Download as pdf or txt
Download as pdf or txt
You are on page 1of 65

p-n junctions

• Intuitive description.
– What are p-n junctions?
p-n junctions are formed by starting with a Si wafer (or ’substrate’) of a given type (say: B-doped p-type,
to fix the ideas) and ‘diffusing’ or ‘implanting’ impurities of opposite type (say: n-type, as from a gas source
of P – such as phosphine – or implanting As ions) in a region of the wafer. At the edge of the diffused (or
implanted) area there will be a ‘junction’ in which the p-type and the n-type semiconductor will be in direct
contact.
– What happens to the junction at equilibrium?
Consider the idealized situation in which we take an n-type Si crystal and a p-type Si crystal and bring
them together, while keeping them ‘grounded’, that is, attached to ‘contacts’ at zero voltage. At first, the
conduction and valence band edges will line up, while the Fermi level will exhibit a discontinuity at the junction.
But now electrons are free to diffuse from the n-region to the p-region, ‘pushed’ by the diffusion term Dn ∇n
in the DDE. Similarly, holes will be free to diffuse to the n region. As these diffusion processes happen, the
concentration of extra electrons in the p-region will build up, as well as the density of extra holes in the n
region. These charges will grow until they will build an electric field which will balance and stop the diffusive
flow of carriers. Statistical mechanics demands that at equilibrium the Fermi level of the system is unique and
constant. Therefore, the band-edges will ‘bend’ acquiring a spatial dependence. This is illustrated in the left
frame of the figure in the next page.
Note:
1. Deep in the n-type region to the right and in the p-type region to the left the semiconductor remains
almost neutral: The contacts have provided the carriers ‘lost’ during the diffusion mentioned above, so that
n = ND in the ‘quasi-neutral n region and p = NA in the quasi-neutral p region.
2. There is a central region which is ‘depleted’ of carriers: Electrons have left the region 0 ≤ x ≤ ln , holes
have left the region −lp ≤ x < 0, so that for lp ≤ x ≤ ln we have np < n2i . This is called the
‘transition region’ or, more often, the ‘depletion region’ of the junction.

ECE609 Spring 2010 105


3. The voltage ‘barrier’ built by the difusion of carriers upon putting the n and p regions in contact with each
other is called the ‘built-in’ potential, Vbi (denoted by Φ0 by Colinge and Colinge). It is easy to calculate
it, since it will be given by the difference between the equilibrium Fermi levels in the the two regions:
     
ND NA ND NA
Vbi = EF n0 − EF p0 = Ei + kB T ln − Ei + kB T ln = kB T ln .
ni ni n2i
(358)

ECE609 Spring 2010 106


– What happens to a biased junction?
Let’s now apply a bias Va to the junction. We consider Va positive when positive bias is applied to the
p-region, as illustrated in the figure. If Va > 0 (forward bias), the field in the depletion region will be
reduced, so that it will not balance anymore the diffusion current of electrons flowing to the left. Thus, electron
supplied from the contact at the extreme right will replenish those electrons entering the p-type region. This
will result in a current density J3 . Having entered the p region, electrons will eventually recombine with holes.
The contact at left will provide the holes necessary for this recombination process, giving rise to a component
J4 of the total current density. A similar sequence of events will happen to holes: Some will diffuse to the
n region (yielding the component J1 of the current density) recombining there with electrons provided by a
current density J2 from the right-contact.
If, instead, we apply a negative Va (reverse bias), the field in the depletion region will increase and the
associated drift current will be larger than the diffusion current. However, the flow of electrons from the p
region will be negligibly small, since there are very few electrons in p-doped Si. Similarly for holes in the n
region. Therefore, the reverse current will be very small. This shows that p-n junctions behave like diodes,
rectifying the current flow.
• Equilibrium.
Let’s consider the band-bending and carrier densities at equilibrium.
– Poisson equation. First, the Poisson equation describing the band bending in the depletion rgion is:

d2 φ(x) e
= − (p − n + ND − NA ) . (359)
dx2 s
This is a nonlinear equation since the carrier densities, p and n, depend on the potential φ(x) itself:
 
EF,n0 −Ei,n0 +eφ(x)
n(x) = ni exp k T = nn0 eeφ(x)/(kB T )
B
  , (360)
Ei,p0 −EF,n0 −eφ(x)
p(x) = ni exp kB T = pp0 e−eφ(x)/(kB T )

ECE609 Spring 2010 107


where nn0 and pp0 are the electron and hole densities in the quasi-neutral n and p regions, respectively.
– Depletion approximation. We can simplify Poisson equation, Eq. (359), by employing the ‘depletion
approximation’: Let’s assume that the electric field vanishes outside the depletion region and let’s also ignore
+
the charge due to free carriers in the depletion region lp ≤ x ≤ ln (indeed we will have ND >> n for

0 ≤ x ≤ ln and NA >> p for −lp ≤ x < 0, since we have depletion of free carriers in these regions), so
that:
d2 φ(x) eND
= − for 0 ≤ x ≤ ln , (361)
dx2 s
d2 φ(x) eNA
= for − lp ≤ x < 0 , (362)
dx2 s
so that, using the boundary conditions
 
dφ  dφ 
= = 0,
dx  x=−lp dx  x=ln

(expressing the fact that the electric field vanishes at the edges of the depletion region) and φ(−lp ) = φp0 ,
φ(x = ln ) = φn0 , expressing the fact that at the edges of the depletion region the carrier concentration
approaches the concentration in the quasi-neutral regions, we have:



eND 2
 − 2s (x − ln ) + φn0 (0 ≤ x ≤ ln )
φ(x) = (363)


 eNA 2
2s (x + lp ) + φp0 (−lp ≤ x ≤ 0) .

Note that the maximum electric field occurs at x = 0:

eNA eND
Fmax = lp = ln , (364)
s s

ECE609 Spring 2010 108


where the last equality follows from charge neutrality which requires

NA lp = ND ln . (365)

– Depletion width. To calculate the width of the depletion region, note that

Vbi = eφn0 − eφp0 . (366)

Moreover, the continuity of the potential at x = 0 implies, from Eq. (363):

eNA 2 eND 2
0 = φ(0−) − φ(0+) = φp0 − φn0 + lp + l . (367)
2s 2s n

Using now the charge-neutrality condition, Eq. (365), this becomes

2
eNA ND 2 eND 2
0 = φp0 − φn0 + l
2 n
+ l . (368)
2s NA 2s n

Inserting into this equation the expression for φp0 − φn0 from Eq. (366) and using Vbi from Eq. (358), we
have:  
NA ND e2 ND 2
Vbi = kB T ln = (N A + N D ) l , (369)
n2i
n
2s NA
so that:   1/2
2s NA Vbi
ln = . (370)
e2 ND (NA + ND )
Similarly:
  1/2
2s ND Vbi
lp = . (371)
e2 NA (NA + ND )

ECE609 Spring 2010 109


– Asymmetric junction. In the simpler case of a very highly asymmetric junction (for example: NA = 1015
cm−3 and ND = 1019 cm−3 , we can ignore NA with respect to ND in Eqns. (370) and (371) above, so
that:
    1/2
2s Vbi 1/2 2s NA Vbi
lp ≈ , ln ≈ 2
<< lp . (372)
e2 NA e2 ND
So, the heavily-doped n-region exhibits essentially no depletion. This is a general property: The larger the
concentration of free carriers, the smaller the voltage which can drop in the region. In the limit of a metal, we
know that no electric field can be sustained, because any voltage drop will be effectively screened by the large
concentration of free carriers.
• Off equilibrium: Shockley’s equation.
Let’s apply a bias Va to the junction. We shall first consider the ideal case of no generation/recombination in
the depletion region (ideal diode). We shall later consider deviations from this ideal condition.
– Ideal diode: No generation-recombination in depletion region.
Let’s make the following simplifying assumptions:
1. The concentrations of free carriers injected into the quasi-neutral regions are small enough so that we can
neglect their charge compared to the charge of the majority carriers when solving Poisson equation (low-level
injection).
2. The concentration of free carriers everywhere is small enough so that we can use Maxwell-Boltznmann
statistics (that is, the high-T limit) instead of the full Fermi-Dirac statistics.
3. The quasi-neutral regions are infinitely long.
4. Finally, there’s no electric field in the quasi-neutral regions, so that only diffusion controls the current-flow
in these regions.
Also, the calculation of all of the components of the current density J1 through J4 in the figure (right frame
at page 106) is difficult. However, we know that J2 = J1 and J3 = J4 , so we need to calculate only the
diffusion currents in the quasi-neutral regions J1 and J3 . Moreover, it will be convenient to compute J1 (that
is, the hole current Jp before it starts decreasing (due to recombination with electrons in the n region) at
x = ln and J3 = Jn at x = −lp (for the same reason).

ECE609 Spring 2010 110


First of all, we can follow again the same procedure we have followed above to obtain the width of the
depletion region simply replacing the built-in potential Vbi with its value modified by the applied bias,
Vbi − Va , obtaining:
 
2s NA (Vbi − Va ) 1/2
ln = . (373)
e2 ND (NA + ND )
Similarly:
 
2s ND (Vbi − Va ) 1/2
lp = , (374)
e2 NA (NA + ND )

Thus, the depletion region shrinks under forward bias (Va > 0) but grows under reverse bias (Va < 0).
From Eq. (360) and from the right frame of the figure at page 106 we see that for the concentrations of the
free minority carriers which spill-over (electrons spilling over into the p regions and holes spilling over into the
n region) we have (using assumption 2 above):


−e(Vbi −Va )/(kB T )

 n(x = −lp ) = nn0 e = np0 eeVa /(kB T )
(375)

 p(x = ln ) = pn0 eeVa /(kB T ) ,

so that the excess carriers at the edges of the depletion region are:


 eVa /(kB T )
 δn(−lp ) = np0 e
 − 1
(376)


 δp(ln ) = pn0 eeVa /(kB T ) − 1 ,

ECE609 Spring 2010 111


Using now assumption 4, the current in the quasi-neutral regions will be:

 dp
 Jp = −eDp dx (x > ln )
(377)

 J = eD dn
n n dx (x < −lp ) .

The continuity equations for the hole current in the quasi-neutral n region (the component J1 in the figure
for x > ln ) and for the electron current in the quasi-neutral p region will be:


 ∂pn
= − 1 ∂Jp − pn −pn0 (x > ln )
 ∂t e ∂x τp
(378)

 −n
 ∂np = 1 ∂Jn − p p0 n
(x < −l ) ,
∂t e ∂x τn p

where τp and τn are the (recombination) lifetimes of the minority holes and electrons in the quasi-neutral
regions. Combining Eq. (377) and Eq. (378) we get:


 ∂pn ∂ 2 pn pn −pn0

 ∂t = D p 2 − τp (x > ln )
∂x
(379)

 2
 ∂np = Dn ∂ np − np −np0

(x < −lp ) .
∂t ∂x2 τn

At steady state, the general solutions are:



 −x/Lp
 pn (x) = pn0 + A e + B ex/Lp (x > ln )
(380)

 n (x) = n −x/Ln
p p0 + C e + D ex/Ln (x < −lp ) ,

where Lp = (Dp τp )1/2 and Ln = (Dn τn )1/2 are the hole and electron diffusion lengths in the quasi-
neutral regions and A, B , C , and D are integration constants to be determined by the boundary conditions.

ECE609 Spring 2010 112


The first of these boundary conditions is determined by the concentrations of the minority carriers far away in
the quasi-neutral regions: pn (x → ∞) = pn0 and np (x → −∞) = np0 , which implies B = C = 0.
Another boundary condition results from the requirement that values of pn and np at the edges of the
depletion region match the values pn (x = ln ) and np (x = −lp ) given by Eq. (376) above, so that:

eVa /(kB T )

 pn (x) = pn0 + pn0 e
 − 1 e−(x−ln )/Lp (x > ln )
(381)


 np (x) = np0 + np0 eeVa /(kB T ) − 1 e(x+lp )/Ln (x < −lp ) .

Finally, by Eq. (377):



 eDp pn0 eVa /(kB T ) −(x−ln )/Lp

 J p (x) = L e − 1 e (x > ln )
p
(382)


 J (x) = eDn np0 eeVa /(kB T ) − 1 e(x+lp )/Ln (x < −lp ) .
n L n

It is important to note that these current densities vary with x because, for example, as Jp (x) decreases as
x increases, the hole current it represents (J1 in the figure at page 106) is transformed – via recombination
processes – into an electron current (J2 in that figure). So, the total current will be independent of x. This
constant value can be obtained thanks to our assumption that there’s no generation/recombination in the
depletion region. In this case we can evaluate the currents in Eq. (382) at a position in which they have not
yet started to ‘decay’ due to recombination, so Jp at x = ln and Jn at x = −lp . Thefore the total current
will be:
 
eDp pn0 eDn np0 eVa /(k T )
eVa /(kB T )

Jtotal = Jp (ln ) + Jn (−lp ) = + e B − 1 = Js e −1 ,
Lp Ln
(383)
where Js is the ‘saturation current density’ (the maximum current density we have under reverse bias, that is,
for Va → −∞). Equation (383) is known as ‘Shockley’s equation’.

ECE609 Spring 2010 113


– Deviation from ideality: Generation/recombination in depletion region.
Let’s now consider the more general case in which GR processes are active in the depletion region. Recall that
these processes are proportional to np − n2i (see Eqns. (333)-(335), page 99 or Eq. (340), page 100), so let’s
consider the product np:
(EF n −EF p )/(kB T )
np = n2i e = n2i eeVa /(kB T ) . (384)

From Eq. (340) at page 100 of the Lecture Notes, assuming σp = σn , and setting τ0 = σvth NT T , we
have:  
2 eV /(k T )
np − n2i ni e a B −1
Us =  E −E  ≈ , (385)
τ p + n + 2n cosh T i τ0 (p + n + 2n i )
0 i kB T
the last step depending on the assumption that only traps at energy ET = Ei contribute effectively to the
GR processes.
Now, from current continuity, at steady state:
dJn (x) dJp (x)
= − = e Us (x) . (386)
dx dx
We have calculated so far the current density Jn at x = −lp which we assumed was identical to Jp (x = ln )
thanks to the fact that there were no GR processes in the depletion region. We can now account for these
processes and evaluate the correction to the current they cause by integrating the equation for Jn from −lp
to ln :
 ln
Jn (ln ) = Jn (−lp ) + e dxUs(x) . (387)
−lp
Therefore for the total current we get:
 ln
eVa /(kB T )
J = Jp (ln )+Jn (ln ) = Jp (ln ) + Jn (−lp ) + e dxUs(x) = Js e − 1 + JGR .
−lp
(388)

ECE609 Spring 2010 114


The term JGR – the ‘generation/recombination current density – acts as a correction to the Shockley term.
In order to avaluate it in a simple approximation, let’s note that JGR will be dominated by processes occurring
at a position where p + n is a minimum, as seen from Eq. (385). But since np is constant,

d(n + p) = 0 → dn = − dp , d(np) = 0 → pdn + ndp = 0 ,

and, combining these two equations, these conditions imply n = p. By Eq. (384),
 
eVa
n = p = ni exp ,
2kB T
so that, from Eq. (385) we have:
 
n2i eVa /(kB T )
e ni eVa /(2k T )
−1
Us ≈ = e B −1 . (389)
2ni τ0 (1 + exp[eVa /(2kB T )] 2τ0

Assuming that Us takes this constant value – let’s call it Umax – throughout the depletion region (not strictly
correct, we are going to overestimate JGR ), we get for JGR :
ni eVa /(2k T )
JGR ≈ e Umax (ln + lp ) = e (ln + lp ) e B −1 . (390)
2τ0
From Eqns. (383) and (388) we have:
ni eVa /(2k T )
eVa /(kB T )
Jtot ≈ Js e − 1 + e (ln + lp ) e B −1 . (391)
2τ0
For a small Va the GR term dominates, while the Shockley term takes over at larger bias. It is customary to
lump these two types of asymptotic behavior by writing:

eVa /(nkB T )
Jtot ≈ Jsat e −1 , (392)

ECE609 Spring 2010 115


where the index n (called ‘non-ideality factor’) takes a value close to 2 al low Va and drops to unity as Va
grows larger.
• Short ‘base’ diodes.
The name ‘short base’ comes from bipolar transistors. These are two p-n junctions butted against each other
to form either p-n-p or n-p-n structures. The ‘central’ region is called ‘base’. Consider for simplicity just
the first p-n junction of a p − n − p structure. When the n-type base is very long, we can study that p-n
junction as we have done above. But if the base is ‘short’ (compared to the diffusion length Lp = (Dp τp )1/2
of minority holes in the n region), then the boundary condition pn (x → ∞) = pn0 used before Eq. (381)
must be replaced by pn (x = xR ) = pn0 , where xR is the edge of the ‘base’. In addition, the minority holes
may transit the base in a time so short that they will not be able to recombine. Thus, the current density of the
minority holes in the base will be constant. Since the DDEs imply

dp
Jp = eµp pF − eDp , (393)
dx
assuming as before that the field vanishes in the quasi-neutral region (F = 0), we have:

Jp x
p(x) = − + B, (394)
eDp

where B is an integration constant which we can find by imposing p(x = ln ) = pn0 exp[eVa /(kB T )]:
 
eVa Jp ln
B = pn0 exp + . (395)
kB T eDp

But the boundary condition at x = xR we’ve considered before implies pn (x = xR ) = pn0 , or:
   
Jp xR eVa Jp ln Jp (ln − xR ) eVa
p(x = xR ) = pn0 = − + pn0 exp + = + pn0 exp .
eDp kB T eDp eDp kB T
(396)

ECE609 Spring 2010 116


Solving for Jp we get:    
eDp pn0 eVa
Jp = exp − 1 . (397)
xR − ln kB T
Since the slope dp/dx is larger than in a long-base junction, the diffusion current will generally be larger.
• Junction capacitance.
We have so far limited our attention to the steady-state behavior of the device. We have seen that when we
move from the equilibrium situation (Va = 0) to forward or reverse bias (Va = 0) we must move charges out
(reverse bias) or in (forward bias) the depletion region. In practical applications it is important to know how
quickly the device can adjust to a new bias condition. One way to estimate this ‘time response’ of the device is
to consider the capacitances involved. Capacitance is a measure of the charge stored per unit change of voltage.
A larger capacitance means that more charge must be moved in or out, so that – for a fixed current – more
time is needed to complete the process. Thus, this translates into a longer delay in responding to the new bias
condition.
In a p-n junction two major capacitances are at play: 1. The capacitance associated with the charge which
must be moved in or out of the depletion region. This is called the ‘depletion capacitance’. 2. The capacitance
present under forward bias due to charges spilling over into the quasi-neutral regions. This is called ‘diffusion
capacitance’. Let’s now consider these two components separately.
– Depletion capacitance.
When the bias Va applied to the junction is varied, the width of the depletion region changes according to
Eqns. (373) and (374). The charge present in each depletion region due to the ionized dopants will be:

Qdepl = eAND ln = eANA lp , (398)

where A is the cross-sectional area of the junction. There will also be a component of charge due to the
motion of majority carriers. But this happens very quickly (in the time scale of picoseconds or less), so we can
ignore this delay. Thus, by definition of capacitance,
      1/2
 dQdepl   dln  e N N 1
Cdepl =   = eAND   = A s A D
   , (399)
dVa dVa 2 (NA + ND ) (Vbi − VA )1/2

ECE609 Spring 2010 117


or, noticing that from Eqns. (373) and (374)
 
2s (Vbi − Va )(NA + ND ) 1/2
lp + ln = , (400)
e NA ND

we can rewrite Eq. (399) as:


s A
Cdepl = , (401)
ln + lp
which is just the capacitance of a parallel-plate capacitor with a dielectric of permittivity s , with plates of
area A separated by a distance ln + lp .
– Diffusion capacitance.
The concentration of excess carriers diffusing in the quasi-neutral regions can be obtained from Eq. (381):

)

 δpn (x) = pn (x) − pn0 = pn0 e
 eV a /(k B T
− 1 e−(x−ln )/Lp (x > ln )
(402)


 δnp (x) = np (x) − np0 = np0 eeVa /(kB T ) − 1 e(x+lp )/Ln (x < −lp ) .

The charge per unit area will be (considering only holes, a similar expression will hold for electrons):
 ∞
eVa /(kB T )
Qdif f,p = e δpn (x) dx = eLp pn0 e −1 . (403)
ln

Under strong forward bias, eeVa /(kB T ) >> 1, so:


2
dQdif f,p e2 Lp pn0 eVa /(k T ) e Lp e
Cdif f,p ≈ = e B ≈ Jp (ln ) = τp Jp (ln ) , (404)
dVa kB T kB T Dp kB T

where we have used Eq. (383) in the last step. Accounting now for the charge of the minority electrons

ECE609 Spring 2010 118


diffusing into the p quasi-neutral region:

dQdif f,p dQdif f,n e


Cdif f ≈ + = [τp Jp (ln ) + τn Jn (−lp )] . (405)
dVa dVa kB T

Note: This is the equation found in most textbooks. However, in Karl Hess’ book one finds instead:
e
Cdif f ≈ [τp Jp (ln ) + τn Jn (−lp )] , (406)
2kB T

where the additional factor of 1/2 is explicitly commented and the claim is made that Eq. (405) is in error.
This factor can be justified in hand-waving fashion by noting that the charges Qdif f,p and Qdif f,n are like
the charges in the two opposite plates of a capacitor, so that the capacitance should be given by the change
w.r.t. the applied bias of the average of the electron and hole charges. A more sophisticated and rigorous
explanation is given by S. E. Laux and K. Hess, IEEE Trans. Electron. Device vol. 46, no. 2 (February 1999),
p. 396. Their argument is based on the observation that – rigorously speaking – the diffusion charge extends
also inside the depletion region, so that the integration in Eq. (403) should extend from 0 to ∞, not from ln
(and similarly for the expression for Qdif f,n ). Since as Va changes charges will leave the depletion region, we
will obtain a lower estimate for the charge, and so for the capacitance. In a way, this argument is equivalent to
our ‘hand-waving’ argument since both reduce to accounting for the charges throughout the entire junction,
not just in the quasi-neutral regions.
– Switching time.
Read the description of the ‘recovery time’ and ‘fall time’ in the text by Colinge and Colinge (pages 123-125).
• Hot electron effects and junction breakdown.
The entire analysis we have carried out so far is based on the DDEs which, as we saw, assume that the
electrons and holes remain at thermal equilibrium. Let’s now consider a few issues which clearly go beyond this
assumption.
– Electron heating.
The first issue is apparently a trivial one: At equilibrium, we may have a very large electric field in the

ECE609 Spring 2010 119


depletion region, especially if we consider heavy doping. We saw at page 86 of the Lecture Notes that the
electron temperature can grow as F 2 as the electric field increases. So, should we expect electron (and hole)
heating in the depletion region? A rigorous solution of the Boltzmann equation, most likely via a Monte Carlo
technique, can provide an answer. But a simple application of the hydrodynamic model (the third moment of
the BTE) is sufficient here: The relevant equation (Eq. (293) at the bottom of page page 85) was:

∂ E
1 E

≈ j·F − , (407)
∂t n τ̂w
or, setting (3/2)kB Tc = E
,

∂Tc 2 T − Tc
≈ j·F − . (408)
∂t 3kB n τ̂w

At equilibrium no current flows (j = 0), we are also at steady-state (∂Tc /∂t = 0), so we must have Tc = T ,
which expresses the fact that at equilibrium the electron temperature is equal to the lattice temperature. In
other words, the ‘average’ electron energy remains (3/2)kB T . Occasionally, an electron will ‘roll down’ the
field and gain kinetic energy above the equilibrium value. But there will also be electrons attempting to ‘climb
the potential hill’, so losing kinetic energy. On average, the two types of process will compensate each other.
– Impact ionization.
The current-voltage characteristics of a p-n junction show an exponentially increasing current as Va grows
positive, while the ‘leakage’ current density saturates to the value (from Eq. (383)):
 
eDp pn0 eDn np0
Js = + , (409)
Lp Ln

as the applied bias takes very large negative values (reverse bias). In practice, as Va grow sufficiently large,
the reverse current ‘snaps’ to very large values. This is called ‘junction breakdown’ and it is mainly due to
avalanche multiplication initiated and sustained by impact ionization, and by Zener tunneling (which we shall
discuss below).

ECE609 Spring 2010 120


Ionization rate. As shown by the top figures at page 94, the frequency at which a carrier – electron or hole
– of kinetic energy E above a ‘threshold’ Eth generates an electron-hole pair can be approximated by the
Keldysh formula:  
1 B E − Eth p
= , (410)
τii(E) τI (Eth ) Eth
for E > Eth , vanishing otherwise. The constant B is a dimensionless parameter and 1/τ (Eth ) is the
scattering rate at the threshold energy. The energy Eth is the minimum energy (dictated by energy and
momentum conservation) a carrier must have in order to be able to generate electron-hole pairs. Clearly,
Eth > EG by energy conservation. But momentum conservation may require that Eth be much higher,
since the recoil carriers and the generated carriers will have to share the available momentum and so cannot in
general have vanishing energy. The coefficient p at the exponent takes the value of 2 for semiconductors with
direct gap. In general, it takes values as high as 6 (as determined by complicated calculations). Lower values
for p imply a ‘soft threshold’ (carriers will ionize with slowly increasing probability as E grows beyond Eth ).
Large values of p imply a ‘hard threshold’: Ionization will occur almost immediately as the energy of the
carrier crosses the threshold energy. Phonon-assisted processes (in which one or more phonons are emitted or
absorbed during the ionization process) can soften the threshold, as well as the so-called U mklapp processes
(in which a carrier ends up in the next Brillouin Zone and G vectors of the reciprocal lattice is needed to ‘map
it back’ into the first BZ) can ‘soften’ the requirements of momentum conservation.
Ionization coefficient. The parameter which is of interest in the breakdown of p-n junctions is the ‘ionization
coefficient’ α (see Eq. (307), page 93). This is defined as the number of pairs generated per unit length, so
that it also represents the rate at which the current density increases (per unit length): Considering for now
only electrons:
dJn
= αn Jn . (411)
dx
Since each electron at energy E = E(k) generates electron-hole pairs at a rate 1/τii (E), if f (k) is the
electron distribution function, then:

1 2 f (k )
αn = d k , (412)
n vd
(2π)3 τii[E(k)]

ECE609 Spring 2010 121


 
where n is the electron density n = 2(2π)−3 dkf (k) and vd
= (2/n)(2π)−3 dkf (k)vF (k) is
the drift velocity (i.e., the group velocity along the direction of the field averaged over the electron distribution
function). It can be seen (perhaps not too trivially) that the functional dependence of τii as given by Eq. (410)
does not affect αn too much. On the contrary, the shape of the distribution function f (k) as a function of
the field dramatically affects αn . Therefore, the whole game consists in estimating how f changes with F .
Shockley had the basic idea that only those electrons which manage to be accelerated to Eth without losing
energy tophonons will contribute to the ionization process. The probability that an electron scatters per unit
time is k P (k , k). If Pnp (t) is the probability that the electrons has not scattered up to time t, the
probability that it will not have scattered up to t + dt will be:
 

Pnp (t + dt) = Pnp (t)  1 − dt P (k , k )  . (413)
k

Solving this equation for a time interval [t1 , t2 ], we get


 
  t  
2
Pnp = exp − P (k , k) dt . (414)
 t1 
k

(Note that k and k are functions of time due to their acceleration in the field, so the integration is not
trivial). Now, the probability that an electron will not scatter with phonons while being accelerated
 from

energy E = 0 to E = Eth can be obtained from Eq. (414) by noticing that 1/τph (k) = k P (k , k),
by using the chain-rule (let’s consider only the 1D case for simplicity):

 
dE dE dk dE eF h̄ dE −1
= = → dt = dE .
dt dk dt dk h̄ eF dk

ECE609 Spring 2010 122


Then,
  E   −1 
h̄ th dE dE
Pnp = exp − . (415)
eF 0 τph (E) dk

Clearly, one can argue about the assumption that after every collision the electrons must start from zero
energy. Nevertheless, this ‘lucky electron’ model captures the basic feature of the ionization coefficient at
small fields F , α ∝ exp(−constant/F ). However, despite its empirical success in this case, Shockley’s
lucky-electron model should not be taken too seriously (indeed it has caused more harm – in the form of
confusion and bad physics in the literature – than good). This may be seen by noticing that at larger electric
field experimentally one finds instead that an expression of the form α ∝ exp(−constant/F 2 ) matches the
data more accurately. This can be explained by assuming that at large enough fields the electron ionize as
soon as they reach the threshold energy Eth , that they lose all of their energy, so that the energy-loss term is
dkE/τiif (k). It may be shown that adding this term to the third-moment equations will yield the desired
behavior.
Avalanche. Let’s generalize Eq. (411) to the case of a p-n junction under strong reverse bias. Under
these bias conditions, the reverse ‘leakage current’ density Js is very small and it is mainly due to those few
electrons present in the p-type quasi-neutral region which flow into the n-type region, and holes undergoing a
similar (but opposite) process. If the electric field in the depletion region is sufficiently large, impact ionization
will occur. Considering one of those few electrons entering the depletion region from the left, it will have a
nonzero probability of creating an electron-hole pair. The generated hole will travel back towards the left and
it may impact-ionize, thus creating another electron-hole pair. The generated electron, in turn, will travel
back towards the right and the cycle will repeat itself. There will be a critical field at which the probability
that impact-ionization occurs will be unity. At this point, the process will run-away, more and more carriers
being generated and the current will grow without limits. This is the avalanche breakdow.
In order to quantify the onset of this condition, let’s note that the initial electron current injected from the
left will grow as:
dJn
= αp Jp + αn Jn , (416)
dx

ECE609 Spring 2010 123


or
dJn
− (αn − αp ) Jn = αp J , (417)
dx

where J = Jn + Jp is the total electron and hole current, constant at steady-state. To determine the
breakdown condition, let’s denote with Jn (−lp ) the electron current density incident from the left-edge of
the depletion region (at x = −lp ) and let’s assume that the electron current exiting the high-field depletion
region (at x = ln ) is a multiple Mn of Jn (−lp ). Since most of the current at x = ln will be carried
by electrons, we can assume Jn (ln ) = Mn Jn (−lp ) ≈ J . With this boundary condition the solution of
Eq. (417) can be written as:
  x   x    
1 x
Jn (x) = J + dx αp exp − dx (αn − αp ) exp dx (αn − αp ) .
Mn −lp −lp −lp
(418)

(Note: This comes from the fact that the general solution of the first-order linear differential equation

dy(x)
+ a(x)y(x) = b(x)
dx

is 
y(x) = C e−A(x) + e−A(x) dx b(x) eA(x) ,

where C is an integration constant determined by the boundary condition and A(x) = x dx a(x )).

Evaluating Jn (x) at x = −lp yields Jn (−lp ) = J/Mn , which is our boundary condition. We must
also require Jn (ln ) = J . With some algebra this implies:
 ln   x 
1
1 − = dx αn exp − dx (αn − αp ) . (419)
Mn −lp −lp

ECE609 Spring 2010 124


The avalanche breakdown voltage is defined as the voltage at which Mn → ∞. Thus, at breakdown the
ionization integral approaches unity:

 ln   x 

dx αn exp − dx (αn − αp ) = 1. (420)
−lp −lp

For a hole-initiated avalanche we get a symmetrical result:

 ln   ln 

dx αp exp − dx (αp − αn ) = 1. (421)
−lp x

Equations (420) and (421) are equivalent: The condition determining the onset of breakdown depends only
on what happens inside the depletion region, not on which type of carrier initiates the ionization process.
Note that for semiconductors for which αn ≈ αp (as in GaP, for example), the breakdown condition becomes
simply
 ln
dx α = 1 , (422)
−lp

with obvious meaning: If the probability of ionizing over the depletion region reaches unity, avalanche will
occur.
Breakdown voltage. The breakdown voltage can be calculated from known doping profile, ionization rates,
etc. from the equations above. In the case of an abrupt junction we have

Fmax (ln + lp )
VBD = , (423)
2

where the maximum field in the depletion region is given by Eq. (364). For ‘linearly graded’ junctions, the

ECE609 Spring 2010 125


formula above must be corrected by a factor 2/3.

ECE609 Spring 2010 126


– Zener breakdown.
Another cause of breakdown is related to the quantum mechanical process of tunneling. At a very large
electric field in a reverse-biased p-n junction, electrons in the valence band may tunnel across the band gap of
the semiconductor, thus creating an electron-hole pair. The process may lead to breakdown either directly (so
many pairs will be crated that the leakage current will grow) or indirectly: The electrons and holes generated
by the tunneling process will impact-ionize and an avalanche process will begin. This breakdown mechanism
is called ‘Zener breakdown’. It affects mostly heavily-doped junctions in which the built-in voltage and the
applied (reverse) bias fall over a narrow depletion region, thus giving rise to large electric fields.
A relatively simple estimate of the strength of the process may be obtained by assuming that the electrons
have to tunnel through a triangular barrier of height EG and length EG /(eF ) (see the figure in the previous
page). Using the WKB approximation (see Homework 1), the tunneling probability is proportional to:
 
∗ 1/2 3/2
4(2m ) EG
PZener (F ) ∝ exp  −  , (424)
3eh̄F

where m∗ is the electron effective mass in the gap (usually approximated by the smaller between the electron
and hole effective masses). More rigorous calculation show that the current is given by:

(2m∗)1/2 e3 F Va
JZener ≈ PZener (F ) . (425)
1/2
4π2 h̄2 EG

This expression is independent of temperature and it is valid only for semiconductors with a direct gap (such
as most of the III-V compound semiconductors). On the contrary, for semiconductors with indirect gap (such
as Si and Ge), the calculations must take into account the fact that crystal momentum must be supplied
(mainly by phonons) in order to allow a transition from the top of the valence band at the symmetry point Γ
to the bottom of the conduction band at other locations in the BZ (at the symmetry points L in Ge, near
the symmetry points X for Si). The role played by phonons in the process renders Zener breakdown quite
strongly temperature-dependent.

ECE609 Spring 2010 127


Other ‘diodes’: Heterojunctions,
Metal-Semiconductor junctions (Shottky contacts), MOS capacitors

p-n junctions are ubiquitous in semiconductor devices: We had to analyze their characteristics in some detail
because this understanding is required to analyze, in turn, the characteristics of many types of devices (bipolar
junction transistors – BJTs – in particular). However, other types of ‘junctions’ play a major role in several types of
transistors: Heterojunctions (that is, the junction between two different semiconductors) make up heterojunction
bipolar transistors (HBTs), high electron-mobility transistors (HEMTs, also known as ‘modulation-doped field-effect
transistors’, MODFETs), and injection lasers; Shottky contacts (that is, the ‘junction’ between a semiconductor
and a metal) enter heavily in the operation of metal-semiconductor FETs (MESFETs), among other devices;
MOS - metal-oxide-semiconductor- capacitors are at the heart of what is arguably the most important type of
transistor in VLSI technology, the ‘metal-oxide-semiconductor field-effect transistor’ (MOSFET).
We shall now discuss in turn each of these ’junctions’ (or ‘diodes’, as they broadly fall into the category of
two-terminal devices).

• Heterojunctions.
– Thermionic emission.
As shown in the left frame of the figure in the next page, let’s consider two different semiconductors brought
into contact. In the figure we use the example of GaAs and the alloy AlxGa1−x As (where x, known as the
‘mole fraction’, is the molar fraction of Al ions to Ga ions). GaAs has a smaller band gap (≈ 1.51 eV)
than Alx Ga1−x As (≈ 1.72 eV for x = 0.3, for example). Most of the discontinuity of the gap falls in the
conduction band. Therefore, as soon as the two materials are brought together, the conduction bands will
line-up as shown in the left frame of the figure.
Let’s now assume that the bands will remain fixed in this situation (we’ll worry later on about the possibility of
a new equilibrium position) and let’s attempt to compute the current caused by electrons flowing from GaAs
to AlxGa1−x As.

ECE609 Spring 2010 128


If we take the z axis as the axis perpendicular to the plane of the interface, we can compute the current by
multiplying the distribution function f (k) in GaAs (i.e., in the left semiconductor) by the z -component of the
velocity and by summing over all states with right-directed velocity, vz (k) > 0 and with sufficient energy and
‘correct’ angle of travel to overcome the barrier. This means, classically:

 
1 ∗ 2 1 h̄kz 2
m vz = m∗ ≥ ∆Ec . (426)
2 2 m∗

Thus the minimum value of the z -component of k an electron in GaAs must have to make it over the barrier
of height ∆Ec is kz0 = (2m∗∆Ec )1/2 /h̄. Thus, this ‘left-to-right’ current can be written as:
 
2e
jLR = dkx dky dkz vz (k) f (k) . (427)
(2π)3 kx ky kz >kz0

Taking for f the non-degenerate Boltzmann limit and switching integration variables from wavevector to

ECE609 Spring 2010 129


velocity we have:
 3  ∞  ∞  
m∗ e EF,L /(kB T ) m∗ 2 2 2
jLR = e dvx dvy dvz vz exp − (vx + vy + vz ) ,
h̄ 4π3 −∞ vz0 2kB T
(428)
where EF,L is the Fermi level in GaAs measured from the conduction-band minimum of GaAs itself.
2
We can use cylindrical coordinates, so that dvx dvy = vdvdφ (with v = (vx + vy2 )1/2 ) and, defining
E = m∗ v 2 /2, we have dE = m∗ vdv , so:
 ∞    2π  ∞
m∗ 2 2 ∗ 2
dvx dvy exp − (vx + vy ) = dφ dv v e−m v /(kB T ) =
−∞ 2kB T 0 0
 ∞
2π −E /(kB T ) 2π
= dE  e = = kB T , (429)
m∗ 0 m∗
so that:
 ∞
e m∗2 EF,L /(kB T ) ∗ 2 (E −|∆Ec |)/(kB T )
jLR = kB T e dvz vz e−m vz /(2kB T ) = A∗ T 2 e F,L .
2π2 h̄3 vz0
(430)
This is called ‘thermionic emission current’, and A = em∗ kB
∗ 2
/(2π2h̄3 ) is the so-called Richardson constant.
Following similar arguments, the current from Alx Ga1−x As to GaAs will be

/(k T )
jRL = A∗ T 2 e F,R B ,
E
(431)

where EF,RL is now the Fermi level in AlxGa1−x As measured from the conduction-band minimum of
Alx Ga1−x As.
Let’s now assume that the Alx Ga1−x As is doped with ND donors per cm3 and that the GaAs is undoped
or weakly n-type doped. Similarly to what we saw for p-n junction, at equilibrium the Fermi levels must

ECE609 Spring 2010 130


‘line up’. Physically, this is caused by the electrons in the AlxGa1−x As flowing into the GaAs-side of the
‘heterojunction’ until a built-in potential will block the flow. We can estimate the time-dependence of this
flow from the continuity equation:
∂n ∂j
e = , (432)
∂t ∂z
The current will be mainly caused by electrons flowing from right to left via thermionic emission. We can
assume that only electrons within a distance Ln from the interface will participate in the conduction, since
electrons in the Alx Ga1−x As farther away will have essentially zero vz (since thay are at equilibrium, suffering
too many collisions). Then, from Eqns. (431) and (432) we have:
∂n /(k T )
≈ −A∗ T 2 e F,R B /Ln .
E
e (433)
∂t
Assuming a time-dependent Fermi level EF,R (t) such that:
 ∞
(E (t)−E)/(kB T )
n(t) = dE ρR (E) e F,R , (434)
0
(where ρR (E) is the density of states at energy E in AlxGa1−x As), we can write:
  3/2
2
E (t)/(kB T ) πh̄
e F,R = n(t)4 = n(t)CR . (435)
2m∗ kB T

From Eqns. (433) and (435) we get:

∂n A∗ T 2
= − n(t)CR , (436)
∂t eLn
whose solution is:   
A∗ T 2 CR
n(t) = nc exp − t , (437)
eLn

ECE609 Spring 2010 131


where nc is the concentration of electrons in the AlxGa1−x As at t = 0. The time constant for the decay
of the electron concentration in the Alx Ga1−x As is thus τ ≈ eLn /(A∗ T 2 CR ) which is of the order of
picoseconds at room temperature. Thus, the Alx Ga1−x As will lose its electrons very quickly, leaving behind
positively ionized donors. The build-up of the positive charge in the Alx Ga1−x As and of the (electron)
negative charge in the GaAs will eventually build a barrier preventing further flow of electrons from the
Alx Ga1−x As to the GaAs. When, finally, equilibrium will be reached, the band diagram will look like what is
shown in the right frame of the figure at page 129.
– The depletion approximation (again).
Let’s quantify this band diagram by solving Poisson equation. Denoting by F the z -component of the electric
field, we have
∂F e + −
= (p − n + ND − NA ) , (438)
∂z s
where s is the dielectric constant of the semiconductor (a function which has a discontinuity at the
GaAs/AlxGa1−x As interface). Let’s take z = 0 at the interface. Let’s also embrace once again the
‘depletion approximation’, by assuming that in the AlxGa1−x As, at a distance W away from the interface the
semiconductor is charge-neutral, with zero electric field, since the free carriers (or the positive charges of the
dopants) will have screened the electric field at this distance. Therefore, this is a valid approximation as long
+
as W >> 2π/βs , where βs is the screening parameter (in our case βs2 = eND /(s kB T )). Then, the
‘built-in’ potential will be:

 W  W
W ∂F
Vbi = − dz F = − zF |0 + dz z . (439)
0 0 ∂z

From this equation, since we have assumed F (z) = 0 for z ≥ W , we have:

 W
∂F
Vbi ≈ dz z . (440)
0 ∂z

ECE609 Spring 2010 132


Neglecting the charges of the free carriers in the depletion layer, from Eq. (438):
 W + +
eND eW 2 ND
Vbi ≈ − dz z = . (441)
0 s 2s
This equation relates Vbi to W , but it gives us neither value. These values can be obtained by solving Poisson
equation on both sides of the interface and matching the solutions. Only when the GaAs-side of the junction is
very heavily doped, so that the bands distort significantly only in the Alx Ga1−x As-side of the heterojunction,
then we have approximately:

|eVbi | ≈ ∆Ec + Ec,L (−∞) − Ec,R (∞) , (442)

(where Ec,L is the energy of the bottom of the CB of GaAs at the far left, and Ec,R (∞) the energy of the
botom of the CB of AlxGa1−x As at the far right) from which W can be obtained via Eq. (441).
– Beyond the depletion approximaton.
A better solution of our problem can be obtained by going beyond the depletion approximation and by solving
Poisson equation on both sides of the interface. First, let’s recall that at the interface between two dielectrics
the normal component of the field D = s F and the ‘in-plane’ components of the field F must be continuous.
Since F = −∇φ, these conditions imply:
 
∂φ  ∂φ 
L = R (443)
∂z  z=0− ∂z  z=0+

and
φ(z = 0−) = φ(z = 0+ ) . (444)
Now let’ consider Eq. (438). As noted when discussing p-n junctions, this equation is nonlinear since the
carrier density n depends on the electrostatic potential itself:
 
eφ(z)
n(z) = nc (−∞) exp , (445)
kB T

ECE609 Spring 2010 133


where nc (−∞) is the equilibrium electron concentration in GaAs far away to the left. Let’s denote with ψ(z)
the quantity −eφ(z)/(kB T ). Then, on the left side of the interface, in the absence of holes and acceptors,
Poisson equation becomes:

∂ 2ψ e2 +
2
= − [ND − nc (−∞)eψ ] . (446)
∂z GaAs kB T
+
If the doping in GaAs is uniform, nc (−∞) = ND and

∂ 2ψ e2 nc (−∞) ψ
= (e − 1) . (447)
∂z 2 GaAs kB T
This equation cannot be solved in closed form. However, we can get some information about the electron
concentration on the left-side of the heterojunction as a function of the value of the potential, φi , and of the
electric field, Fi , at the interface. In order to do this, let’s multiply Eq. (447) by −∂ψ/∂z and integrate.
The left-hand-side becomes:
     
∂ 2 ψ ∂ψ 1 ∂ ∂ψ 2 1 ∂ψ 2
− dz = − dz = − , (448)
∂z 2 ∂z 2 ∂z ∂z 2 ∂z
while for the right-hand-side we get:
 
e2 nc (−∞) ψ ∂ψ e2 nc (−∞) ψ e2 nc (−∞) ψ
(e −1) dz = (e −1) dψ = (e −ψ+C) , (449)
GaAs kB T ∂z GaAs kB T GaAs kB T

where C is an integration constant. Therefore, from Eqns. (448) and (449), fixing C = −1 by requiring
∂ψ/∂z = 0 for z = −∞, we have the relation:

2nc (−∞)kB T
F2 = − (eψ − ψ − 1) . (450)
GaAs

ECE609 Spring 2010 134


Equation (447) can be integrated from z = −∞ to the interface z = 0 obtaining:

GaAs Fi = Qtot , (451)

where Qtot is the total excess charge (due to the electrons which have spilled over from Alx Ga1−x As-side of
the heterojunction) at the left of the interface. Note that electrons accumulate exponentially – the Boltzmann
law - with increasing interface potential, since from Eqns. (450) and (451) Qtot ∝ eψi /2 . Thus, one speaks
of an ‘accumulation layer’ at the GaAs side of the interface. Had we considered, instead, p-doped GaAs,
the derivation would have been very similar. However, in the presence of holes, for small values of ψi the
electrons would have been recombining with the holes – thus forming a depletion layer – before starting
to form an ‘inversion layer’ (so called because the p-type GaAs would become n-type, so ’inverted’) in
which the electron density would grow, as before, as ∝ eψi /2 . As the interface potential reaches the value
ψi = 2kB T ln(NA /ni ), then one talks of ‘strong inversion’, since the electron density at the interface
exceeds the hole density in the bulk of the p-type GaAs.
Equations (450) and (451), together with the continuity conditions Eqns. (443) and (444), can be used to
obtain a complete solution of Poisson equation.

• Shottky contacts.
– Metal-semiconductor junction.
When we bring a metal and a semiconductor together, as shown in the figure in the next page, we have a
situation not too dissimilar from what we have seen regarding heterojunctions: If the semicondutor (assumed
to be n-type) Fermi level is larger than Fermi level in the metal (as shown in the figure), electrons will flow
from the semiconductor untill a new equilibrium will be reached. This will result in the creation of a depletion
region in the semiconductor side of the ‘contact’. Similarly to what we have sen in the case of heterojunctions,
there will be an energy ‘barrier’ of height eφB between the bottom of the CB of the semiconductor and the
metal Fermi level. Denoting by eφM the metal workfunction (that is, the energy required to excite an electron
at the metal Fermi level to the level of the vacuum) and by eχ the electron affinity of the semiconductor, the

ECE609 Spring 2010 135


Shottky barrier height will be (looking at the figure):

eφBn = e (φM − χ) , (452)

where the subscript n reminds us that we have considered an n-type semiconductor. For a p-type
semiconductor, instead, we would have obtained

eφBp = EG − e (φM − χ) . (453)

ECE609 Spring 2010 136


– Non-ideal junctions.
This is what we expect in an ideal case. In practice, it often happens that the Fermi level of the semiconductor
is ‘pinned’ at some particular energy in the gap due to the presence of electron traps at the interface (interface
states). Years ago J. Bardeen had suggested that electronic states associated to unterminated bonds present
at the semiconductor interface may be an intrinsic cause of this ‘pinning’. Now it seems that things are more
complicated: For example, bare surfaces may ‘reconstruct’ and, when exposed to metals, may ‘deconstruct’ in
complicated ways, leading to amorphization, interdiffusion and clustering in the interfacial region. It is fair to
say that we do not know exactly what pins the Fermi level, so that in many cases the Shottky barrier height
should be regarded as an experimentally measured quantity.
– Image force effects.
Another cause of deviation from the ideal picture we used to reach Eq. (452) is caused by ‘image force effects’.
This is a basic idea of elementary electrostatics. Let’s consider an interface between two dielectrics (with
permittivities 2 – the metal, assumed to be in the semi-infinite space z < 0, and 1 – the semiconductor).
Let’s consider the electron in the semiconductor, which fills the half-space z ≥ 0, at a distance z = d from
the interface. By the method of images we can write the potential along the line (the z -axis) normal to the
interface and passing through the particle location as:
# $
e2 1 2 − 1 1
V (z) = − . (454)
4π1 |d − z| 2 + 1 |d + z|

The first term is the ‘usual’ potential due to the electron iteself. The second term is the image potential,
Vim (z), which we shall consider now. For the electron sitting at z = d the force due to its own image
(located at z = −d) will be:

dVim (z)  e2 2 − 1
Fim (d) = −  = − . (455)
dz z=d 16π1 d2 2 + 1

Now, the work performed against the image required to bring the electron from infinity to the position z > 0

ECE609 Spring 2010 137


will be:
 ∞
e2 2 − 1
∆φim (z) = ds Fim (s) = − . (456)
z 16π1 z 2 + 1
This can be viewed as a force which lowers (for 2 > 1 ) or raises (in the opposite case) the barrier height.
Coming to our explicit case, assuming for the metal 2 → ∞, and setting 1 = s , and adding the effect of
a uniform field Fz , we have a total potential energy

e2
Φim (z) = − − eFz z . (457)
16π1 z

This potential exhibits a maximum at some distance from the interface and it yields a barrier whose height is
reduced by the amount:
 
eFz 1/2
∆φB = . (458)
4πs
An additional barrier-reducing effect is purely quantum-mechanical and it’s caused by the penetration of
electronic wavefunctions of the metal into the semiconductor band gap. We shall not discuss it further.

ECE609 Spring 2010 138


– Thermionic current.
We can now compute the thermionic current. The current flowing from the left (the metal) to the right (the
semiconductor) can be computed following exactly the arguments used to derive Eq. (430). Thus:

∗ 2 −eφB /(kB T )
jLR = A T e . (459)

At equilibrium, the same current will flow from the semiconductor to the metal, so the net current will be
zero. If we now apply an external bias, we shall for now assume that the (quasi) Fermi levels remain constant
in the metal, EF,L and in the semiconductor, EF,R , and we assume that the difference between the Fermi
levels is just the applied bias, eVa = EF,R − EF,L, taken positive when we apply positive bias to the
semiconductor. (The assumption that the applied voltage drops in the depletion region of the semiconductor
while the quasi-Fermi levels remain constant will be discussed below.) For Va > 0 we can consider JLR
given, unchanged, by Eq. (459), since the barrier heights will remain unchanged by the application of the
external bias (approximately: the image force lowering will by slightly modified by the applied bias via the field
Fz in Eq. (458), but we shall ignore this small effect). On the contrary, the current flowing from the right will
be enhanced (for Va > 0) and will be:

∗ 2 −eφB /(kB T ) eVa /(kB T )


jRL = A T e e , (460)

ECE609 Spring 2010 139


so that the total current, j = jRL − jLR will be:

∗ 2
j = A T e−eφB /(kB T ) eVa /(kB T )
e −1 . (461)

Note the similarity with the Shockley equation, Eq. (383), for the current in p-n-junctions.
The thickness of the depletion region can be obtained as we did before for p-n-junctions and we obtain
 
2s (eVbi − Va ) 1/2
W = , (462)
eND

where the built-in potential Vbi is (see the figure) eφB − (Ec − EF ).
– Diffusion current.
The limitation of the derivation leading to Eq. (461) lies in the assumption that the quasi-Fermi level remains
constant in the semiconductor. Consider electrons flowing from the right, ‘jumping’ over the barrier towards
the metal. They will do so with a velocity vz roughly given by 2eφB /m∗ , of the order of at least 107
cm/s (and up to 108 cm/s) for most barrier-heights and semiconductors. This velocity is quite large because
only electrons streaming from right to left contribute to its average, since, having entered the metal, it’s very
unlikely that they will be reflected back. In other words, only the left-going half of the electron distribution
contributes to this average velocity. Therefore, the time constant characterizing the out-flow of these carriers
into the metal will be of the order of λ/vz , where λ is the electron mean-free-path. This flow of electrons
out of the depletion region towards the metal must be balanced by electrons coming into the depletion region
from the right at z = W . These electrons will not be ‘streaming’ in one direction (to the left), but be
diffused in all directions by scattering with impurities, phonons, etc. Thus, their characteristic velocity (now
averaged over left- and right-going carriers) will be of the order of the ‘drift’ velocity vd , typically of the order
of 105 cm/s, or even smaller, for the typical concentration gradients of interest. Therefore, the associated
time constant, λ/vd , will be at the very least two orders of magnitude smaller. This is true unless the width
W of the depletion region is smaller than λ. In this case also the supply of electrons from the right will occur
‘ballistically’ at the fast time scale λ/vz , since scattering will be negligible also for this ‘supply’ of electrons
from the right. This is the situation in which the supply of electrons from the right is able to keep up with

ECE609 Spring 2010 140


the escape of electrons to the left and our assumption of constant quasi-Fermi levels is valid. If, however,
the supply from the right cannot match the escape-rate to the left (which happens, as we have seen, when
W >> λ), then our picture is clearly inconsistent (since it leads to a situation away from steady-state, in
contrast to our assumption). In this case a diffusion-dominated model better describes the current flow in
Shottky diodes.
From current continuity
∂n
j = enµn Fz + eDn . (463)
∂z
Using Einstein’s relation to express µn in terms of Dn , we can re-write this equation as:
 
e ∂φ ∂n
j = eDn − n + . (464)
kB T ∂z ∂z
Multiplying by exp[−eφ(z)/(kB T )] this becomes:

−eφ(z)/(kB T ) ∂ne−eφ(z)/(kB T )
je = eDn . (465)
∂z
In steady-state the current does not depend on z , so
 W
−eφ(z)/(kB T ) −eφ(W )/(kB T ) −eφ(0)/(kB T )
j e dz = eDn n(W ) e − n(0) e . (466)
0
Using the solution φ(z) of Poisson equation in the depletion approximation we could in principle compute the
function  W
−eφ(0)/(kB T )
F (ND ) = e e−eφ(z)/(kB T ) dz , (467)
0
(which depends only on the doping density ND ) so that from Eq. (466), noticing that under an applied external
bias and assuming the the entire external potential drops in the depletion region, Va = φ(W ) − φ(0), we
have
−e(Vbi −Va )/(kB T )
jF (ND ) = eDn n(W ) e − n(0) . (468)

ECE609 Spring 2010 141


Now, since in the depletion approximation n(W ) is equal to its equilibrium value and, at equilibrium,
neq (W ) exp[−eVbi /(kB T )] = neq (0), we have:

eVa /(kB T )
jF (ND ) = eDn neq (0) e − n(0) . (469)

As last step, we must find the non-equilibrium concentration n(0). To do that, let’s observe that the current
flowing to the left at the junction (z = 0) is the thermionic emission current with concentration n(0) and
zero barrier height (since we are considering the current at the top of the barrier). Therefore,
 
0 ∗ 2 n(0) neq (0)
jth = A T − , (470)
Nc Nc

having used Eqns. (430), (431), and the known expression neq ≈ Nc exp[EF − Ec )/(kB T )]. The
second term in the equation above represents the current coming back, which is essentially an equilibrium
current, since no voltage drops in the left side of the junction. The term A∗ T 2 /(eNc ) is called ‘interface
0
recombination velocity’ and it is denoted by vR . Thus, using now the fact that jth is equal to j (since this is
the current at the interface and j does not depend on z ), combining Eq. (469) and (470) we finally get:

eDn neq (0)[eeVa /(kB T ) − 1]


j = . (471)
F (ND ) + D/vR

Note, once again, that is of the form j ∝ exp[eVa /(kB T )] − 1, the typical form of the current in many
junctions.

ECE609 Spring 2010 142


– Tunneling current.

The last component of the current we must consider is of quantum-mechanical origin. It is the tunneling
current illustrated in the figure above.
In the example illustrated in the figure, electrons in the metal can tunnel across the Shottky barrier and enter
the semiconductor. Similarly, electrons can tunnel from the semiconductor into the metal.
Such a current can be calculated using the WKB approximation we have seen in Homework 1. Let the z -axis
be normal to the plane of the interface, let z = 0 be the location of the interface and let vz (k) be the
z -component of the velocity of the electrons hitting the interface and attempting the tunneling process. The
probability Pt that an electron will tunnel across the barrier can be estimated using the WKB approximation:
#  z $
t
Pt ≈ exp −2 dz κz (z) , (472)
0

ECE609 Spring 2010 143


where zt is the tunneling distance (that is, the position at wich the electrons will emerge from the barrier
in the semiconductor, κ(z) = {2m ∗ ∗[eφ(z) − Ez ]}1/2 /h̄ is the imaginary component of the electron
wavevector in the barrier, φ(z) the electrostatic potential in the depletion region, and Ez = h̄2 kz2 /(2m∗M )
the component of the electron kinetic energy in the metal due to the z -component of the electrons wavector
in the metal with effective mass m∗M . (Note that this separation of kinetic energy is only valid for parabolic
bands and for directions along the principal axes of ellipsoidal band-structures.)
Tunneling processes, in principle, conserve the component of the momentum parallel to the interface. Indeed,
the translation symmetry on the plane of the interface ensures the conservation of this quantity. In practice, at
sufficiently high temperatures, phonons can provide the ‘parallel momentum’ required to relax this conservation
law. Also, interface roughness and deviations fron ideality of the interface, often have the same effect. Since
it is a realistic effect, and since it simplifies considerably the calculations, we shall assume that parallel
momentum is not conserved and we can replace Ez with the total electron kinetic energy E in the expression
for κ(z) above.
The tunneling current from the metal to the semiconductor will then be:

 #  z $
dk t
jM S = 2e v (k) fM (k) [1 − fS (E k)] exp
3 z
−2 dz κz (z) , (473)
kz >0 (2π) 0

where fM (k) is the Fermi function in the metal and the factor [1 − fS (E k)] ensures that we tunnel only
into unoccupied states in the semiconductor . Performing the integral in polar coordinates, recalling that
vz (k) = h̄kz /m∗M = h̄k cos θ/m∗M , we have:

 ∞ #  z $  π/2
eh̄ 3 t
jM S = dk k fM (k) [1−fS (k)] exp −2 dz κz (z) dθ sin θ cos θ ,
2π2 m∗M 0 0 0
(474)
the upper integration limit for θ being π/2 since we restrict the intergration only over right-directed electrons

ECE609 Spring 2010 144


(kz > 0 in Eq. (473)). Thus, changing integration variable k to E, as usual:
 ∞ #  z $
em∗M t
jM S = dE E fM (E) [1 − fS (E)] exp −2 dz κ(z) . (475)
2π2h̄3 0 0

Similarly, the current due to electrons tunneling from the semiconductor to the metal will be:
 ∞ #  z $
em∗S t
jSM = dE E fS (E) [1 − fM (E)] exp −2 dz κ(z) . (476)
2π2h̄3 0 0

The total current will be, obviously, jT = jSM − jM S


 .zt
Let’s consider two cases in which the WKB integral, 0 dzκ(z) can be handled analytically. In the first
case, quite a realistic one, we approximate the potential φ(z) with its expression valid in the depletion
approximation,
 2
z
φ(z) = φ B 1 − , (477)
W
where φ B , as illustrated in the figure, is the total voltage drop in the semiconductor, φ B = Vbi − Va =
φB − Va + EF,R − Ec and W = [2s (Vbi − Va )/(eND )]1/2 is the width of the depletion region. In
this case, with some algebra, we find:

 z  z
t (2m∗S )1/2 t
dz κ(z) = dz [eφ(z) − E]1/2 =
0 h̄ 0

 W [1−(E/eφ )1/2 ]   1/2


(2m∗S )1/2 B eφ B 2
= dz 2
(W − z) −E , (478)
h̄ 0 W
having used the fact that zt = W [1 − (E/eφ B )1/2 ].

ECE609 Spring 2010 145


Changing integration variable y = (eφ B /W 2 )(W − z)2 − E , we get:

 z   1/2  eφ −E   1/2

t mS W B y
dz κ(z) = dy =
0 2eφ B
h̄ 0 y + E
  1/2   1/2 
m∗S W E 1
= 2eφB  1− −  . (479)
2eφ B h̄ eφ B 2
Thus, from Eqns. (475) and (479), we get:
 ∞   
em∗M 2(2m∗S eφ B )1/2  1/2

jM S = dE E fM (E) [1−fS (E)] exp − W 1 − E/(eφB ) − 1/2
2π2h̄3 0 h̄
(480)
At this point the integration must be performed numerically.
Another example consists in approximating the Shottky barrier as a triangular barrier

φ(z) = φ B − Fz z , (481)

where Fz may be chosen as the interfacial field, 2φ B /W , or some suitable average of the field over the
depletion region. In this case we get:
 ∞  
em∗M 4(2m∗S )1/2
jM S = 3
dE E fM (E) [1 − fS (E)] exp − (eφ B − E)3/2 . (482)
2π2 h̄ 0 3eh̄Fz

Given the complexity of these calculation, one may wonder how ohmic contacts could ever be realized. One
possible explanation consists in assuming that height of the Shottky barrier, eφB , may be zero. This may
happen in a few cases. But more likely is the scenario in which heavy amorphization of the interface may
result in a very heavily doped – almost metallic – thin layer of semiconductor. The very high doping will result

ECE609 Spring 2010 146


in such a thin depletion layer, that tunneling across this barrier may be a dominant process, almost killing the
resistance of the contact. This can be seen easily from Eq. (480): For W << h̄/(2m∗S eφ B )1/2 , the WKB
exponential approaches unity, thus resulting in a very large current. Similarly from Eq. (482), noticing that
the field Fz is inversely proportional to W .

• MOS capacitors.
The Metal-Oxide-Semiconductor (MOS) – or Metal-Insulator-Semiconductor (MIS) – diode is arguably the
most used and useful device in VLSI technlogy. Its ideal structure is shown in the figures below.

METAL INSULATOR SEMICONDUCTOR METAL INSULATOR SEMICONDUCTOR

eφM eχ
eφB

EC
EC Ei
EF,M Ei EF,M EF,S
eVFB < 0 eψB EF,S EV
EV

FLAT BANDS ZERO BIAS

ECE609 Spring 2010 147


METAL INSULATOR SEMICONDUCTOR
METAL INSULATOR SEMICONDUCTOR

EC
Ei
EF,S
EF,M EV
eVG > 0
eVG < 0 EC
EF,M
Ei
EF,S
EV

ACCUMULATION INVERSION
Starting from a semiconducting substrate (assumed to be p-type in the figures), an insulator is grown or
deposited on the substrate. Typically, the natural oxide of Si, SiO2 , is thermally grown by heating the Si wafer
to temperatures in the range 850-1000oC in oxygen-rich ambient. The relative simplicity of this process (which,
however, must be extremely clean) and the unsurpassed electronic properties of SiO2 are probably the reasons
why Si has been the dominant material in microelectronics. After the growth or deposition of the insulator a
metal (or highly-doped polycrystalline Si) is deposited over it. In the figures a metal is considered.
At ‘flat band’ the alignement of the bands is illustrated in the first figure. In the ideal case, charge would flow
across the insulator, so that the Fermi levels in the metal, EF,M , and in the semiconductor, EF,S , would
line-up and the difference, φM S , between the metal and the semiconductor work-functions would vanish:
 
EG
φM S = φM − χ+ + ψB = 0 for p − type semiconductor , (483)
2e

ECE609 Spring 2010 148


and  
EG
φM S = φM − χ+ − ψB = 0 for n − type semiconductor . (484)
2e
In practice, the time required for the Fermi level to line-up is extremely long and this ideal situation is never
achieved. The figure shows that the application of a small bias, the flat-band voltage VF B , is required to
line-up the Fermi levels. Its value will be given by the nonvanishing φM S of Eq. (483) or (484) above.
The application of a negative bias, VG , to the metal (usually called the ‘gate’) drives the MOS diode into
‘accumulation’: As seen in the figure, the bias causes an accumulation of holes at the Si-SiO2 interface. The
application of a positive gate bias, instead, results in the ‘inversion’ of the semiconductor surface, electrons now
piling up at the interface. Let’s now consider these processes in some detail.

– Interface space-charge region.

SiO2 Semiconductor
EC

Ei
eψ eφB
EF
eψs

EV

The figure above shows in more detail the band-diagram near the semiconductor-SiO2 interface. Let z be the

ECE609 Spring 2010 149


coordinate along the normal to the interace, z = 0 be the location of the interftace, and let’s define by ψ(z)
the potential, taken as zero in the bulk and measured from the intrinsic Fermi level Ei . Then, assuming the
non-degenerate limit, the electron and hole concentrations will be:
 
eψ(z)
np (z) = np0 exp = np0 exp(βψ) , (485)
kB T

pp (z) = pp0 exp(−βψ) , (486)

where ψ is positive downward (as in the figure), β = e/(kB T ), and np0 and pp0 are the equilibrium
electron and hole concentrations in the bulk.
Let ψs be the surface potential so that

ns = np0 exp(βψs ) , ps = pp0 exp(−βψs ) (487)

are the surface concentrations of the carriers.


We can deal with the ‘exact’ Poisson equation as we have done above (see pages 134 and 135), extracting
some information before embracing the ‘usual’ depletion approximation. We have:

∂ 2ψ e + −
2
= − [ND + p(z) − NA − n(z)] , (488)
∂z s
where s is the (static) dielectric constant of the semiconductor. From Eqns. (485) and (486) we have:

pp − np = pp0 e−βψ − np0 eβψ , (489)


+ −
so that, since by charge neutrality in the bulk ND − NA = np0 − pp0 , we have

∂ 2ψ e −βψ βψ
= − [p p0 (e − 1) − n p0 (e − 1)] . (490)
∂z 2 s

ECE609 Spring 2010 150


Let’s now integrate Eq. (490) from an arbitray location z towards the bulk (z → ∞, as we have done on
page 134-135, Eqns (448)-(451): Let’s multiply Eq. (490) by −∂ψ/∂z and integrate. The left-hand-side
becomes:  ∞ 2  ∞    
∂ ψ ∂ψ 1 ∂ ∂ψ 2 1 ∂ψ(z) 2
− 2 ∂z
dz = − dz = , (491)
z ∂z 2 z ∂z ∂z 2 ∂z
having used the fact that the field vanishes as z → ∞. For the right-hand-side we get:
 ∞  0
e −βψ ∂ψ e
[pp0 (e −1) − np0 (eβψ
−1)] dz = [pp0 (e−βψ −1) − np0 (eβψ −1)] dψ =
s z ∂z s ψ

e −βψ βψ
= [pp0 (e + βψ − 1) − np0 (e − βψ − 1)]. (492)
s β
Therefore we have the following relationship between field and potential at any location z :
 2    
2k T ep β n
F2 = (e−βψ + βψ − 1) +
B p0 p0
(eβψ − βψ − 1) . (493)
e 2s pp0

Introducing the Debye length, LD = [kB T /(e2 pp0 )]1/2 (the dielectric screening length in the p-type
non-degenerate bulk Si), and denoting as G(βψ, np0 /pp0 )2 the term in square brackets in Eq. (493), we
have:  
1/2
∂ψ 2 kB T n p0
F = = ± G βψ, , (494)
∂z eLD pp0
the plus (minus) sign valid for ψ > 0 (ψ < 0).
The charge at the interface can now be expressed using Gauss law and the value of the field at the interface
(obtained by setting ψ = ψs in Eq. (494)):
 
21/2s kB T np0
Qs = −s Fs = ∓ G βψs , . (495)
eLD pp0

ECE609 Spring 2010 151


This represents the total charge per unit area, shown in the figure below. For negative ψs the charge is
positive (holes), corresponding to accumulation. At flat band the total charge is obviously zero. In depletion
1/2
and weak inversion the term βψ in the function G dominates, so that he charge grows as ψs . Finally,
in strong inversion the charge is negative, the term (np0 /pp0 )eβψs being the dominant one. By definition,
strong inversion begins at ψs = 2ψB , the value of the surface potential at which the electron concentration
at the interface equals the hole concentration in the bulk.

1015
p–type Si 300K
NA = 4x1015 cm–3

1014 ~ exp(βψs/2)
strong inversion

1013 ~ exp(–βψs/2)
Qs (cm–2)

accumulation

flat
band
1012
weak
depletion inversion

1011 ~ ψs1/2

ψB EC
10
10
–0.4 0.0 0.4 0.8
ψs (Volt)

ECE609 Spring 2010 152


In order to obtain separately the electron and hole charges, ∆n and ∆p, we must integrate p(z) and n(z)
from the surface to the bulk:
 ∞  0
−βψ epp0 LD e−βψ − 1
∆p = pp0 (e − 1) dz = dψ , (496)
0 21/2 kB T ψs G(βψ, np0 /pp0 )

and  ∞  0
βψ epp0 LD eβψ − 1
∆n = np0 (e − 1) dz = dψ . (497)
0 21/2 kB T ψs G(βψ, n /p
p0 p0 )
The differential capacitance of the semiconductor depletion layer is given by:

∂Qs s 1 − e−βψs + (np0 /pp0 )(eβψs − 1)


CD = = . (498)
∂ψs 21/2 LD G(βψs , np0 /pp0 )

At flat-band condition, ψs = 0, so:


s
CD,F B = . (499)
LD

– Ideal capacitance-voltage characteristics.


It is very important to understand the capacitance-voltage characteristics of an MOS diode (or capacitor),
in view of their relevance to the operation of an MOS field-effect transistor. Let’s recall that we are now
interested in the ‘differential capacitance’. That is, we apply a dc (or slowly-varying) gate bias to the diode,
but, in addition, we apply a small (of the order of kB T /e or less) ac bias at a given frequency.
If we apply a gate bias VG to the gate (while keeping the semiconductor substrate grounded), part of the
voltage, ψs , drops in the semiconductor and part, Vox , in the insulator. The latter will be given obviously by:

Qs tox
Vox = Fox tox = , (500)
ox

ECE609 Spring 2010 153


having used the fact that s Fs = ox Fox , and having denoted with tox the thickness of the insulator. Thus,
the ’oxide capacitance’ will be
dQs ox
Cox = = . (501)
dVox tox
But the total capacitance will also depend on the charge induced by the voltage drop in the semiconductor,
Eq. (498). Therefore the total capacitance will be the series-capacitance of the insulator, Cox and of the
depletion region, CD :
CoxCD
Ctot = . (502)
Cox + CD
For a given insulator thickness, the oxide capacitance is thus the maximum capacitance.

Cox Cox
GATE CAPACITANCE

low
frequency
CFB

high–frequency

GATE VOLTAGE
Looking at the figure above, in accumulation (VG < 0) holes pile-up very close to the semiconductor-

ECE609 Spring 2010 154


insulator interface. As the gate bias is reduced, the depletion capacitance begins to matter, depressing the
total capacitance. To estimate its value, in the depletion approximation we can write the potential in the
semiconductor as:  
z 2
ψ(z) ≈ ψs 1 − , (503)
W
where the depletion width W can be obtained in the usual way (see Eq. (441) above):
 
2s ψs 1/2
W ≈ . (504)
eNA

Thus, the depletion capacitance will be the result of charges responding to the ac-bias at the edge of the
depletion region, so that
s
CD = , (505)
W
and
ox
Ctot = . (506)
tox + (ox /s )W
At flat band we should replace W with LD , while the onset of strong inversion (also called the ‘turn-on
voltage’ or, more commonly, the ‘threshold voltage’, as this marks the onset of strong conduction in MOS
field-effect transistors), we have
   
2s 2ψB 1/2 4s kB T ln(NA /ni ) 1/2
W = Wmax ≈ = . (507)
eNA eNA

When the surface potential reaches the strong-inversion value, ψs = 2ψB , the electron concentration at the
insulator-semiconductor interface is so large (provided enough time is given to the minority carriers so that
an inversion layer is indeed formed, as we shall see below) that it will screen the field and it will prevent any
further widening of the depletion region. Indeed, as we can see in the plot Qs vs. ψs at page 152, a small
increment of surface potential will result in a huge increase of the electron charge. Thus, one can assume

ECE609 Spring 2010 155


that in strong inversion most of the additional VG will drop in the insulator and no significant increas of
ψs will occur. Therefore, Eq. (507) gives the maximum width of the depletion region. The only exception
to this is given by the application of a very quickly-varying dc bias: If the gate bias VG is increased very
quickly to positive values, there will be little or no time for generation/recombination processes in the bulk
to provide/absorb enough minority carriers (electrons) to feed the inversion layer. In this case ψ will exceed
2ψB , W will grow beyond Wmax and the capacitance will drop below its minimum value Cmin given below
by E. (508).
As the gate bias becomes even more positive, we must distinguish two different situations. If the applied bias
is varied slowly and the applied ac bias is of low frequency, generation and recombination processes in the
bulk may be able to follow the ac signal. Thus, electrons in the inversion layer will respond, the response of
the depletion layer will be screened by the ac-varying inversion charge and the differential capacitance will rise
back to the value given by Cox. If, on the contrary, the frequency of the ac signal is large, than the inversion
charge will not be able to follow the signal. The depletion capacitance will be ‘clamped’ at the minimum value
s /Wmax , and so the total capacitance will remain at the minimum value
ox
Cmin = , (508)
tox + (ox /s )Wmax

independent of gate bias.


Experimentally the transition between the low-frequency and the high-frequency behavior happens around
1-50 Hz. However, MOS diodes built on substrates of excellent quality (small density of SRH centers) may
exhibit the low-frequency behavior only when there is no applied ac signal and the dc gate bias is varied
sufficiently slowly (even a fast varying dc VG – say, a C-V sweep in a few seconds, may push the diode into
its high-frequency response).
– Deviations from ideality: Oxide charges and interface traps.
So far we have considered an ideal MOS structure in which both the insulator and the interfaces are free from
defects and impurities. This was not the case early on when MOS devices were first fabricated. Even today,
after decades of research and development has enabled the realization of almost-ideal structures, during the
operation of a device defects may be created by energetic carriers ‘hitting’ the interface or being injected into

ECE609 Spring 2010 156


the insulator. These defects may be broadly classified into two categories: Fixed charges in the oxide and
electrically-active interface states. (The word ‘fixed’ labelling the charges in the oxide refers to the fact that
their charge-state – or electronic population – does not depend on the applied bias, while their location – as
in the case of mobile Na or K ions – may depend on the bias and thermal history of the device.)
Insulator charges are typicaly either impurities (historically mobile Na and K ions were the subject of many
efforts), or defects, such as dangling bonds, local stress in the SiO2 ionic network, oxygen vacancies induced
by growth, processing – such as irradiation. Such charges have the main effect of shifting the threshold voltage

Qs (2ψB )
VT = Vox,si + 2ψB = + 2ψB , (509)
Cox

or the flat-band voltage. In order to derive a general expression for the flat-band voltage shift due to
an arbitrary distribution of charges ρ(z) inside the insulator (where z now denotes the distance from the
gate-insulator interface) let’s consider first an ideal (free of charge) MOS capacitor originally at flat-band
condition and now add a sheet of charge λ (charge per unit area) at z . This charge will induce polarization
(image) charges both in the metal and in the semiconductor. The latter charge will bend the bands and so
modify the gate bias we must apply to recover flat-band. Let’s now apply an additional bias to the metal, so
to bring the semiconductor at flat band. In this situation there will be no field in the metal (by definition),
no field in the semiconductor (as we are at flat band). There will only be a constant electric field λ/ox
for −tox < z < 0. Thus, the metal potential will have moved by an amount zλ/ox with respect to the
initial situation. This will be the shift of the flat-band voltage caused by the sheet of charge. Therefore, for a
distribution ρ(z) of charges inside the insulator, the flat-band shift will be given by:
 tox  tox
ρ(z) 1 z
∆VF B = z dz = ρ(z) dz . (510)
0 ox Cox 0 tox

Note that charges close to the gate-insulator interface have no effect on VF B (the polarization charges at the
metal surface screen completely the oxide charges), while charges near the semiconductor-insulator interface
have maximum effect.

ECE609 Spring 2010 157


Interface traps (or states) are defects – typically Si dangling bonds – whose occupation depends on the
position of the Fermi level at the insulator-semiconductor interface. They have a twofold effect: Depending
on their occupation, they shift the flat-band voltage, as in Eq. (510). Since for these states z = tox , their
electrostatic effect is strong. More importantly, becaue of the intrinsic delay in responding to an ac signal (the
time constants for emission and capture are those of SRH centers), they store charge, thus contributing to the
total capacitance of the MOS diode. Their capacitance, Cit , is in series with Cox and in parallel to CD , so
that   −1
1 1 Cox(CD + Cit )
Ctot = + = . (511)
Cox CD + Cit Cox + CD + Cit
Several methods have been devised to measure the density, Dit , of the interface traps. They all rely on a
measurement of their capacitance. Ideally, if one knew very accurately the theoretical C-V characteristics of
the diode, a comparison of the theoretically computed and experimentally measured characteristics will yield
the desired capacitance, Cit, and so the density Dit , since the density of the interface-trap charge at energy
E in the gap (the position of the Fermi level in the gap at a surface potential ψs ) will be ∝ Cit (ψs ): Having
obtained Cit from the total capacitance via Eq. (511), at a given gate bias we have:
dQit Nit
Cit = = e = e Dit , (512)
dψs dE
where Nit is the total number of traps up to energy E in the gap and Dit is the trap density per unit
area and energy in the gap. Recalling that VG = Vox + ψs , we have dVG = dψs + dVox . Since
dQ = Cox dVox = Ctot dVG , we have dVG = dψs + (Ctot/Cox) dVG , and so we can obtain
dψs /dVG from the relation
dψs Ctot
= 1 − . (513)
dVG Cox
Thus, from Eqns. (506), (512), and (513) we can extract dQit /dψs :
 
dQit Ctot dψs −1 C
Dit (E) = = − d . (514)
dψs e dVG e

ECE609 Spring 2010 158


This is the density of interface traps per unit energy in the gap at the energy E which, as we said above,
indicates the position of the Fermi level inside the semiconductor gap at the interface.
In practice, detailed theoretical curves C − V are hard to compute. Therefore, typically one replaces the
theoretical curve with a C-V curve obtained under conditions such that the interface traps do not respond.
Since the characteristic time of the response of the trap is of the order of
 
1 e(ψB − ψs )
τ = exp − , (515)
vth σni kB T

(for a p-type substrate), either a low-temperature measurement (so that at a given ac frequency ω the response
time τ becomes so long that the trap occupation does not vary) or a high-frequency measurement will provide
almost ideal C-V characteristics. A comparison between high-frequency and low-frequency measurements (or
high-T and low-T measurements) will provide Cit .

There are two major corrections we should make to the analysis followed so far: We have used the non-degenerate
limit (using Maxwell-Boltzmann instead of Fermi-Dirac statistics in relating carrier density to (quasi)Fermi levels)
and we have ignored completely all quantum mechanical properties of the electrons.
The first correction – important at large densities such as strong inversion and accumulation – complicates the
mathematical analysis so that it becomes impossible to derive analytically even the relations field/potential or
surface-field/total-charge given by Eqns. (494) and (495). Only numerical work can give us reliable answers.
Yet, the analysis followed so far is qualitatively correct and gives a quantitatively correct picture in the important
region covering weak-accumulation to weak inversion.
Quantum mechanical properties are usualy important in transport. So far, we have limited our attention to
electrostatics. Yet, even the electrostatic poperties can be affected by quantum mechanics when the charge
carriers are confined within regions of size comparable to (or, a fortiori, smaller than) the thermal wavelength
of the electrons. Accumulation and inversion layers are such reasons and they do present this problem. Let’s
consider this situation in more detail, limiting our attention to inversion layers.

ECE609 Spring 2010 159


– Quantization effects in inversion layers.
The thermal wavelength, λth , of an electron is given by the deBroglie relation λth = h/pth = 2π/kth ,
where pth = h̄kth is the average momentum of electrons at thermal equilibrium. For the average carrier
energy at thermal equilibrium we have Eth = (3/2)kB T = h̄2 kth 2
/(2m∗ ), so that, at 300K,

 
2πh̄ m0 1/2
λth = ≈ 6.2 nm (516)
(3m∗kB T )1/2 m∗

Consider the Si-SiO2 interface in inversion/strong-inversion. The electron density ns (the ‘sheet density’ per
unit area, see the figure at page 152) is of the order of 1011 -to-1013 cm−2 , corresponding to an interfacial
field Fs = ens /s of the order of 104-to-106 V/cm. Thus, an electron of thermal energy ((3/2)kB T ≈ 40
meV) will be ‘squeezed’ by the field against the interface over a confining distance ∆z ≈ 3kB T /(2Fs ),
very much like a particle in a box considered before. This distance is of the order of 40-to-0.4 nm, comparable
or even smaller than the electron thermal wavelength. We are not allowed to ignor the wave-like nature
of electrons when we confine them so tightly: We expect that discrete energy levels will emerge from
the confinement. If we confine a particle in a region of width ∆z , by Heisenberg’s principle the particle
momentum will suffer an uncertainty ∆k ∼ 1/∆z , so that the confined particle will have a minimum energy
E0 ∼ h̄2 ∆k2 /(2m∗) ∼ h̄2 /(2∆z 2 m∗ ), called the ‘zero-point energy‘. In strong inversion this energy
may be comparable to (or even larger than) the thermal energy, and quantum effects due to the confinement
should not be ignored.
A simplified case exemplifies the situation. Consider a ‘triangular well’: Electrons are confined at the left
(z = 0) by an infinitely high potential wall (mimicking the Si-SiO2 barrier) and at the right by a potential
eφ(z) = Fs z , where Fs is constant field given for example, by Gauss law, Fs = ens /s , as we just saw.
Assuming either spherical bands or ellipsoidal valleys with principal axes aligned with the plane of the interface
and its normal, since the potential is a function of the coordinate z alone, the ‘effective mass’ (or ‘envelope’)
wave equation (see Eq. (104), page 32 of the Lecture Notes) can be solved by separating variables (see
Eq. (113), page 35 of the Lecture Notes): Writing ψ(x, y, z) = ψx (x)ψy (y)ψz (z), the factors ψx (x)
and ψy (y) will be just right-travelling or left-travelling plane waves of the form ∼ exp(±ikx x) and

ECE609 Spring 2010 160


∼ exp(±ikx x), respectively. The factor ψz (z), instead, obeys the Schrödinger-like effective mass equation:

h̄2 d2
− ψz (z) + eFz zψz (z) = E ψz (z) , (517)
2mz dz 2

where mz is the effective mass along the direction normal to the interface. The boundary conditions on ψz
are, obviously, ψz (0) = 0 and ψz (z → ∞) = 0.
Equation (517) can be written as

 
d2 2mz eFs E
ψz (z) − z − ψz (z) = 0 . (518)
dz 2 h̄2 eFs

Let’s now change independent the variable to t = (2mz eFs /h̄2 )1/3 [z − E/(eFs )] and set u(t) = ψz (z),
so that Eq. (518) becomes:

u − tu = 0, (519)
where the prime denotes derivative with respect to t and the boundary conditions are u(t0 ) = 0 and
u(t → ∞) = 0, having indicated with t0 the value of t corresponding to z = 0:

 
E 2mz eFs 1/3
t0 = − (520)
eFs h̄2

As trivial as this equation may appear, it is actually one of the possible forms in which the Bessel equation
may be cast. Its solutions – shown in the figure together with the similar function Bi(t) we are not interested
in now – are functions called ‘Airy functions’:

ECE609 Spring 2010 161


        
1 1/2 2 2/3 2 2/3 t 1/2 2 2/3
u(t) = Ai(t) = t I−1/3 t + I1/3 t = K1/3 t ,
3 3 3 3π 3
(521)
where I±1/3 (x) are ‘modified Bessel functions of the first kind’ and K1/3 (x) is the ‘modified Bessel function
of the second kind’, all of order 1/3. An integral representation of the Airy function (useful for numerical
evaluation) is:
 ∞
1 i(tz+z 3 /3)
Ai(t) = e dz . (522)
2π −∞
This shows that our solution u(t) = Ai(t) vanishes as t → ∞, so the boundary condition at the right is
always satisfied. Not so for the other boundary conditions: Only particular values of E such that Ai(t0 ) = 0
(where t0 is given by Eq. (520) above) are acceptable. The first few roots of Ai(t) are at t=-2.33811, -
4.08795, -5.52056, -6.7867144, ..., so that only discrete values of E (let’s call them En ) are acceptable. An

ECE609 Spring 2010 162


approximated formula for the roots of the Airy functions yields:

  1/3   1/3     2/3


2 2
h̄ 2/3 h̄ 3πeFs 3
En = (eFs ) |tn | ≈ n+ , (523)
2mz 2mz 2 4

for n = 0, 1, 2, ..., having approximated the roots of the Airy function by tn ≈ −[(3π/2)(n + 0.75)]2/3.
Thus, the solutions of Eq. (517) are associated to a discrete spectrum of eigenvalues En and can be written
as:
   
2mz eFs 1/3 En
ψz,n (z) = Ai z − . (524)
h̄2 eFs

These functions are shown in the figure in the next page, together with the confining potential eFs z . Note
that there are two families (called ‘ladders’) of solutions: Indeed we are illustrating the case of Si. We
have two minima of the conduction band sitting at the center of ellipsoidal equi-energy surfaces with the
longitudinal axis along z . For this two-fold degenerate family, mz = mL ≈ 0.9m0 . Their energy levels and
wavefunctions are shown as solid lines. In addition, there are 4 minima such that mz = mT ≈ 0.19m0. By
Eq. (523), for a given n their energy is higher, thanks to the smaller mass. This second four-fold degenerate
ladder of eigenlevels is called the ‘primed’ ladder (with eigenvalues denonetd by En ) and the corresponding
eigenvalues and (squared) wavefunctions are shown as dashed lines in the figure.

ECE609 Spring 2010 163


0.15

SQUARED WAVEFUNCTIONS (arb. units) and POTENTIAL ( V )


0.10

0.05

0.00
0 5 10 15 20 25
DISTANCE FROM INTERFACE ( nm )

ECE609 Spring 2010 164


1.1 0.60 1.5
1.0

electron density (1020 cm–3)


0.50
0.9 1.2
0.8 0.40
0.7 quantum
potential (V)

potential (V)
0.30 0.9
0.6 classical
0.5 0.20
0.4 0.6
0.10
0.3
0.2 –0.00
0.1 0.3
–0.10
0.0
–0.1 –0.20 0.0
0 20 40 60 80 0 5 10 15 0 1 2 3 4 5 6
z (nm) z (nm) z (nm)
The ‘energy levels’ are actually ‘bands’ (properly called ‘subbands’) since only the motion along the z axis is
quantized. The motion of the electrons on the plane of the interface is still ‘free’, so each index n labels a
band of energy En (kx , ky ) = (h̄2 /2)(kx 2
/mx + ky2 /my ) + En .
The ‘triangular well approximation’ just considered is ‘instructive’ but not generally accurate. To properly
account for confinement effects we must solve the envelope Schrödinger-like equation

h̄2 d2
− ψz (z) + eV (z)ψz (z) = E ψz (z) , (525)
2mz dz 2

where the potential eV (z) is the solution of the Poisson equation

d2 V (z) e +
2
= [NA − n(z)] , (526)
dz s

+
where NA is the density of ionized acceptors in the p-type substrate and the electron density n(z) is given

ECE609 Spring 2010 165


by:


n(z) = ni(EF ) |ψz,i (z)|2 , (527)
i=0
where   
md kB T FF − Ei
ni (EF ) = ln 1 + exp , (528)
πh̄2 kB T

is the occupation of the subband i (see homework assignement 5). Thus, we must solve simultaneously (‘self-
consistently’) Eqns. (525) and (526) (note that the electron charge density depends on the wavefunctions

which, in turn, depend on the potential), while computing the Fermi level so that i ni (EF ) yields the
desired total electron sheet density. The problem is usually solved iteratively, by starting with an appropriate
‘guess’ for the potential (the semiclassical solution or even the depletion approximation), by solving the
Schrödinger equation, Eq. (525), computing the Fermi level EF at the desired density for the subband
energies Ei obtained from the solution of the Schrödinger equation, computing the charge density en(z)
and solving the Poisson equation Eq. (526). The cycle is then repeated, by solving Eq. (525) with the new
potential, etc. When some convergence criterion is met (the potential or the charge density change from one
iteration to the next by no more than some preset amount), the procedure is halted. The figure above shows
a typical solution (subbands, wavefunctions, potential), as well as a comparison between the classical and
quantum-mechanical charge distribution.
There are three major effects due to carrier confinement in inversion layers:
1. Threshold voltage shift. The ‘zero-point energy’ E0 (the energy of the ground-state subband) can be
noticeably above the bottom of the conduction band at the interface. Moreover, the density of states in the
CB will be reduced with respect to the bulk, 3D value. Therefore, a negative gate bias of a larger magnitude
will be required to induce a given charge density. This results in a shift of the threshold voltage, VT , which
is approximately given by E0 /e.
2. Reduction of the gate capacitance. As shown in the right-most frame of the previous figure, the electron
charge density in the inversion layer peaks at the interface in the classical model, but reaches a maximum
at z = z
, a few tenths of nm away in the quantum-mechanical case (up to a full nm for ns ∼ 1013

ECE609 Spring 2010 166


cm−3 ). For an insulator thickness of the order of 1 or 2 nm, this can be a significant effect. Indeed the
oxide capacitance Cox = ox /tox in inversion will be become

ox ox
Cox → < . (529)
tox + (ox /s ) z
tox

This reduction of gate capacitance is indeed a major concern in VLSI technology.


3. Modification of the transport properties. From the discussion above, it should be clear that electrons in
inversion layers are confined in two dimensions, since they are not free to move along the direction normal
to the interface. They constitute what’s known as a ‘two dimensional electron gas’ (2DEG). Only electrons
occupying subbands at very high energy can be described once more as bulk, 3D particles, since the energetic
separation of the subbands is so small that a continuum representation is suitable. Under very strong
confinement, only a few subbands will be populated. Furthermore, the energetic separation of the subbands
grows with confining field (see Eq. (523), so the density of states (DOS) is very different from the bulk, 3D
DOS. Finally, the wavefunctions associated to the electronic states are not plane-waves any longer, so that
the matrix elements of the various collision processes will be significantly different from their 3D counterparts.
The net effect is that the transport properties (e.g., mobility and saturated velocity) are significantly affected
by the quantum confinement. Note that we can still employ the semiclassical Boltzmann Transport Equation
to describe transport (under the same approximations), but now the BTE describes a 2D system with only
two degrees of freedom, those on the plane of the interface. We shall see in detail some of these effects
when dealing with MOS Field-Effect Transistors.

– Tunneling effects.
The final effect we must consider is electron tunneling across the insulator. This is becoming an increasingly
important process. Let’s see why. As illustrated in the figure, electrons can tunnel either via ‘direct tunneling’
(at left, for thin insulators and/or small insulator fields) or via ‘Fowler-Nordheim tunneling’ (at right, for thick
insulators and/or strong fields). Using the usual WKB approximation, the tunneling probabilities in the two

ECE609 Spring 2010 167


cases are: For direct tunneling:

   
4(2emox) 1/2
3/2 2(2moxeφ B )1/2
Pd ≈ exp − φB [1 − (1 − tox Fox /φ B )]3/2 ∼ exp − tox
3h̄Fox h̄

= e−2κ tox , (530)


where Fox is the field in the insulator, mox the effective mass in the gap of the insulator, φ B = φB − E0
is the effective barrier height, reduced by the energy E0 of the bottom subband in the inversion layer, tox is
the insulator thickness, the last step has been made assuming a thin insulator (tox << φ B /Fox ), and we
have defined an ‘average’ decay constant κ = (2moxeφ B )1/2 /h̄. For FN-tunneling, instead:

 
4(2emox)1/2 3/2
PF N ≈ exp − φB = = e−(4/3)κ zt , (531)
3h̄Fox

where zt = φ B /Fox is the tunneling distance across the triangular barrier. As device scaling progresses,
thinner and thinner insulators are used, together with a reduced applied bias. When insulators were 10 nm
thick and the applied bias was of the order of 5 V, FN tunneling was the only concern. This possibly caused
only minor ‘leakage’ when the devices were turned-on very strongly, so only during the ‘on’ state. Today,
instead, insulators are as thin as 2 nm (or even less). Even if the applied bias has been reduced to less than
1 V, from Eq. (530) it is clear that direct tunneling is becoming increasingly important. The major cause
of concern is its independence of bias: Electrons can tunnel across the ‘trapezoidal’ (or almost ‘rectangular’
barrier) under any bias condition. This leakage in the ‘off-state’ causes unwanted power dissipation and it
constitutes one of the problems (if not ‘the’ problem) we must face attempting to scale devices to even smaller
dimensions.

ECE609 Spring 2010 168


3.5 3.5

3.0 3.0

2.5 2.5

2.0 2.0
potential (V)

potential (V)
1.5 eφB – E0 1.5 eφB – E0

1.0 1.0

0.5 0.5

0.0 0.0

–0.5 direct tunneling –0.5 Fowler–Nordheim tunneling

–1.0 –1.0
–20 0 20 40 60 80 –80 –60 –40 –20 0 20 40 60 80
z (nm) z (nm)

ECE609 Spring 2010 169

You might also like