p E + -. . grad (p . k . 1) = - -- e t mp mp 'p

    These equations can be regarded also as macroscopic force balance equations. A "closed

    solution" of these equations is, unfortunately, not possible. In order to

    obtain an approximate solution we introduce effective carrier mobilities n and P.

    (2.3-22) ls equation with en tum space.

    (2.3-15)

    (2.3-23)

    (2.3-24)

    ms to integrals,

    well established

    We rewrite (2.3-21) and (2.3-22) after multiplication with the corresponding average

    collision times Tv and charge constant q, and - remembering (2.3-1) and (2.3-2) - we end up with:

    I

    I

    l

    l

  • 16

    2 Some Fundamental Properties 2.3 Carrier Transport Equations

    Then we can use the substitutions (2.3-34) and (2.3-35) which by means of physical

    interpretation are termed the .Einstein relations

    k . T Dn = 11 - (2.3-34)

    q

    k . T D = .--

    P p q (2.3-35)

    All scattering processes h: polar optical phonon scat

    has been neglected.

    The spatial variations oft This implies a slowly var

    path.

    Effects of degeneracy ha

    V( integral.

    The spatial variation oft

    varying electric field vect

    The influence of the Le induction.

    The time and spatial

    furthermore, lattice and c

    the diffusion of hot carrier

    overcome this problem b)

    [2.67], [2.68], [2.69], [2

    Parabolic energy bands

    degenerate semiconducto

    the realistic band structi

    [2.20]. However, for a re

    system of Boltzmann equ:

    one (cf. [2.96]).

    The zero order term of t

    collision time only has

    conductivity phenomena.

    The semiconductor has b distribution function is c

    vicinity of boundaries, for

    expected that the drift-di

    free paths of boundaries.

    In the literature we can fine

    different form of the current

    eguivalent to the drift-diffu

    slightly different relations v

    statistics that the equilibriu

    Fermi functions.

    (2.3-25)

    a JP _ (- 1 ( k . T)) -r . - + J = q. . p. E - - grad p --

    P 8t P p p q

    The average collision times -r v are very small, typically in the order of tenth of picoseconds.

    Therefore, equations (2.3-25) and (2.3-26) can be understood as being

    singular!)'. pertu~ This suggests to expand the solution into powers of the

    perturbation parameter which is the collision time.

    (2.3-26)

    a)

    Jn(-rn ) = I t: (-rn Y

    i=O

    (2.3-27)

    00

    Jp(,p)= I Jpi (rp)1 (2.3-28) i=O

    We have an algebraic equation for the zero order term of the current density.

    _ (- 1 ( k. T)) J 110 = q ,, n E + --;:;- grad n . -q-

    - (- 1 ( k. T)) J po= q P p E - p grad p -q-

    These equations are formally approximations of order rv.

    Jn = Jno + 0

    (Tn) Jp=]p0+0(rp)

    (l We further assume that the lattice temperature is constant; T=const.

    (2.3-29)

    (2.3-30)

    (2.3-31)

    (2.3-32)

    (2.3-33)

    to define the diffusion constants Dv , and, finally, we are able to write down the current

    relations in the well known, established form as sums of a drift and a diffusion component.

    1,/~:! q n 11 E + q D1 1 grad

    n Jr~ q P v E - q DP grad p

    (2.3-36)

    (2.3-37)

    ~ k)-

    i: x, - (

    l+exp

    In the following I should like to summarize the most important assumptions which

    had to be performed over and above to the ones necessary for the validity of the Boltzmann

    equation to obtain the current relations (2.3-36) and (2.3-37).

    fp0(X,k)=--l

    +exp (

    I

    (

    I

  • (2.3-31)

    (2.3-32)

    (2.3-33)

    ins of physical

    (2.3-34) 3" I

    (2.3-35)

    -rite down the

    a drift and a

    (2.3-36)

    (2.3-37)

    nptions which

    validity of the

    .3-37).

    nental Properties

    (2.3-25)

    (2.3-26)

    . er of tenth of

    stood as being

    powers of the

    (2.3-27)

    (2.3-28)

    en t density.

    (2.3-29)

    (2.3-30)

    2.3 Carrier Transport Equations

    All scattering processes have been assumed to be elastic. Therefore, for instance, polar optical phonon scattering, which is a major scattering mechanism in GaAs, has been neglected.

    The spatial variations of the collision time and the band structure are neglected. This implies a slowly varying impurity concentration over a carrier mean free path .

    Effects of degeneracy have been neglected in the approximation for the scattering integral.

    The spatial variation of the external forces is neglected which implies a slow!)". varying electric field vector.

    The influence of the Lorentz force is ignored by assuming zero magnetic induction.

    The time and spatial variation of carrier temperature is neglected and, furthermore, lattice and carrier temperature are assumed to be equal. Therefore, the diffusion of hot

    carriers is improperly described. Several authors have tried to overcome this problem by

    using modified Einstein relations [2.9], [2.52], [2.53], [2.67], [2.68], [2.69], [2.85].

    Parabolic energy bands are assumed which is an additional reason why degenerate

    semiconductor materials cannot be treated properly. Calculations of the realistic band

    structure of various semiconductors can be found in, e.g., [2.20]. However, for a

    realistic band structure it can become necessary to use a system of Boltzmann equations

    to describe the carrier distribution instead of just one (cf. [2.96]).

    The zero order term of the series expansions of Jn and JP into powers of the collision time only has been taken into account. Thus, all time dependent EOnductivity

    phenomena, like velocity overshoot, are not included.

    The semiconductor has been assumed to be infinitely large. In a real device the distribution function is changed in a complex, highly irregular manner in the vicinity of

    boundaries, for instance contacts [2.58] and interfaces [2.36]. It can be I expected that the drift-diffusion approximation fails within a few carrier mean free paths of boundaries.

    In the literature we can find quite a few papers and books whose authors use a different

    form of the current relations. These are based upon special assumptions equivalent to the

    drift-diffusion approximations. The procedure to derive these slightly different relations

    will be outlined next. We know from semiconductor statistics that the eguilibrium

    distribution functions for electrons and holes are Fermi functions.

    1 i: (x, I Z ) = (Ee (x, k)- E Jn (x))

    1 +exp k . T(x)

    1

    (2.3-38)

    fpo (x, k) = (E JP (x) - Ev (x, tZ)) 1 -l-exp k. T(x)

    (2.3-39)

    17

  • 18

    2 Some Fundamental Properties 2.3 Carrier Transport Equations

    Ee and E,, denote the conduction and the valence band energy, respectively.

    - h2.f.f EJx, k) = E,0 - q I/I (x) + (2.3-40)

    2 m: _ n2kk E0(:x,k)=E00-q l/J(x)- (2.3-41)

    2. m; Eco and E"0 are the conduction band and the valence band edge, respectively; I/I (x) is

    the electrostatic potential as defined in Section 2.1. -

    E Jn and E fp in (2.3-38) and (2.3-39) determine the Fermi energy for electrons and holes. We

    shall try now to calculate a correction term (2.3-42) to the equilibrium distribution function.

    By assuming vanishing vai

    relations (2.3-47) to (2.3-50) i

    for the distribution function

    fn ~fno - in fno (1

    . The current densities can no velocity and distribution ft

    i: 1. I, (x, t; t ) = i.; (x, k) + i.. (x, k, t)

    (2.3-42) J - -q f n - 4 . n3 . Un . J,

    We recall the Boltzmann equation with the relaxation time approximation:

    a I, five d f ~ d f I, - fvo (2 3 43) - + - gra k v + U v gra x v = - --- - ot n 'tv

    Vk

    By assuming

    J - q f : 4 . 7[3 . iip. j

    Vk

    (2.3-46)

    ifJn and

  • 2.3 Carrier Transport Equations 19

    ental Properties

    (2.3-40)

    By assuming vanishing vanat10n of temperature grad, T=O and substituting relations

    (2.3-47) to (2.3-50) into (2.3-46) we obtain expressions (2.3-51) and (2.3-52) for the

    distribution functions for electrons and holes, respectively.

    ectively.

    (2.3-41) (2.3-51)

    .ively; t/J (x) is ~ . av

    fv=fvo+

  • 20

    2 Some Fundamental Properties 2.3 Carrier Transport Equations

    _ ( k . T ( p )) JP= -q P p grad ijJ +-q-. In n;e

    Then we evaluate the "grad" operator and obtain relations (2.3-62) and (2.3-63) for the electron

    and hole current, respectively.

    (2.3-61) quasi-Fermi potential v

    mean free path.

    Only to first order is the

    Fermi potential. Away

    important.

    Various authors have publ.

    carrier transport in semicor out

    a one dimensional si rm [2.31].

    They use for the de

    OU OU ~ -=-U-+-

    Ot ox

    _ - (k. T ) J;= n n E +q Dn grad n-q ., n -q- gradln(n;e) (2.3-62)

    (2.3-63)

    The last term in these relations represents a current which is caused by a possible dependence on

    position of the intrinsic concentration. It thus accounts for variations in the balldgap, and it will

    describe the bandgap narrowing effect observed in heavily doped semiconductors (see also

    [2.91]). If we assume a constant intrinsic concentration we obviously do not have a gradient of

    the intrinsic concentration, and then relations (2.3-62) and (2.3-63) are indeed identical to

    (2.3-36) and (2.3-37). For practical purpose it is often useful to define effective fields for the

    drift current components of the electron and hole current density.

    OW ow -=-U-+q

    ot ox

    J =A . n . . E + q . D . grad n n n 11 11 (2.3-66)

    (2.3-67)

    v denotes the electron drift

    electron energy and w rep (2.3-68) is almost equivalei

    obtained during the deriva

    energy w has been assumer

    3. k. T m' W= "+-

    2

    Equation (2.3-69) is obtair

    energy E and then integrat

    procedure we had to go thrc of

    the order u. (u. u) one ha "displaced" Maxwellian [2

    _ _ k . T

    En=E--- grad ln(n;J q

    - - k. T EP = E + -- -grad In (nie) q

    (2.3-64)

    (2.3-65)

    By rewriting (2.3-62) and (2.3-63) we obtain a form of the current relations which is very similar

    to the classical drift-diffusion approximations but which can, as in our example, take into account

    positional variations of the band gap.

    It is worth noting that Boltzmann statistics for the carrier concentrations have been used in

    (2.3-66) and (2.3-67). With Fermi statistics for the carrier concentrations an equally simple form

    of the current relations compared to the classical drift-diffusion approximations is not derived so

    easily.

    In the following I would like to summarize some of the simplifying assumptions which became

    clear in the derivation of (2.3-51) and (2.3-52), and which, additionally have been implicitly

    used for the derivation of the drift diffusion relations (2.3-34) and (2.3-35) (see also [2.10]).

    Higher order derivatives of the quasi-Fermi potentials have been neglected (cf. (2.3-46)).

    This means that we transform a non-local solution of the Boltzmann equation into an

    approximate one depending only upon the local gradient of the quasi-Fermi potential.

    The dependence of the distribution function upon the gradient of the quasi-Fermi potential

    has been linearized. That means that the scale of length over which the

    - ( E f;.(x,k)~exp --

    If one uses such a model

    circumvents the assumption

    always in equilibrium wit]

    required to derive (2.3-69),

    [2.42], [2.95]), and the addi

    lead to many open question

    these types of models becom

    such an electron temperat1

    principles particle model has

    which is that simulation res

    upon classical current relati, A

    two. dimensional simulat

    [2 .16]' [2.22].

    1

  • amental Properties

    (2.3-61)

    and (2.3-63) for

    L (n;,)) (2.3-62) 50

    (n;e)) (2.3-63)

    :d b~ a possible

    ts for variations ect observed in

    instant intrinsic

    ; concentration,

    36) and (2.3-3 7).

    he drift current

    (2.3-64)

    (2.3-65)

    lations which is

    .h can, as in our

    (2.3-66)

    (2.3-67)

    .tions have been

    ncentrations an

    al drift-diffusion

    ng assumptions

    2), and which,

    : drift diffusion

    en neglected (cf.

    the Boltzmann

    l gradient of the

    the quasi-Fermi

    i over which the

    2.3 Carrier Transport Equations 21

    quasi-Fermi potential varies by k . T/q must be large compared to the carrier mean free

    path.

    Only to first order is the carrier transport driving force the gradient of the quasiFermi potential. Away from equilibrium the electric field vector will become important.

    Various authors have published approaches for a more sophisticated treatment of carrier

    transport in semiconductors. Froelich and Blakey, for instance, have carried out a one

    dimensional simulation using energy and momentum conservation laws [2.31]. They use

    for the description of electron transport:

    av av q. E 2 a ( ( m*. v2 )) v at= - v. a;:+----;;- - 3. m*. n . ax n. w- --2- - -:;::

    (2.3-68)

    aw =-V aw +qEV--2-,!_(nV(W- m*.v2))- w-w0 at ax 3. n ax 2

  • 22

    2 Some Fundamental Properties

    Thornber [2.88] has published a generalized current equation for the simulation of

    submicron devices by supplementing the drift and diffusion current components with

    so-called gradient,~ and relaxation current comQonents in order to include the most

    important features of velocity overshoot.

    oE oE) on on J0=q n v(E)+ W(E) - +B(E). - -q D(E) - -q. A(E). - ox ot ox ot

    (2.3- 72)

    Graphs of the gradient coefficient W(E), the rate coefficient B (E) and the relaxation

    coefficient A (E) have been presented by Thornber for electrons in silicon at room

    temperature. v (E) and D (E) are the well known terms for the drift velocity and the diffusion

    coefficient. In the classical drift-diffusion current approximations all terms except those two

    are assumed to be negligibly small. Thornber stated in his article [2.88] that relations of the

    form (2.3- 72) are adequate to represent current densities whenever the characteristic

    distances over which the particle concentration or electric field changes exceed 20 nm in

    silicon (200 nm in GaAs) and the characteristic time intervals of such changes exceed about

    0.4 picoseconds in silicon (2 picoseconds in GaAs). However, as far as I know, relations of

    the form (2.3- 72) - a generalization of which to higher dimensional form is supposed to be

    straightforward - have not been tested for practical applicability, although it can be

    speculated that the range of validity of such treatments is greatly extended.

    In recent work a new concept of device operation has been brought about, namely ballistic

    transport. It was argued that by properly selecting the material, temperature, geometry and

    bias, a device could be built which is much smaller than the mean free path between

    scattering events [2.41], [2.78]. This question, however, is still open, although quite a few

    activities in that field can be observed [2.3], [2.6], [2.7], [2.37], [2.45], [2.74], [2.78], [2.87],

    [2.93].

    As review papers on which kind of problems have to be faced especially for the simulation

    of miniaturized devices, references [2.10], [2.29], [2.30] can be suggested. Investigations on

    how the carrier transport equations, i.e. current relations, are changed, particularly by heavy

    doping effects, are presented in [2.59], [2.63], [2.70], [2.73], [2.90], [2.91]. In the next

    section we shall address these problems with emphasis on adequate models for the carrier

    concentrations. However, considering the current relations, throughout this text we shall

    favour current relations which have a structural appearing like (2.3-66) and (2.3-67). These

    formulations will allow us to best characterize, in a more mathematical sense, the problem

    of carrier transport. From a pragmatic physical point of view equations (2.3-66) and

    (2.3-67) offer, considering the state of the art in understanding their background, a

    sufficiently large set of parameters (effective mobilities "'effective fields E., effective diffusivities D,) to be invoked in order to reach a specific goal for agreement between

    results obtained by simulation and measurement. One must keep in mind that all these

    equations are just models in any case, which only imitatively simulate a real process, more

    or less accurately, in a qualitative and quantitative sense.

    2.4 Carrier Concentrations

    2.4 Carrier Conce:

    Accurate models for the

    necessity if qualitatively < obtained. I first shall revi

    concentrations, which gii

    approaches are certainly

    physics. However, I shall

    assumptions which are ver:

    keeps pace with the currei

    Assuming a parabolic and

    the conduction band (2.4-1

    given in the well known f<

    4 n (2 pcCE) h

    4 n (2

    Pv(E) = h3

    I have avoided in (2.4-1), (2

    - introducing a reference t

    differences are of relevanc

    literature because that pre

    seems to be very conveni:

    arbitrary reference point v Ee

    and Ev are the so-called J the width of the forbidden

    Eg=Ec-Ev

    Numerical values of the ba.

    frequently used materials

    Table 2.4-1.

    Table 2.4-1. Band gaps in undope

    material E9 [eV]

    Si 1.l2

    Ga As 1.35

    Ge 0.67

    For silicon the temperatu

    accurately with (2.4-4) or (

    7.0 Eg= 1.17eV- -

    I

    (

  • ndamental Properties 2.4 Carrier Concentrations 23

    the simulation of

    rrent components

    n order to include

    2.4 Carrier Concentrations

    In on --q. A(E)'x . ot

    (2.3-

    72)

    Accurate models for the carrier concentrations in semiconductor devices are .a necessity if

    qualitatively and quantitatively correct simulation results are to be obtained. I first shall

    review the "classical" approaches of modeling the carrier concentrations, which give fairly

    simple, closed form algebraic results. These approaches are certainly well documented in

    many books on semiconductor physics. However, I shall place particular emphasis on

    properly pointing out assumptions which are very possibly going to be violated when

    device sophistication keeps pace with the current trends.

    Assuming a parabolic and isotropic band structure the density of possible~ for the

    conduction band (2.4-1) and the valence band (2.4-2) as a function of energy E are given in

    the well known form with properly averaged effective masses [2.42]:

    4 . n (2 m:)312 Pc(E)= ,, VE-Ee

    (2.4-1)

    and the relaxation

    in silicon at room

    ft velocity and the

    imations all terms

    tated in his article

    t current densities

    concentration or

    the characteristic

    .on (2 picoseconds

    '. m (2.3-72) - a sd to

    be straightthough

    it can be ly

    extended.

    ;ht about, namely

    naterial, tempera-

    Iler than the mean

    1, however, is still

    [2.3], [2.6], [2.7],

    (2.4-2)

    I have avoided in (2.4-1 ), (2.4-2) - as I shall do for all other formulae in this section -

    introducing a reference energy with a specific absolute value, because only energy

    differences are of relevance. One should, however, be quite careful in reading the literature

    because that problem is absolutely not treated in a unique manner. It seems to be very

    convenient, a matter of taste, to some people to introduce an arbitrary reference point with

    zero energy on the energy scale. Ee and Ev are the so-called band edges. Their difference E9 denotes the band gap, i.e. the width of the forbid d en band between condu ctio n band and valen ce band.

    especially for the I, [2.30] can be ions,

    i.e. current

    resented in [2.59],

    iall address these

    trations. However,

    tll favour current d

    (2.3-67). These

    matical sense, the

    of view equations

    iderstanding their

    ilities .,

    effective a specific

    goal for ement.

    One must case,

    which only a

    qualitative and

    (2.4-3)

    Numerical values of the band gap and its linear temperature coefficient for the most

    frequently used materials semiconductor devices are made of are summarized in Table

    2.4-1.

    Table 2.4- l. Band gaps in undoped material at T = 300 K

    material E. [eV] dE./d T [eV K - t]

    Si 1.12 -2.7. 10-4

    Ga As 1.35 -5.0. 10-4

    Ge 0.67 -3.7. 10-4

    For silicon the temperature dependence of the band gap can be modeled more accurately with (2.4-4) or (2.4-5) [2.32], [2.34].

    7.02. 10-4eV. (-fJ E9=1.17eV- (T)

    1108+ - K

    (2.4-4)

  • 24 2 Some Fundamental Properties

    Eg=l.1785eV-9.025-10-5eV- (;)-3.05-10-7eV- (;)2 (2.4-5) (2.4-4) is the older formula; it has also been published with slightly different constants, e.g. [2.5], [2.33].

    The temperature dependence of the effective masses m: and m; of electrons and holes for the density of states in silicon can be modeled with polynomials which are fitted to

    experimental data [2.32], [2.34].

    m~ =m0 (t.045 +4.5 10-4 (;)) (2.4-6)

    m;=m0 (o.523+1.4.10-3. (;)-1.4s.10-6. (;)2) (2.4-7)

    Values for the effective masses at room temperature are summarized for some of the

    relevant semiconductor materials in Table 2.4-2.

    Table 2.4-2. Effective mass ratios at T = 300 K

    material m:fmo [ ] m;/m0 [ ]

    Si 1.18 0.5

    Ga As 0.068 0.5

    Ge 0.55 0.3

    In order to obtain expressions for the carrier concentrations we have to integrate the density

    of states function multiplied with the corresponding carrier distribution function over the

    energy space.

    00

    n= J pc(E). f,,(E). dE (2.4-8) E, E

    p= f Pv(E)f,,(E)-dE (2.4-9)

    - 00

    The lower integration bound in (2.4-8) is E e because no possible states for electrons do exist

    for energies below the conduction band edge. For a similar reason the upper bound in

    (2.4-9) is Ev. The distribution functions f,, (E) and fp (E) are Fermi functions.

    fn(E)= (E-Efn) 1 +exp k . T

    (2.4-10)

    1 (2.4-11)

    fp(E) = (E1p-E) 1 -l-exp k. T

    2.4 Carrier Concentrations

    E Jn and EfP denote the h exact meaning will be dis:

    evaluated to

    2 n=Nc Vn F112

    2 p=Nv. Vn. F1;2

    where N c and N" denote th

    the valence band, respecti

    Nc=2 cn: Nv=2 cn:

    F 112 (x) is the Fermi integ

    closed form solution.

    00

    F1;2(X)=J _t l+e

    0

    The asymptotic behavior'

    however, is analytic.

    F1;2 (x) ~ ~ x31; 3

    The qualitative behavior Fig.

    2.4-1. For arguments

    approximated with an ext

    Vn 2

    F1;2(x)~ c(x)+(

    Quite a few suggestions ha

    very crude approximation

    c (x) = 1/4, -1 <

    However, due to this simp.

    expressions where F 112 (x)

    been proposed and used i

  • 2.4 Carrier Concentrations 25 ital Properties

    r (2.4-5) E Jn and E I denote the Fermi energies for electrons and holes, respectively. Their exact meaning will be discussed later. The integrals in (2.4-8) and (2.4-9) can be evaluated to

    :ly different n=N ,_2_.F (Efn-Ec) c l c 1/2 k V tt . T

    (2.4-12)

    ectrons and

    ls which

    are

    (2.4-6)

    2 (Ev-Efv) p=Nv. -- . F1;2 --~~

    }In k.T where N c and N" denote the effective density of states in the conduction band and in the

    valence band, respectively.

    (2.4-13)

    (2.4- 7)

    Ne= 2. (2 n :~ T m: )3'2 (2.4-14) some of the

    Nv=2 cnI~~ T.m;y12

    (2.4-15)

    F112(x) is the Fermi integral of order 1L2 which, unfortunately, does not have a closed form

    solution.

    .ntegrate the

    distribution

    oo VY F x = -d 1/2 ( ) . - v y

    0

    The asymptotic behavior of F112 (x) for large negative and large positive argument,

    however, is analytic.

    (2.4-16)

    (2A-17)

    (2.4-8) ~ 2 3/2 F 112 (x) = - x , x 1

    3 (2.4-18)

    (2.4-9)

    The qualitative behavior of F112 (x) and its asymptotic expansions are

    shown in Fig.2.4-1. For arguments close to zero it has been shown

    that F112(x ) can be approximated with an expression of the following type:

    (2.4-10)

    Vn 2

    F (x)~--- 112 - c(x)+e-x Quite a few suggestions have been made in the literature for c (x). A most simple but very

    crude approximation reads:

    (2.4-19)

    for

    electrons on

    the upper ni

    functions.

    c(x)=l/4, -l

  • 2

    6

    2 Some Fundamental Properties 2.4 Carrier Concen tra ti ons

    10 l -:j {ri. x I

    -e I '7 2 I /' I I I

    ~ 10. ~

    I I

    I I

    /. I

    /, I

    LL b I 1....x31

    2 I 3 I

    . ,

    10-1

    However, this approach

    range of arguments. A

    inverse function is pre

    formulae with high ac

    Chebyshev approximati

    To come back to the ca

    (2.4-17) for the Fermi ii

    holds. The validity of t

    electrons is sufficiently ~

    energy for holes is suffici

    are equivalent to the use

    will then simplify to

    n=Ncexp(~

    p=Nv. exp e

    1 o-i

    -

    5

    5 x

    Fig. 2.4-1. Fermi-integral of order 1/2 and its asymptotic expansions

    0

    10

    15

    I

    c(x)=0.31-0.044.x, x

  • arnental Properties 2.4 Carrier Concentrations 27

    However, this approach will only deliver formulae which are valid for a restricted range of

    arguments. A review on approximations for Fermi integrals and their inverse function is

    presented in [2.12]. For the purpose of implementation of formulae with high accuracy on

    large computers it is better to use rational Chebyshev approximations as demonstrated in

    [2.19].

    To come back to the carrier concentrations, we can use the asymptotic expansion (2.4-17)

    for the Fermi integral in the expressions (2.4-12), (2.4-l3)j[

    Efn-Ec -1 k . T

    (2.4-25)

    Ev-Efp

  • 2

    8

    2 Some Fundamental Properties 2.4 Carrier Concentrations

  • ental Properties 2.4 Carrier Concen tra ti ons 29

    hey describe

    ibrium state.

    ations to the

    y

    differences :

    relevant for

    iitrarily. It is

    1e zero if the

    etowhichno :

    zero for the

    um. Thus,

    E, .ulated in

    the ng the

    above

    band and it is, therefore, well separated from both band edges. Thus Boltzmann statist ics

    for intrinsic semiconductors in equilibrium are usually valid.

    For many applications it is convenient to define a so-called intrinsic concentration n; as the

    geometric average of the carrier concentrations in a semiconductor in equilibrium.

    (2.4-40)

    The existence of dopants is allowed in (2.4-40). If Boltzmann statistics are valid for

    describing the carrier concentrations, n, is evaluated with small algebraic effort:

    n; = VNc N,, -exp (- 2. ~9- T) (2.4-41)

    (2.4-33)

    We see that n; is position dependent if the band gap E9 is position dependent. The carrier

    concentrat10ns can now be rewnften into the well known form with five parameters:

    intrinsic concentration n.; electrostatic potential t /; , quasi-Fermi potentials

  • 30

    2 Some Fundamental Properties 2.4 Carrier Concentrations

    (2.4-48)

    o E e of the conduction barn causes a shift o Ev of the va majority electrons also scree

    donor ions. As already s

    description of the interactio

    these subjects can be found i

    on the results which are esi

    derivation.

    Kane [2.48] bas derived a1 conduction and the valeno that the local potential fluct function can be defined as density of states functions states functions over the la

    + _ k. T (NZ) N n NA --> l jJ b ~ -- ln --

    q n,

    _ + . k T (NA_) NA Nn --> tj;b;;;; ---.In --

    q n;

    However, it is to note that the validity of Boltzmann statistics becomes a very poor

    assumption for high doping concentrations, because, as already mentioned, the Fermi

    energies E fm E f P are shifted towards one of the band edges. If the error introduced by the

    assumption of Boltzmann statistics is not acceptable, one has to solve (2.4-49) for the

    built-in potential.

    2 (Ev-Efp) 2 (Efn-Ec) + - Nv,cF112 -Nc,cF112 +Nn-NA=O vn k-T vn k-T

    (2.4-49)

    (2.4-47)

    Again, this can only be done with numerical methods. It is obvious that the sum of the

    intrinsic Fermi energy E; and built-in potential t f;b, which is often termed the extrinsic Fermi

    energy, can be calculated simultaneously from (2.4-49). Most semiconductor devices contain regions with doping levels above 1018 cm - 3 and the transport of carriers through these heavily doped parts can play an essential role in determining device behavior and performance. Therefore, the models for the carrier concentrations have to properly reflect the underlying physics of heavy doping effects. In the preceding considerations we have only addressed the problem of carrier statistics in this context. All of the possible problems associated with the density of states functions have been ignored, except that shift energies for the conduction and the valence band, which have been assumed to be parabolic, have been allowed. In the following we shall examine more in depth why and how the band structure is changed in heavily doped semiconductors. However, the statements we shall make have to some extent a speculative character, because, as it has to be said, our understanding of the physics of heavily doped semiconductors is fairly limited. The density of states function for electrons and holes is influenced by essentially two categories of phenomena [2.62]. The first category consists of interactions between carriers and between carriers and ionized impurity atoms. The second category comprises the effects of electrostatic potential fluctuations which account for the random distribution of impurities together with the overlap of the electron wave functions at the impurity states causing bandtails [2.48] and impurity bands [2.66]. While the second category of heavy doping phenomena alters the shape of the density of states functions for electrons and holes, the first category produces only rigid shifts of both the conduction and the valence bands towards each other. To give an example we shall discuss the possible carrier interaction phenomena in ntype silicon. For p-type material the facts are analogous. Inn-type semiconductors three phenomena become apparent: electron-donor interaction, electron-electron interaction and electron-hole interaction. The electron-donor interaction does not yield changes in the band edges, but the number of electrons in the semiconductor becomes so large that they screen the donor ions, which effectively reduces the impurity ionization energy so that the donor levels ultimately disappear into the conduction band (see also [2.62]). Electron-electron interaction yields a rigid shift

    Pc(E)= 4-n:-(2-1 h3

    Pv(E)= 4-n-(21

    with:

    x

    y(x)= iJ;. f V -0)

    A simple approximation fc

    Slotboom [2.81].

    x:::;0.601 y(x);;;;;

    x?:0.601

    Results which are fairly sim

    same time by Bonch-Bru:

    infinite tails for the condu-

    states functions are princi

    band, but they fall off rapid

    edge. As expected (2.4-5

    asymptotically equivalent

    (2.4-2), respectively. Amor

    functions has been carried

    are remarkably more comj

    ductors I NZ I;;;;; IN.4 I 0 superior to Kane's methoi

    {

  • ndarnental Properties

    (2.4-47)

    (2.4-48)

    omes a very poor

    y mentioned, the

    dges, If the error

    otable, one has to

    (2.4-49)

    JS that the sum of

    often termed the

    2.4-49).

    above 1018 cm-3 i

    play an essential

    the models for the

    physics of heavy

    essed the

    problem .sociated

    with the ergies for

    the conabolic,

    have been nd how

    the band the

    statements we sit

    has to be said, is

    fairly limited. by

    essentially two

    eractions between

    second category

    h account for

    the :he electron

    wave rity bands

    [2.66]. the shape

    of the >ry

    produces only Is

    each other. To

    phenomena in ne

    semiconductors

    electron-electron

    eraction does not

    he semiconductor

    ively reduces the

    isappear into the

    yields a rigid shift

    2.4 Carrier Concentrations

    Ee of the conduction band towards the valence band. Electron-hole interaction causes a shift o E; of the valence band towards the conduction band, because the majority electrons also screen the mobile minority holes in addition to the immobile donor ions. As already

    said, completely analogous statements hold for the description of the interaction

    phenomena in p-type material. An excellent review on these subjects can be found in [2.57].

    We shall primarily concentrate in the following on the results which are established without

    going very much into details of their derivation.

    Kane [2.48] has derived approximations for the density of states function for the

    conduction and the valence bands in heavily doped semiconductors by assuming that the

    local potential fluctuations are sufficiently slow that a local density of states function can

    be defined as if the local potential were constant. The macroscopic density of states

    functions which are the statistical average of the local density of states functions over the

    lattice can then be expressed as:

    _ 4.n.(2m~)312 .'C-. (E-Ec) Pc (E)- , ~ V a.; Y

    (Jcv (2.4-50)

    (2.4-51)

    with:

    x

    y(x)= J; f V x-u exp(-u2). du - 00

    (2.4-52)

    A simple approximation for the unwieldy equation (2.4-52) has been suggested by

    Slotboom [2.81].

    x~0.601 y(x)~

    x;;::0.601

    1 2. Vn e-x2 (1.225 -0.906. (1-ez"'))

    irx-(1--1 ) 16. x2

    (2.4-53)

    Results which are fairly similar to (2.4-50), (2.4-51) have been presented at almost the

    same time by Bonch-Bruevich [2.15]. These density of states functions include infinite tails

    for the conduction and the valence bands. That means the density of states functions are

    principally different from zero everywhere in the forbidden band, but they fall off rapidly

    with increasing distance from the corresponding band edge. As expected (2.4-50) and

    (2.4-51) are for small doping concentrations asymptotically equivalent to the parabolic

    density of states functions (2.4-1) and (2.4-2), respectively. A more rigorous approach to

    the derivation of density of states functions has been carried out by Halperin and Lax

    [2.38]. However, their results are remarkably more complex. For strongly compensated,

    heavily doped semiconductors IN; I~ I NA. I 0 only, the Halperin and Lax theory is expected to be superior to Kane's method (cf. [2.72]).

    31

    {

  • 32

    2 Some Fundamental Properties 2.4 Carrier Concentrations

    a.; is the characteristic standard deviation of the Gaussian tails of the density of states

    functions (2.4-50), (2.4-51). The best established model for O"cv has been published by

    Morgan [2.66].

    O"cv= q1 . v(Nii + N-:i_).). . exp (-_a_) e 4-n 2.J..

    (2.4-54)

    For the derivation of this fo valence band structures are r: (2.4-50), (2.4-51). Fermi stati accounted for in the calcula screening length in metals [2 that the Fermi energy lies i forbidden band. This does extraordinarily high which appropriate for semiconduct In Fig. 2.4-2 a comparison of given. The solid line correspc Mock, Polsky et al.); the da: infinite (the model of Slotbooi model ofVanOverstraeten et (2.4-56) as a reference.

    Jc denotes the screening length, and a is the crystal lattice constant, numerical values of which are summarized in Table 2.4-3.

    Table 2.4-3. Crystal lattice constants

    material a [10-9 m]

    Si GaAs Ge

    0.543072

    0.565315

    0.565754

    Kane [2.48] as well as Morgan [2.66] in their original work have used a so-called cutoff

    radius instead of a/2 in the exponential term of (2.4-54). However, there is evidence to

    relate this quantity to the lattice constant [2.66]. VanOverstraeten et al. [2.90] and Slotboom

    [2.81] have in their investigations fully neglected the exponential factor of (2.4-54). For the

    screening length Jc two models are most frequently in use. The first one has been proposed

    by Stern [2.84].

    E 0

    ~+ V1_an 1+1-ope l+-Ni_+N_-;;. oEfn oE!P k T;00

    For non-degenerate material when Boltzmann statistics can be applied this formula reduces

    to the well known Debye length.

    (2.4-55)

    en . . c. . , . .>

    O J c C J) 10-'

    (2.4-56)

    0

    )

    c

    c C J ) C J ) L 0 en

    Jc=~ -vt . k T

    q n+ P

    T;0n in (2.4-55) represents an effective temperature for ion screening. In the original paper

    of Stern [2.84] this quantity is treated as adjustable parameter in order to fit experimental

    data. Stern has speculated that at room temperature T;011 should be in the range from about 7000 K to 9000 K. Mock [2.63] and Polsky et al . [2.73] have used 9000 K in thei r work; Nakagawa

    [2.70] has claimed that 6000 K is more appropriate to obtain quant itatively correct results; and

    Slotboom [2.81] has assumed T;011 to be infinite so that the last term in the denominator of (2.4-55) vanishes. In [2.51] and [2.90] a different model for the screening length which has also been suggested

    by Stern [2.83] in an early work has been used.

    10-s ,--~- 10 17

    d

    Fig. 2.4-2. Screening length '

    (2.4-57)

    When the doping concentra

    described by a delta function

    the electrons of the impurir

    impurity band. Morgan [2.6( ~

    for the impurity band.

    J

  • il Properties

    density of

    has been

    (2.4-54)

    numerical

    t

    so-called .

    r, there is

    ieten et al.

    iected the

    are most

    (2.4-55)

    is formula

    (2.4-56)

    re original

    irder to fit

    ould be in

    ~.73]

    have (is

    more

    2.81] has

    J f (2.4-55)

    which has

    (2.4-57)

    2.4 Carrier Concentrations

    For the derivation of this formula it has been assumed that the conduction and valence band

    structures are parabolic so that (2.4-57) seems to be inconsistent with (2.4-50), (2.4-51).

    Fermi statistics for the carrier distribution functions have been accounted for in the

    calculation of (2.4-57) which can also be identified as the screening length in metals [2.50].

    A requirement for the applicability of (2.4-57) is that the Fermi energy lies in one of the

    carrier bands, and not as usual in the forbidden band. This does not happen unless the

    doping concentration is extraordinarily high which should lead to the conclusion that

    (2.4-57) is inappropriate for semiconductor device modeling.

    In Fig. 2.4-2 a comparison of the models for the screening length inn-type silicon is given.

    The solid line corresponds to (2.4-55) with T;0n equal to 9000 K (the model of Mock,

    Polsky et al.); the dashed line is also (2.4-55) but with T;0n assumed to be infinite (the

    model ofSlotboom); and the dot-dashed line corresponds to (2.4-57) (the model

    ofVanOverstraeten et al.). The dotted line denotes the classical Debye length (2.4-56) as a

    reference.

    silicon

    E o

    - en L

    ..., Ol c Q) 1

    o-> OJ c ,_ c Q) Q)

    L o en

    - -- - Mock Slolboom Uan Overslraelen el.al. De bye

    10 -9 ,---,--- 10 17

    1 0 18 1 0 IS 1 0 ZO 10 21

    donor concentration [cm-31

    Fig. 2.4-2. Screening length versus donor concentration in silicon at 300 K temperature

    When the doping concentration is large, the impurity energy level cannot be described by a

    delta function as it is the case in simple theory. The wave function of the electrons of the

    impurity atoms overlap, thus causing the formation of an impurity band. Morgan [2.66] has

    developed a theory which predicts a Gaussian sfiape for the impurity band.

    33

  • 34

    2 Some Fundamental Properties

    (2.4-59)

    2.4 Carrier Concentrations

    1C 23

    iu= j valence band

    - de I I 0 21 ::>

    Q)

    M I

    E

    l 0 20

    o - (I)

    Q)

    .., 10 19

    cu .., (I) c, 10 18

    0

    . .., -

    (/) I 0 17

    c

    Q)

    'O

    I 0 16

    I 0 IS

    -.7 -.6

    Fig. 2.4-3. Band structure f,

    (2.4-58)

    En and EA are the activation energies for specific donor and acceptor atoms, respectively.

    Numerical values for Ev and EA are well documented in the literature, e.g. [2.86]. The

    expression for uDA like equation (2.4-54) for ucu has also been proposed by Morgan [2.66].

    q2 v(Nj; +NA)). ( 1 ) a DA= - 1.0344 - exp - ~========= s 4.n Vn.3206n(Nj;+NA).).3

    (2.4-60)

    Some discussion about the models for avA can be found in, e.g., [2.39], [2.72].

    In order to obtain a density of states function for electrons and holes, the density of states

    functions of the conduction band (2.4-50), the valence band (2.4-51) and the impurity bands

    (2.4-58), (2.4-59) have to be combined. Kleppinger and Lindholm [2.51] have simply added

    up the corresponding functions for that purpose. VanOverstraeten et al. [2.30], however,

    have assumed that the total density of states function of the mobile carriers is composed of

    the envelope of the conduction (valence) density of states and the corresponding impurity

    band density of states function. This approach is physically much more sound since adding

    up the density of states functions implies that a substitute impurity atom and a silicon atom

    that were at that same lattice site both contribute to the density of states (cf. [2.72]). The

    concentration of electrons and holes can now, finally, be calculated by:

    (2.4-62)

    1023

    1022

    - 7 10 2t ::> Q)

    M I E 10 u

    - (IJ Q)

    .., 10'" cu

    .., (IJ

    .._ 10 18 0 ::Jl

    .., - 10 17 (/) c Q) 'O

    I 0"

    10 16 ,

    valence band

    I I 11111111

    -.7 -.6

    00

    (2.4-61)

    - 00

    00

    - 00

    The integration bounds are now - co and co in contrast to (2.4-8) and (2.4-9) because of the

    infinite tails of the density of states functions. It is obvious that the

    integrals (2.4-61), (2.4-62) do not have a closed form algebraic solution; they have to

    be solved with numerical methods. Details on how to design efficient algorithms for the

    self-consistent solution of the carrier concentrations and the built-in potential are given in,

    e.g., [2.46].

    Fig. 2.4-3, Fig. 2.4-4 and Fig. 2.4-5 summarize the results we have obtained in a graphical

    way. They show the density of states function for electrons max(pc(E),pD(E)) and the density

    of states function for holes max(pv(E),pA(E)). The dashed line in the conduction band

    denotes the distribution function of electrons, i.e. the integrand of (2.4-61). Fig. 2.4-3

    corresponds to a doping of : =1016cm-3, NA =0, i.e. fairly low doping concentration. Fig.2.4-4 has been

    Fig. 2.4-4. Band structure f(

    l

  • ne Fundamental Properties 2.4 Carrier Concentrations 35

    (2.4-58)

    10 2Z

    donor concentration

    valence band conduction band

    (2.4-59) - I ;:) I 0 21 [2.72]).

    , be calculated by:

    10 ,.

    10 ts -.

    6 -.

    5 -.4 .

    4

    .5 .6 . 7

    energy [et.JJ

    Fig. 2.4-3. Band structure for N't, = I016'cm - 3, N;, =0 in silicon at 300 K temperature

    - ~ 10" Q)

    1J

    1022 donor concentration

    (2.4-61)

    (2.4-62)

    I ~ JO"

    I 'll I

    E o I 0.,

    ~t to (2.4-8) and (2.4-9) is.

    It is obvious that the .ic

    solution; they have to 1 1

    efficient algorithms for

    nd the built-in potential

    "' OJ -;;; I 0'" - .>

    '- 0 10

    :n - .>

    10'"

    ; we have obtained in a

    function for electrons

    ioles max (P v (E),p A (E)). distribution function of

    sponds to a doping of tion.

    Fig. 2.4-4 has been

    valence band conduction band

    I I I I I I

    -.6 -.5 -.4 .

    4

    .

    s .6

    .7

    energy leUJ

    Fig. 2.4-4. Band structure for N't, = 1018 cm -3, N;, =0 in silicon at 300 K temperature

  • Fig. 2.4-5. Band structure for NZ= 1021 cm-3, N;; = 0 in silicon at 300 K temperature

    1023

    ~ acceptor concentration= 10 16cm-3

    10221

    donor concentration= 1016cm-3

    - .... I 10., ::> Q J i" 'l

    I E 10 20 0

    - (/)

    QJ

    .,.> 10" C D

    .,.> C f )

    (/) 10 17

    c QJ

    1J

    10,.

    10'" ~ ' ~~~~~~~' ~~~'d' I, l I ,l, 'I'' ~~~~~~~li'D'~ '~~~~ -.? -.6 -.5 -.4 .4 .5 .6 .

    7 energy [eUJ

    36

    10 23

    10 22

    - .... I ::

    > 10 21

    QJ i" 'l I

    E 0 l 0 20 - [I)

    QJ

    .,.> l 0 1 9 C D .,.>

    [I

    )

    (0

    L

    ..., c 10 12 (j)

    o c ~ 0 T=30( o

    ~ 10

    14 E

    o

    al (T)= -1.99765 10+1 +2.01814. 10-1. T-1.97040. 10-4. T

    2 (2.4-68)

    The coefficients a2 ( T) and a3 ( T) which fit the empirical formulae best to Mock's

    model in the doping range [1012, 1020]cm-3 are:

    a2(T)=9.60563. 10-1-3.94127 .10-3. T+4.41488. 10-6. T2

    a3(T)=l.29363 .10-1 +l.10709 .10-3. T-9.56981-10-7. T2

    U 10 II (/) c

    L

    ~ 1010

    with:

    10 s -t--- 10 17

    (2.4-69

    ) whereas for Slotboom's model in the doping range [1012, 3. 1020] cm -3 they read: Fig. 2.4-7. Intrinsic concentr

    (

    ' I

  • undamental Properties

    2.61], have proved

    of bipolar devices,

    oach with (2.4-63),

    t the concept of an

    tal impurity con)f

    compensation is

    erial any

    approach . s to note

    that for rom

    equilibrium is

    eason can be found

    I) compared to the

    concentration has

    (2.4-66)

    'imental data up to

    ons the theoretical

    o fit experimental

    in [2.11].

    :asured band gaps

    for heavily doped

    inly the rigid shifts

    to lattice disorder,

    h contributions to

    ive electrical band

    gations.

    omplex models

    of ] which have

    been ~ structure

    for the

    :m-3

    (2.4-67)

    . 10-4. T2 (2.4-68)

    ae best to Mack's

    ~ o-6. T2 .0-1.

    T2 (2.4-69

    ) ] cm -3 they read:

    2.4 Carrier Concentrations

    a2(T)=7.95811. 10-1-3.20439. 10-3 . T+3.54153. 10-6. T2 a3

    (1)=2.97104. 10-1 +6.75707. 10-4. T-4.90892. 10-7. T2

    (2.4-70)

    and for VanOverstraeten's et al. model in the doping range [ 101 7, 1021] cm - 3 they

    evaluate to:

    a2(T)=2.38838 .10-1-9.57814. 10-4 . T+ 1.07551. 10-6. T2 a3 ( T) =

    5.10190. 10-1+5.75190. 10-4. T- 7.01029. 10-1. T2

    (2.4-71)

    The temperature T has to be given in Kelvin in (2.4-68) to (2.4- 71 ). The maximum

    relative difference of formula (2.4-67) with the above given coefficients and the exactly

    evaluated models is always smaller than ten percent (cf. [2.46], [2.47]) in the temperature

    range [250, 400] K. Formulae for an effective intrinsic concentration for compensated

    material are also given in [2.46], however, they are much more complicated.

    In Fig. 2.4-7 the effective intrinsic concentration for Mock's model (solid line), Slotboom's

    model (dashed line) and VanOverstraeten's et al. model (dot dashed line) are shown in

    conjunction with the experimental values of Mertens et al. [2.61], Slotboom [2.81], Wieder

    [2.92] and Wulms [2.94]. Although the agreement between the models and the

    experimental data is not overwhelming, it can be considered pragmatically to be

    quite.good, because of the fairly pronounced scatter

    of the measured data. However, a judgment as to which of the models is to be prefered can

    not, therefore, be given.

    10 15

    ~

    Moel< I t

  • 40 2 Some Fundamental Properties 2.6 The Basic Semiconductor Eq1

    2.5 Heat Flow Equation J =qn- .E+1 II n

    For the design of power devices it is often desired to simulate interaction of electrothermal

    phenomena. Changes in the temperature and its distribution in the interior of a device can

    influence significantly the electrical device behavior. Particularly, two effects usually have

    to be considered. Thermal runawa:t is one, a rather common mechanism where the

    electrical energy dissipated causes a temperature rise over an extended area of a device

    resulting in increased power dissipation. The device temperature increases which leads to

    an irrecoverable device failure (burn out), unless an equilibrium situation can occur with a

    heat sink removing all of the energy dissipated. The existence of such an equilibrium

    situation is the second question which is sometimes quite difficult to answer [2.54]. In order

    to account for thermal effects in semiconductor devices tqe heat flow equation (2.5-1) has

    to be solved.

    JP= q. p - p. EThe

    last expression in (2.5-'.

    temperature field as the dr

    being constant for the deriv.

    As can be proved with m

    perature in (2.3-29), (2.3-30)

    verified these relations witl

    solution of the Boltzmann e1

    the thermal diffusion coeffi

    ar p. c. - -H =div k(T). grad T at (2.5-1)

    D T p

    DP~ 2. T

    material c[m2s-

    2K-

    1] p [VAs

    3m-

    5]

    Si 703 2328

    Si02 782 2650 typical

    Si3N4 787 3440 typical

    GaAs 351 5316 Ge 322 5323

    These coefficients are small

    the procedure just sketchec

    result owing to the comple

    obtain and discrepancies

    demonstrated that Stratton

    presence of dopands the th e

    factor of five. Some more cc

    [2.76]. However, one need 1

    the con text of semiconduct:

    these relatively rough mod

    on the current densities, e.

    p and c are the specific mass density and specific heat of the material. Numerical values for

    p and cat room temperature are summarized in Table 2.5-1 for the most frequently used

    materials in device processing.

    I Table 2.5-L Specific heat and density constants al T= 300 K

    The temperature dependence of p and c can be assumed to be negligibly small in

    consideration of practical device applications [2.50]. If one is not interested in thermal

    transients one can assume for the simulation that the partial derivative of the temperature

    with respect to time vanishes, which eases the problem of solving the heat flow equation by

    one dimension. However, one is absolutely incorrect in using this assumption in a

    simulation for which an equilibrium condition does not exist. The simulation program will

    "blow up" in a manner analogous to the real device.

    k ( T) and H denote the thermal conductivity and the locally generated heat. Models for these

    quantities will be given and discussed in Section 4.3 and Section 4.4, respectively.

    To just calculate the temperature distribution and the associated thermal power dissipation

    without taking into account the current induced by gradients of the temperature is a fairly

    crude approach which is only appropriate for limited application [2.35]. In a more rigorous

    approach the current density equations have to be supplemented by additional terms.

    2.6 The Basic Sem

    We shall now summarize tr

    in order to be able to wri

    equations, which we shall 1

    sake of transparency and e

    and complexity of our mod

    major number of engim

    Certainly, conditions do ex

    doubt. However, as I tried 1

    results in semiconductor

    applicable and still sufficiei

    The basic semiconductor

    continuity equations for el

    for electrons (2.6-4) and hr

    this set the heat flow equ:

  • nental Properties

    interaction of

    ibution in the

    rice behavior.

    iway is one, a

    ted causes a

    reased power

    verable device

    1 a heat sink

    rium situation

    4]. In order to

    ~(2.5-1)

    (2.5-1)

    al. Numerical

    1 for the most

    gibly small in interested in

    I derivative of em of solving

    y incorrect in

    ition does not

    us to the real

    heat. Models j

    Section 4.4,

    iermal

    power .dients

    of the e for

    limited

    [uations have

    2.6 The Basic Semiconductor Equations 41

    - - '/' 111 = q n 1 1 E + q D11 grad n + q n D11 grad T

    - - T J p = q . p . v . E - q D p grad p - q p . D p . grad T

    (2.5-2

    )

    (2.5-3

    )

    The last expression in (2.5-2), (2.5-3) represents a drift current component with the

    ~mperature field as the driving force. In Section 2.3 we did assume temperature being

    constant for the derivation of the classical drift-diffusion relations (cf. (2.3-33)). As can be

    proved with minor algebraic effort, by assuming non-constant temperature in (2.3-29),

    (2.3-30) we obtain equations (2.5-2), (2.5-3). Stratton [2.85] has verified these relations

    with a much more rigorous approach, from a perturbation solution of the Boltzmann

    equation. He also derived in his paper approximations for the thermal diffusion

    coefficientsn;'.

    DT~~ - ~ f (2.5-4)

    II - 2. T - l..~ ..

    D '-

    DT~_P -:;: - r-~ (2.5-5) p-2.r 11 These coefficients are smaller by a factor of two compared to those we obtain with the

    procedure just sketched above. However, as pointed out in [2.85] a more exact result

    owing to the complexity of the problem is cumbersome, if at all possible, to obtain and

    discrepancies of that order are not at all surprising. Dorkel [2.27] demonstrated that

    Stratton's result is applicable for intrinsic semiconductors; in the presence of dopands the

    thermal diffusion coefficient is underestimated by at most a factor offive. Some more

    considerations on this subject can be found in, e.g. [2.14], [2.76]. However. one need not

    worry as all publications on non-isothermal effects in the context of semiconductor device

    modeling certify more or less the applicability of these relatively rough models for

    describing the feedback of temperature gradients on the current densities, e.g. [2.1], [2.18].

    2.6 The Basic Semiconductor Equations

    We shall now summarize the results which we have obtained in the previous sections in order to be able to write down a set of equations, the "basic" semiconductor equations,

    which we shall use in all further investigations. It is obvious that for the sake of transparency and efficiency, we shall perform a trade-off between accuracy and complexity of our model. The equations we shall concentrate on are valid for the major number of engineering applications, particularly for silicon devices. Certainly, conditions do exist for which their validity is not guaranteed, or at least in doubt. However, as I tried to express in the previous sections, the more sophisticated results in semiconductor physics are too complex to give a rigorous, generally applicable and still sufficiently simple model for the purpose of device simulation. The basic semiconductor equations consist of Poisson's equatjon (2.6-1), the continuity equations for electrons (2.6-2) and holes (2.6-3) and the current relations for electrons (2.6-4) and holes (2.6-5). For some applications it is desired to add to this set the beat flow equations (2.6-6).

  • 42

    2 Some Fundamental Properties

    div grad t/J= .!!:__. (n-p-C) (2.6-1)

    - op div] +q -= -q. R p at

    (2.6-3)

    2. 7 References

    [2.12] Blakemore, J. S.: Appro Used

    to Describe Eleen 1067- 1076

    (1982). [2.13] Blatt, F. J.: Physics

    of E

    [2.14) Bletekjaer, K.: Transpon Electron Devices ED-17, [2.15] Bonch-Bruevich, V. L.: C State4, No.10, 1953-1' [2.16) Buot, F. A., Frey, J.: Et Design Information. Sol; [2.17] Capasso, F., Pearsall, T. Carlo Simulations of Higl EDL-2, 295 (1981). [2.18] Chryssafis, A., Love, W.: in n-p-n Power Transistc [2.19] Cody, W. J., Thacher, H. Orders - 1 /2, I /2 and 3 / [2.20] Cohen, M. L., Bergstres Fourteen

    Semiconductor 789- 796 (1966). [2.21] Conwell, E. M.: High F [2.22) Cook, R. K., Frey, J.: 1 in Si and GaAs MES (1982). [2.23] Curtice, W. R.: Direct C.

    (Monte-Carlo) Model fo 1942-1943 (1982). [2.24] DeMan, H. J. J.: The J Transistor. IEEE Trans. [2.25] DelA!amo, J. A., Swan Narrowing in Moderate! [2.26] Dhariwal, S. R.,

    Ojha, ' Electron. 25, No. 9, 909 [2.27] Dorkel, J. M.: On El( Electron. 26, No. 8, 819 [2.28] Engl, W. L., Dirks, H. K {1983).

    [2.29] Frey, J.: Transport Pb

    Semiconductor Devices : [2.30]

    Frey, J.: Physics Preble

    Semiconductor Devices : [2.31]

    Froelich, R. K., Blakey, Wave lmpatt Diodes. Pr [2.32) Gaensslen,

    F. H., Jaege Depletion Mode Devio

    Meeting, pp. 520- 524 ( [2.33)

    Gaensslen, F. H., Rideoi

    Temperature Operation. [2.34)

    Gaensslen, F. H., Jaeger Mode

    MOSFET's. Solid [2.35) Gaur, S. P.,

    Navon, D.

    Nonisothermal Conditio [2.36) Gnadinger, A. P., Talley, and the Thickness of the l 13' 1301 -1309 (1970). [2.37) Grondin, R. 0., Lugli, P Device Lett. EDL-3, No

    - on div J -q. - = q. R n ot

    (2.6-2)

    Jn=q n 11 E11+q D11 gradn

    JP=q. p . P. EP-q. DP. gradp ar p . c . - - H =div k ( T) . grad T ot

    (2.6-4)

    (2.6-5)

    (2.6-6)

    I

    To almost this level of completeness, these equations were first presented by

    VanRoosbroeck [2.89].

    Models for C, the net doping concentration, for R, the net generation/ recombination,

    for 1 1, P, the carrier mobilities, for H, the thermal generation and fork ( T), the thermal

    conductivity will be d iscussed in the following chapters. E,, and EP, the effective fields in the current relations are to first order the electric field, however, we may use

    supplementary correction terms to account for heavy do12ing (cf. Section 2.3, Section

    2.4) o r thermally induced currents (cf. Section 2.5). For such mathemat ical

    investigations, relat ively slight perturbations are of only secondary importance. Hence,

    for most applicat ions, accounting for some specific effect is possible by properly

    modeling the parameters in the basic equations.

    J

    2. 7 References

    [2.1] Adler, M. S.: Accurate Calculations of the Forward Drop and Power Dissipation in Thyristors.

    IEEE Trans. Electron Devices ED-25, No. I, 16-22 (1978). [2.2] Adler, M. S.: An Operational Method to Model Carrier Degeneracy and Band Gap Narrowing.

    Solid-State Electron. 26, No. 5, 387-396 (1983). [2.3] Ankri, D., Eastman, L. F.: GaA!As-GaAs Ballistic Hetero-Junction Bipolar Transistor.

    Electronics Lett. 18, No.17, 750-751 (1982). [2.4] Anselm, A. I.: Einfiihrung in die Halbleitertheorie. Berlin: Akademie-Verlag 1964.

    [2.5] Antoniadis, D. A., Dutton, R. W.: Models for Computer Simulation of Complete IC Fabrication Process.

    IEEE J. Solid State Circuits SC-14, No. 2, 412-422 (1979).

    [2.6] Awano, Y., Tomizawa, K., Hashizume, N., Kawashima, M.: Monte Carlo Particle Simulation of a GaAs

    Submicron n+-i-n+ Diode. Electronics Lett. 18, No. 3, 133-135 (1982).

    [2. 7] Awano, Y., Tomizawa, K., Hashizume, N., Kawashima, M.: Monte Carlo Particle Simulation of a GaAs

    Short-Channel MESFET. Electronics Lett. 19. No. I, 20-21 (1983).

    [2.8] Aymerich-Humett, X., Serra-Mestres, F., Millan, J.: An Analytical Approximation for the Fermi-Dirac

    Integral F3/2(x). Solid-State Electron. 24, No. 10, 981-982 (1981).

    [2.9] Baccarani, G., Mazzone, A. M.: On the Diffusion Current in Heavily Doped Silicon. Solid-State Electron.

    18, 469-470 (1975).

    [2.10] Baccarani, G.: Physics of Submicron Devices. Proc. VLSI Process and Device Modeling, pp. 1-23.

    Katholieke Universiteit Leuven, 1983.

    [2.11] Bennett, H. S.: Improved Concepts for Predicting the Electrical Behavior of Bipolar Structures in Silicon.

    IEEE Trans. Electron Devices ED-30, No. 8, 920-927 (1983).