[go: up one dir, main page]

0% found this document useful (0 votes)
40 views33 pages

E02 MS Junctions

Uploaded by

Manu Manohar
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
40 views33 pages

E02 MS Junctions

Uploaded by

Manu Manohar
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 33

Junctions 251

complicate calculations of junction properties, and a computer must be used


in solving the problem accurately.
Most of the conclusions we have made regarding carrier injection,
recombination and generation currents, and other properties are qualitatively
applicable to graded junctions, with some alterations in the functional form
of the resulting equations. Therefore, we can apply most of our basic concepts
of junction theory to reasonably graded junctions as long as we remember
that certain modifications should be made in accurate computations.

Many of the useful properties of a p-n junction can be achieved by simply 5.7
forming an appropriate metal–semiconductor contact. This approach is obvi- Metal–
ously attractive because of its simplicity of fabrication; also, as we shall see seMicOnDuctOr
in this section, metal–semiconductor junctions are particularly useful when JunctiOns
high-speed rectification is required. On the other hand, we also must be able
to form nonrectifying (ohmic) contacts to semiconductors. Therefore, this
section deals with both rectifying and ohmic contacts.

5.7.1 schottky barriers


In Section 2.2.1 we discussed the work function q≥m of a metal in a vacuum.
An energy of q≥m is required to remove an electron at the Fermi level to
the vacuum outside the metal. Typical values of ≥m for very clean surfaces
are 4.3 V for Al and 4.8 V for Au. When negative charges are brought near
the metal surface, positive (image) charges are induced in the metal. When
this image force is combined with an applied electric field, the effective work
function is somewhat reduced. Such barrier lowering is called the Schottky
effect, and this terminology is carried over to the discussion of potential bar-
riers arising in metal–semiconductor contacts. Although the Schottky effect
is only a part of the explanation of metal–semiconductor contacts, rectifying
contacts are generally referred to as Schottky barrier diodes. In this section
we shall see how such barriers arise in metal–semiconductor contacts. First
we consider barriers in ideal metal–semiconductor junctions, and then in
Section 5.7.4 we will include effects which alter the barrier height.
When a metal with work function q≥m is brought in contact with a
semiconductor having a work function q≥s , charge transfer occurs until the
Fermi levels align at equilibrium (Fig. 5–40). For example, when Φm 7 Φs,
the semiconductor Fermi level is initially higher than that of the metal before
contact is made. To align the two Fermi levels, the electrostatic potential of
the semiconductor must be raised (i.e., the electron energies must be low-
ered) relative to that of the metal. In the n-type semiconductor of Fig. 5–40
a depletion region W is formed near the junction. The positive charge due
to uncompensated donor ions within W matches the negative charge on the
metal. The electric field and the bending of the bands within W are similar
to effects already discussed for p-n junctions. For example, the depletion
width W in the semiconductor can be calculated from Eq. (5–21) by using
252 Chapter 5

Metal Semiconductor

 
m >  s n-type   n
 

q s q

qm Ec
EF s q(m   s) = qV0
qB = q( m  )
Ec
EF m EF m EF s

Ev

Metal Semiconductor
Ev

W
(a) (b)

Figure 5–40
A Schottky barrier formed by contacting an n-type semiconductor with a metal having a larger work
function: (a) band diagrams for the metal and the semiconductor before joining; (b) equilibrium band
diagram for the junction.

the p+ -n approximation (i.e., by assuming the negative charge in the dipole


as a thin sheet of charge to the left of the junction). Similarly, the junction
capacitance is APs/W, as in the p+ -n junction.17
The equilibrium contact potential V0, which prevents further net elec-
tron diffusion from the semiconductor conduction band into the metal, is the
difference in work function potentials Φm - Φs. The potential barrier height
ΦB for electron injection from the metal into the semiconductor conduction
band is Φm -x, where qx (called the electron affinity) is measured from the
vacuum level to the semiconductor conduction band edge. The equilibrium
potential difference V0 can be decreased or increased by the application of
either forward- or reverse-bias voltage, as in the p-n junction.
Figure 5–41 illustrates a Schottky barrier on a p-type semiconductor,
with Φm 6 Φs. In this case aligning the Fermi levels at equilibrium requires
a positive charge on the metal side and a negative charge on the semiconduc-
tor side of the junction. The negative charge is accommodated by a depletion
region W in which ionized acceptors (Na- ) are left uncompensated by holes.
The potential barrier V0 retarding hole diffusion from the semiconductor to
the metal is Φs - Φm, and as before this barrier can be raised or lowered

17
While the properties of the Schottky barrier depletion region are similar to those of the p+ -n region, it is
clear that the analogy does not include forward-bias hole injection, which is dominant for the p+ -n region,
but not for the contact of Fig. 5–40.
Junctions 253

Metal Semiconductor Figure 5–41


A Schottky barrier
 
m < s p-type   between a p-type
p
  semiconductor
and a metal
qs q having a smaller
qm work function:
Ec Ec
(a) band
diagrams before
EF m joining; (b) band
diagram for
the junction at
EFs EF equilibrium.
Ev Ev

Metal Semiconductor q(s  m) = qV0

(a) (b)

by the application of voltage across the junction. In visualizing the barrier


for holes, we recall from Fig. 5–11 that the electrostatic potential barrier for
positive charge is opposite to the barrier on the electron energy diagram.
The two other cases of ideal metal–semiconductor contacts (Φm 6 Φs
for n-type semiconductors, and Φm 7 Φs for p-type) result in nonrectifying
contacts. We will save treatment of these cases for Section 5.7.3, where ohmic
contacts are discussed.

5.7.2 rectifying contacts

When a forward-bias voltage V is applied to the Schottky barrier of Fig. 5–40b,


the contact potential is reduced from V0 to V0 - V (Fig. 5–42a). As a result,
electrons in the semiconductor conduction band can diffuse across the deple-
tion region to the metal. This gives rise to a forward current (metal to semi-
conductor) through the junction. Conversely, a reverse bias increases the
barrier to V0 + Vr, and electron flow from semiconductor to metal becomes
negligible. In either case flow of electrons from the metal to the semiconduc-
tor is retarded by the barrier Φm - x. The resulting diode equation is similar
in form to that of the p-n junction

I = I0(eqV>kT - 1) (5–79)

as Fig. 5–42c suggests. In this case the reverse saturation current I0 is not
simply derived as it was for the p-n junction. One important feature we can
254 Chapter 5

 V   Vr 

q( m  )
q( V0  V )
Ec
q( m  ) EF m
EFs
qV
EFm q(V0  Vr)

Ev

Ec
Forward bias EF s
(a)

I
Ev

Reverse bias
(b)
V

(c)

Figure 5–42
Effects of forward and reverse bias on the junction of Fig. 5–40: (a) forward bias; (b) reverse bias;
(c) typical current–voltage characteristic.

predict intuitively, however, is that the saturation current should depend


upon the size of the barrier ≥B for electron injection from the metal into the
semiconductor. This barrier (which is Φm - x for the ideal case shown in
Fig. 5–42) is unaffected by the bias voltage. We expect the probability of an
electron in the metal surmounting this barrier to be given by a Boltzmann
factor. Thus

I0 ∝ e-qΦB > kT (5–80)

The diode equation (5–79) applies also to the metal–p-type semicon-


ductor junction of Fig. 5–41. In this case forward voltage is defined with the
semiconductor biased positively with respect to the metal. Forward current
increases as this voltage lowers the potential barrier to V0 - V and holes flow
from the semiconductor to the metal. Of course, a reverse voltage increases
the barrier for hole flow and the current becomes negligible.
In both of these cases the Schottky barrier diode is rectifying, with
easy current flow in the forward direction and little current in the reverse
direction. We also note that the forward current in each case is due to the
Junctions 255

injection of majority carriers from the semiconductor into the metal. The
absence of minority carrier injection and the associated storage delay time
is an important feature of Schottky barrier diodes. Although some minority
carrier injection occurs at high current levels, these are essentially major-
ity carrier devices. Their high-frequency properties and switching speed are
therefore generally better than typical p-n junctions.
In the early days of semiconductor technology, rectifying contacts were
made simply by pressing a wire against the surface of the semiconductor.
In modern devices, however, the metal–semiconductor contact is made by
depositing an appropriate metal film on a clean semiconductor surface and
defining the contact pattern photolithographically. Schottky barrier devices
are particularly well suited for use in densely packed integrated circuits,
because fewer photolithographic masking steps are required compared to
p-n junction devices.

5.7.3 Ohmic contacts


In many cases we wish to have an ohmic metal–semiconductor contact, hav-
ing a linear I–V characteristic in both biasing directions. For example, the
surface of a typical integrated circuit is a maze of p and n regions, which must
be contacted and interconnected. It is important that such contacts be ohmic,
with minimal resistance and no tendency to rectify signals.
Ideal metal–semiconductor contacts are ohmic when the charge
induced in the semiconductor in aligning the Fermi levels is provided by
majority carriers (Fig. 5–43). For example, in the Φm 6 Φs (n-type) case
of Fig. 5–43a, the Fermi levels are aligned at equilibrium by transferring
electrons from the metal to the semiconductor. This raises the semicon-
ductor electron energies (lowers the electrostatic potential) relative to the
metal at equilibrium (Fig. 5–43b). In this case the barrier to electron flow
between the metal and the semiconductor is small and easily overcome
by a small voltage. Similarly, the case Φm 7 Φs (p-type) results in easy
hole flow across the junction (Fig. 5–43d). Unlike the rectifying contacts
discussed previously, no depletion region occurs in the semiconductor in
these cases since the electrostatic potential difference required to align the
Fermi levels at equilibrium calls for accumulation of majority carriers in
the semiconductor.
A practical method for forming ohmic contacts is by doping the
semiconductor heavily in the contact region. Thus if a barrier exists at the
interface, the depletion width is small enough to allow carriers to tunnel
through the barrier. For example, Au containing a small percentage of Sb
can be alloyed to n-type Si, forming an n+ layer at the semiconductor sur-
face and an excellent ohmic contact. Similarly, p-type material requires a
p+ surface layer in contact with the metal. In the case of Al on p-type Si,
the metal contact also provides the acceptor dopant. Thus the required p+
surface layer is formed during a brief heat treatment of the contact after
the Al is deposited.
256 Chapter 5

Metal Semiconductor

 
m < s n-type   n
 
~
~ ~
~
q m
q s q Ec
EFm EFm EFs
q ( s   m )
Ec
q (   m )
EFs
Ev

Ev

(a) (b)

Metal Semiconductor
m > s p-type
 
  p
~
~ ~
~ q  
q s

Ec
q m
Ec
EFs
Ev
q ( m   s )
EFm EFm EFs
Ev

(c) (d)

Figure 5–43
Ohmic metal–semiconductor contacts: (a) Φm 6 Φs for an n-type semiconductor, and (b) the equilibrium
band diagram for the junction; (c) Φm 7 Φs for a p-type semiconductor, and (d) the junction at equilibrium.
Junctions 257

5.7.4 typical schottky barriers

The discussion of ideal metal–semiconductor contacts does not include cer-


tain effects of the junction between the two dissimilar materials. Unlike a
p-n junction, which occurs within a single crystal, a Schottky barrier junc-
tion includes a termination of the semiconductor crystal. The semiconduc-
tor surface contains surface states due to incomplete covalent bonds and
other effects, which can lead to charges at the metal–semiconductor inter-
face. Furthermore, the contact is seldom an atomically sharp discontinuity
between the semiconductor crystal and the metal. There is typically a thin
interfacial layer, which is neither semiconductor nor metal. For example,
silicon crystals are covered by a thin (10-20 Å) oxide layer even after etch-
ing or cleaving in atmospheric conditions. Therefore, deposition of a metal
on such a Si surface leaves a glassy interfacial layer at the junction. Although
electrons can tunnel through this thin layer, it does affect the barrier to cur-
rent transport through the junction.
Because of surface states, the interfacial layer, microscopic clusters
of metal–semiconductor phases, and other effects, it is difficult to fabricate
junctions with barriers near the ideal values predicted from the work func-
tions of the two isolated materials. Therefore, measured barrier heights are
used in device design. In compound semiconductors the interfacial layer
introduces states in the semiconductor band gap that pin the Fermi level
at a fixed position, regardless of the metal used (Fig. 5–44). For example,
a collection of interface states located 0.7∼0.9 eV below the conduction
band pins EF at the surface of n-type GaAs, and the Schottky barrier height
is determined from this pinning effect rather than by the work function of
the metal. An interesting case is n-type InAs (Fig. 5–44b), in which EF at
the interface is pinned above the conduction band edge. As a result, ohmic

~ 0.8 eV
Ec Ec
EF EF
Interface Ev
states

Ev

Metal nGaAs Metal nInAs


(a) (b)

Figure 5–44
Fermi level pinning by interface states in compound semiconductors: (a) EF is pinned near Ec - 0.8 eV
in n-type GaAs, regardless of the choice of metal; (b) EF is pinned above Ec in n-type InAs, providing
an excellent ohmic contact.
258 Chapter 5

contact to n-type InAs can be made by depositing virtually any metal on


the surface. For Si, good Schottky barriers are formed by various metals,
such as Au or Pt. In the case of Pt, heat treatment results in a platinum
silicide layer, which provides a reliable Schottky barrier with ΦB ≃ 0.85 V
on n-type Si.
A full treatment of Schottky barrier diodes results in a forward current
equation of the form
I = ABT 2e-qΦB > kTeqV > nkT (5–81)

where B is a constant containing parameters of the junction properties and n


is a number between 1 and 2, similar to the ideality factor in Eq. (5–74) but
arising from different reasons. The mathematics of this derivation is similar
to that of thermionic emission, and the factor B corresponds to an effective
Richardson constant in the thermionic problem.

5.8 Thus far we have discussed p-n junctions formed within a single semicon-
heterO- ductor (homojunctions) and junctions between a metal and a semicon-
JunctiOns ductor. The third important class of junctions consist of those between
two lattice-matched semiconductors with different band gaps (hetero-
junctions). We discussed lattice-matching in Section 1.4.1. The interface
between two such semiconductors may be virtually free of defects, and
continuous crystals containing single or multiple heterojunctions can be
formed. The availability of heterojunctions and multilayer structures in
compound semiconductors opens a broad range of possibilities for device
development. We will discuss many of these applications in later chapters,
including heterojunction bipolar transistors, field-effect transistors, and
semiconductor lasers.
When semiconductors of different band gaps, work functions, and
electron affinities are brought together to form a junction, we expect dis-
continuities in the energy bands as the Fermi levels line up at equilib-
rium (Fig. 5–45). The discontinuities in the conduction band ∆Ec and the
valence band ∆Ev accommodate the difference in band gap between the
two semiconductors ∆Eg. In an ideal case, ∆Ec would be the difference in
electron affinities q(x2 - x1), and ∆Ev would be found from ∆Eg - ∆Ec .
This is known as the Anderson affinity rule. In practice, the band discon-
tinuities are found experimentally for particular semiconductor pairs.
For example, in the commonly used system GaAs–AlGaAs (see Figs. 3–6
and 3–13), the direct band gap difference ∆ EΓg between the wider band
gap AlGaAs and the narrower band gap GaAs is apportioned approxi-
mately 23 in the conduction band and 13 in the valence band for the het-
erojunction. The built-in contact potential is divided between the two
Chapter 7, MESFET and Related Devices 4 225

The MODFET and its two-dimensional electron gas.


A comparison of three field-effect transistors-MOSFET, MESFET, and
t: MODFET.
J'

I'
7.1 METAL-SEMICONDUCTOR CONTACTS
h
The first practical semiconductor device was the metal-semiconductor contact in the form
of a point contact rectifier, that is, a metallic whisker pressed against a semiconductor.
The device found many applications beginning in 1904. In 1938, Schottky, suggested that
the rectifying behavior could arise from a potential barrier as a result of the stable space
charges in the semiconductor. The model arising from this concept is known as the Schottky
barrier. Metal-semiconductor contacts can also be nonrectifying; that is, the contact has
a negligible resistance regardless of the polarity of the applied voltage. This type of con-
tact is called an ohmic contact. All semiconductor devices as well as integrated circuits
need ohmic contact to make connections to other devices in an electronic system. We
consider the energy band diagram and the current-voltage characteristics of both the rec-
tifying and ohmic metal-semiconductor contacts.

7.1.1 Basic Characteristics


The characteristics of point contact rectifiers were not reproducible from one device to
another. They have been largely replaced by metal-semiconductor contacts fabricated by
planar processes (see Chapters 10-14). A schematic diagram of such a device is shown
in Fig, la. To fabricate the device, a window is opened in an oxide layer, and metal layer
is deposited in a vacuum system. The metal layer covering the window is subsequently
defined by a lithographic step. We consider a one-dimensional structure of the metal-
semiconductor contact shown in Fig. lb, which corresponds to the central section in Fig.
la, between the dashed lines.

! Fig. 1 (a) Perspective view of a metal-semiconductor contact fabricated by the planar process.
(b) One-dimensional structure of a metal-semiconductor contact.

I
226 b Chapter 7. MESFET and Related Devices

Figure 2a shows the energy band diagram of an isolated metal adjacent to an iso-
lated n-type semiconductor. Note that the metal work function qqm is generally differ-
ent from the semiconductor work function qq8.The work function is defined as the energy
difference between the Fermi level and the vacuum level. Also shown is the electron affin-
ity g, which is the energy difference between the conduction band edge and the vac-
uum level in the semiconductor. When the metal makes intimate contact with the
semiconductor, the Fermi levels in the two materials must be equal at thermal equilib-
rium. In addition, the vacuum level must be continuous. These two requirements deter-
mine a unique energy band diagram for the ideal metal-semiconductor contact, as shown
in Fig. 2b.
For this ideal case, the barrier height qqB, is simply the difference between the metal
work function and the semiconductor electron affinity§:
q q ~ n= q @ m - q X . (1)
Similarly, for the case of an ideal contact between a metal and a p-type semiconductor,
the barrier height q q B p is given by

Vacuum level
----------
-----t---- 7-----7---

Metal Semiconductor

Fig. 2 (a) Energy band diagram of an isolated metal adjacent to an isolated n-type
semiconductor under thermal nonequilibrium condition. (b) Energy band diagram
of a metal-semiconductor contact in thermal equilibrium.

Both q$Bn( in electron'volts) and eB,(in volts) are referred to as the barrier height.
Chapter 7. MESFET and Related Devices 4 227

where Egis the bandgap of the semiconductor. Therefore, for a given semiconductor and
for any metal, the sum of the barrier heights on n-type and p-type substrates is expected
to be equal to the bandgap:

On the semiconductor side in Fig. 2b, Vbiis the built-in potential that is seen by elec-
trons in the conduction band trying to move into the metal.
Vbi = q ~ -Vn
n ' (4)
The qVn is the distance between the bottom of the conduction band and the Fermi level.
Similar results can be given for the p-type semiconductor.
Figure 3 shows the measured barrier heights for n-type silicon2and n-type gallium
arsenidea3Note that qqBnincreases with increasing q&. However, the dependence is not
as strong as predicted by Eq. 1.This is because in practical Schottky diodes, the disrup-
tion of the crystal lattice at the semiconductor surface produces a large number of sur-
face energy states located in the forbidden bandgap. These surface states can act as donors
or acceptors that influence the final determination of the barrier height. For silicon and
gallium arsenide, Eq. 1generally underestimated the n-type barrier height and Eq. 2 over-
estimates the p-type barrier height. The sum of qqBnand q@Bp,however, is in agreement
with Eq. 3.
Figure 4 shows the energy band diagrams for metals on both n-type and p-type semi-
conductors under different biasing conditions. Consider the n-type semiconductor first.
When the bias voltage is zero, as shown in the left side of Fig. 4a, the band diagram is
under a thermal equilibrium condition. The Fermi levels for both materials are equal.
If we apply a positive voltage to the metal with respect to the n-type semiconductor, the
semiconductor-to-metal barrier height decreases as shown on the left side of Fig. 4b. This
is a forward bias. When a forward bias is applied, electrons can move easily from the semi-
r
conductor into the metal because the barrier has been reduced by a voltage VF.For a

Metal work function q& (eV)

Fig. 3 Measured barrier height br metal-silicon and metal-gallium arsenide contact^.^,^


228 b Chapter 7. MESFET and Related Devices

-
n type semiconductor p - type semiconductor

Fig. 4 Energy band diagrams of metal n-type and p-type semiconductors under different bias-
ing conditions: (a) thermal equilibrium; (b) forward bias; and (c) reverse bias.

reverse bias (i.e., a negative voltage is applied to the metal), the barrier has been increased
by a voltage VR,as depicted on the left side of Fig. 4c. It is more difficult for electrons
to flow from the semiconductor into the metal. We have similar results for p-type semi-
conductor, however, the polarities must be reversed. In the following derivations, we con-
sider only the metal-n-type semiconductor contact. The results are equally applicable to
a p-type semiconductor with an appropriate change of polarities.
The charge and field distributions for a metal-semiconductor contact are shown in
Fig. Sa and Sb, respectively. The metal is assumed to be a perfect conductor; the charge
transferred to it from the semiconductor exists in a very narrow region at the metal sur-
face. The extent of the space charge in the semiconductor is W, i.e., p, = qND for x c W
and ps= 0 for x > W. Thus, the charge distribution is identical to that of a one-sided abrupt
p+-njunction.
The magnitude of the electric field is decreasing linearly with distance. The maxi-
mum electric field c&, is located at the interface. The electric field distribution is then
given by
Chapter 7. MESFET and Related Devices 4 229

(b)
Fig. 5 (a) Charge distribution and (b) electric-field distribution in a metal-semiconductor
contact.

where E, is the dielectric permittivity of the semiconductor. The voltage across the space-
charge region, which is represented by the area under the field curve in the Fig. Sb, is
given by

The depletion-layer width W is expressed as

and the space-charge density, Qsc, in the semicoqductor is given as

where the voltage V equal to +V, for forward bias and to -V, for reverse bias. The
depletion-layer capacitance C per unit area can be calculated by using Eq. 9 :

and
230 a. Chapter 7. MESFET and Related Devices

We can differentiate 1/C2with respect to V. Rearranging terms we obtain :

Thus, measurements of the capacitance C per unit area as a function of voltage can pro-
vide the impurity distribution from Eq. 12. If N D is constant throughout the depletion
region, we should obtain a straight line by plotting 1/C2versus V. Figure 6 is a plot of
the measured capacitance versus voltage for tungsten-silicon and tungsten-gallium arsenide
Schottky diode^.^ From Eq. 11,the intercept at 1/C2= 0 corresponds to the built-in poten-
tial Vbi. Once Vbcis determined, the barrier height eBn
can be calculated from Eq. 4.

8 EXAMPLE 1
Find the donor concentration and the barrier height of the tungsten-silicon Schottky diode shown
in Fig. 6.

SOLUTION The plot of 1/C2 versus V is a straight line, which implies that the donor concentra-
tion is constant throughout the depletion region. We find

Fig. 6
Chapter 7. MESFET and Related Devices *I231

1 From Eq. 12,

2.86 x lo1'
V,, = 0.0259 x In
[ 2-7'x l0l5 ] = o,24

Since the intercept Vbiis 0.42 V, then the barrier height is ,$q = 0.42 + 0.24 = 0.66 V. 4

7.1.2 The Schottky Barrier


A Schottky barrier refers to a metal-semiconductor contact having a large barrier height
(i,e., qB,,or qBp>> kT)and a low doping concentration that is less than the density of
states in the conduction band or valence band.
The current transport in a Schottky barrier is due mainly to majority carrier, in con-
trast to a p-n junction, where current transport is due mainly to minority carriers. For
Schottky diodes operated at moderate temperature (e.g., 300 K), the dominate transport
mechanism is thermionic emission of majority carriers from the semiconductor over the
potential barrier into the metal.
Figure 7 illustrates the thermionic emission p r o ~ e s sAt
. ~ thermal equilibrium (Fig.
7a),the current density is balanced by two equal and opposite flows of carriers, thus there
is zero net currents. Electrons in the semiconductor tend to flow (or emit) into the metal,
and there is an opposing balanced flow of electrons from metal into the semiconductor.
These current components are proportional to the density of electrons at the boundary,
I As discussed in Section 3.5 of Chapter 3, at the semiconductor surface an electron
can be thermionically emitted into the metal if its energy is above the b a r n i height. Here
the semiconductor work function qqSis replaced by qqbBn,and

(a) (b) (4
;I Fig. 7 Current transport by the thermionic emission process. (a) Thermal equilibrium; (b) for-
[. ward bias; and ( c )reverse bias5 ,
232 +I Chapter 7. MESFET and Related Devices

where N , is the density of states in the conduction band. At thermal equilibrium we have

where],, is the current from the metal to the semiconductor, J,, is the current from
the semiconductor to the metal, and C, is a proportionality constant.
When a forward bias V, is applied to the contact (Fig, 7b), the electrostatic poten-
tial difference across the barrier is reduced, and the electron density at the surface increases
to

The current J,, that results from the electron flow out of the semiconductor is there-
fore altered by the same factor (Fig. 7b). The flux of electrons from the metal to the semi-
conductor, however, remains the same because the barrier qB, remains at its equilibrium
value. The net current under forward bias is then

Using the same argument for the reverse-bias condition(see Fig. 7c), the expression for
the net current is identical to Eq. 16 except that V, is replaced by -V,.
The coefficient CINc is found to be equal to A", where A* is called the efective
Richardson constant (in units of A/K2-cm2),and T is the absolute temperature. The value
of A* depend on the effective mass and are equal to 110 and 32 for n- and p-type sili-
con, respectively, and 8 and 74 for n- and p-type gallium arsenide, re~pectively.~
The current-voltage characteristic of a metal-semiconductor contact under thermionic
emission condition is then

where ], is the saturation current density and the applied voltage V is positive for for-
ward bias and negative for reverse bias. Experimental forward I-V characteristics of two
Schottky diodes4 are shown in Fig. 8. By extrapolating the forward I-V curve to V = 0 ,
we can findJ,. From], and Eq. 17a we can obtain the barrier height.
In addition to the majority carrier (electron) current, a minority-carrier (hole) cur-
rent exists in a metal n-type semiconductor contact because of hole injection from the
Chapter 7. MESFET and Related Devices 4 233

I Fig. 8 Forward current density versus applied voltage of W-Si and W-GaAs diodesS4
k
, metal to the semiconductor. The hole injection is the same as in a p+-n junction, which
is described in Chapter 4. The current density is given by

Jp =Jpo(e*vlkT- 1),

where

Y
Under normal operating conditions, the minority-carrier current is orders of magnitude
I
smaller than the majority-carrier current. Therefore, a Schottky diode is a unipolar device
t
(i.e., predominately only one type of carrier participates in the conduction process).

EXAMPLE 2
1 For a tungsten-silicon Schottky diode with N , = 1016~ m - find
~ , the barrier height and depletion-
layer width from Fig. 8. Compare the saturation current ], with,,] assuming that the minority-
carrier lifetime in Si is s. -
Chapter 7. MESFET and Related Devices

SOLUTION From Fig. 8, we have], = 6.5 x A/cm2.The barrier height can be obtained fro
Eq. 17a:
GB, = 0.0259 x In 'lox
( 6 . 5,~1 )
300' = 0,67 V,

This result is in the close agreement with the C-V measurement (see Fig. 6 and Ex. 1).
The built-in potential is given by GBn- Vn,where

Therefore,
Vbl= 0.67 - 0.17 = 0.50 V.
The depletion-layer width at thermal equilibrium is given by Eq. 8 with V = 0:

w=;4 2e~vbt 2.6 x


, cm,

To calculate the minority-carrier current densityJp,, we need to know Dp,which is 10 cm2/sfor


N , = 1016~ m -and
~ , Lp which is %= = 3.1 x cm. Therefore,

'
4Dpni - 1.6 x 10"' x 10 x (9.65 x 10')'
1p0 = -- = 4.8 x lo-'' ~/cm'.
L~ND (3.1 x lo-3) x lo16
The ratio of the two current densities is
J,= 6.5 x
= 1.3 x 10'.
lpo4.8x10-~'
From the comparison, we see that the majority-carrier current is over seven orders of magni-
tude greater than the minority-carrier current.

A.3 The Ohmic Contact


An ohmic contact is defined as a metal-semiconductor contact that has a negligible con-
tact resistance relative to the bulk or series resistance of the semiconductor. A satisfac-
tory ohmic contact should not significantly degrade device performance and can pass the
required current with a voltage drop that is small compared with the drop across the active
region of the device.
A figure-of-merit for ohmic contacts is the specific contact resistance Re, defined as

For metal-semiconductor contacts with low doping concentrations, the thermionic-emission


current dominates the current transport, as given by Eq. 17. Therefore,

Equation 20 shows that a metal-semiconductor contact with a low barrier height should
be used to obtain a snlall Re.
Chapter 7. MESFET and Related Devices 4 235

For contacts with high doping concentration, the barrier width becomes very nar-
row, and the tunneling current becomes dominate. The tunneling current, as described
in the upper inset of Fig. 9, is proportional to the tunneling probability, which is given
in Section 3,6 of Chapter 3:

where W is the depletion-layer width, which can be approximated as V(2&Jq~,)(qBn - V ),


rn, is the effective mass, and A is the reduced Planck constant. Substituting W into Eq,
21, we obtain

where C, equal 4 G J h . The specific contact resistance for high dopings is thus

( 10-l' cm3/z)

Fig. 9 Calculated and measured values of specific contact resistance. Upper inset shows the
tunneling process. Lower inset sbows thermionic emission over the low barrier.6
236 w Chapter 7. MESFET and Related Devices

Equation 23 shows that in the tunneling range the specific contact resistance depends
strongly on doping concentration and varies exponentially with the factor @Bn .
The calculated values of R, are plotted6 in Fig. 9 as a function of I/%. For N D >
1019~ m -R, ~ ,is dominated by the tunneling process and decreases rapidly with increase
doping. On the other hand, for ND < 101'cm3, the current is due to thermionic emis-
sion, and Rc is essentially independent of doping. Also shown in Fig. 9 are experimen
tal data for platinum silicide-silicon (PtSi-Si) and aluminum-silicon (Al-Si)diodes. They
are in close agreement with the calculated values. Figure 9 shows that a high doping con-
centration, a low barrier height, or both must be used to obtain a low value of Rc. These
two approaches are used for all practical ohmic contacts.

b EXAMPLE 3
An ohmic contact has an area of 105 cm2and a specific contact resistance of n-cm2. The ohmic
contact is formed in an n-type silicon. If ND = 5 x 1019 cm3, and q&, = 0.8 V, and the electron
effective mass is 0.26 ?no,find the voltage drop across the contact when a forward current of 1A
flows through it.

SOLUTION The contact resistance for the ohmic contact is

%. = 1 0 - 6 -~ cm2/10-5 cm2 = 10-la,


A

440~26x 9.1 x lo4' x ( 1 . 0 5 10-lo)


~
C, = 4ljm,E,/h =
1.05x
= 1.9 x 10" (rnj', I V ] .

From Eq. 22,

or

I. =
A IhG
Rc

= lox [d 1.9 x 1014


m
1 1.9 x l0l4 x 0.8
5 x 10IQx lo6
=8.13x lo8 A.
at I = lA, we have

or
V = 0.8 - 0.763 =0.037'V = 37 mV.
Hu_ch04v4.fm Page 133 Friday, February 13, 2009 5:54 PM

4.16 ● Schottky Barriers 133

destination after different times of travel. Consequently, a short LED pulse at the
originating point would arrive at the destination as a longer broadened pulse. For
this reason, lasers, with their extraordinary purity of wavelength, are the light
source of choice for long-distance high data rate links.

4.15 ● PHOTODIODES ●

Figure 4–25 shows that a reverse current flows through a diode when illuminated
with light and the current is proportional to the light intensity. A reverse-biased PN
diode can thus be used to detect light, and the device is called a photodiode. If the
photodiode is biased near the avalanche breakdown voltage, photo-generated
carriers are multiplied by impact ionization as they travel through the depletion
layer (see Fig. 4–13) and thereby the sensitivity of the detector is increased. This
device is called an avalanche photodiode. Photodiodes are used for optical
communication, DVD reader, and other light-sensing applications.

PART III: METAL–SEMICONDUCTOR JUNCTION


There are two kinds of metal–semiconductor junction. The junctions between metal
and lightly doped semiconductors exhibit rectifying IV characteristics similar to
those of PN junctions. They are called Schottky diodes and have some interesting
applications. The junction between metal and heavily doped semiconductors
behaves as low-resistance ohmic contacts (basically electrical shorts). Ohmic
contacts are an important part of semiconductor devices and have a significant
influence on the performance of high-speed transistors.

4.16 ● SCHOTTKY BARRIERS ●

The energy diagram of a metal–semiconductor junction is shown in Fig. 4–34. The


Fermi level, EF, is flat because no voltage is applied across the junction. Far to the
right of the junction, the energy band diagram is simply that of an N-type silicon
sample. To the left of the junction is the energy band diagram of a metal—with the
energy states below EF almost totally filled and the states above EF almost empty.
The most striking and important feature of this energy diagram is the energy barrier
at the metal–semiconductor interface. It is characterized by the Schottky barrier
height, φB. φB is a function of the metal and the semiconductor. Actually, there are
two energy barriers. In Fig. 4–34a, qφBn is the barrier against electron flow between
the metal and the N-type semiconductor.6 In Fig. 4–34b, qφBp is the barrier against
hole flow between the metal and the P-type semiconductor. In both figures, there is
clearly a depletion layer adjacent to the semiconductor–metal interface, where EF is
close to neither Ec nor Ev (such that n ≈ 0 and p ≈ 0).

6 The hole flow in Fig. 4–34a is usually insignificant because there are few holes in the N-type
semiconductor.
Hu_ch04v4.fm Page 134 Friday, February 13, 2009 5:54 PM

134 Chapter 4 ● PN and Metal–Semiconductor Junctions

Depletion
Metal Neutral region
layer

qfBn
Ec
EF

Ev
(a)

Ec

EF
Ev
qfBp

(b)
FIGURE 4–34 Energy band diagram of a metal–semiconductor contact. The Schottky barrier
heights depend on the metal and semiconductor materials. (a) φBn is the barrier against
electron flow between the metal and the N-type semiconductor; (b) φBp is the barrier against
hole flow between the metal and the P-type semiconductor.

TABLE 4–4 Measured Schottky barrier heights for electrons on N-type silicon (φBp)
and for holes on P-type silicon (φ Bp). (From [7].)

Metal Mg Ti Cr W Mo Pd Au Pt
φBn (V) 0.4 0.5 0.61 0.67 0.68 0.77 0.8 0.9
φBp (V) 0.61 0.50 0.42 0.3
Work Function 3.7 4.3 4.5 4.6 4.6 5.1 5.1 5.7
ψM (V)

It will become clear later that φB is the single most important parameter of a
metal–semiconductor contact. Table 4–4 presents the approximate φBn and φBp for
several metal–silicon contacts. Please note that the sum of qφBn and qφBp is
approximately equal to Eg (1.12 eV), as suggested by Fig. 4–35.
φ Bn + φ Bp ≈ E g (4.16.1)
Why does φBn (and φBp) vary with the choice of the metal? Notice that Table 4–4 is
arranged in ascending order of φBn. There is a clear trend that φBn increases with
increasing metal work function (last row in Table 4–4). This trend may be partially
explained with Fig. 4–2a.
φ Bn = ψ M – χ Si (4.16.2)
Hu_ch04v4.fm Page 135 Friday, February 13, 2009 5:54 PM

4.16 ● Schottky Barriers 135

Vacuum level, E 0

xSi ⫽ 4.05 eV
qcM

qfBn
Ec
EF

Ev
(a)

Vacuum level, E 0
xSi ⫽ 4.05 eV
qcM

qfBn
Ec
⫹ ⫺ EF

Ev
(b)
FIGURE 4–35 (a) An “ideal” metal–semiconductor contact and (b) in a real
metal–semiconductor contact, there is a dipole at the interface.
ψM is the metal work function and χSi is the silicon electron affinity. See Sec. 5.1 for
more discussion of these two material parameters. Equation (4.16.2) suggests that
φBn should increase with increasing ψM (in qualitative agreement with Table 4–4) by
1 eV for each 1 eV change in ψM (not in quantitative agreement with Table 4–4). The
explanation for the quantitative discrepancy is that there are high densities of energy
states in the band gap at the metal–semiconductor interface.7 Some of these energy
states are acceptor like and may be neutral or negative. Other energy states are
donor like and may be neutral or positive. The net charge is zero when the Fermi
level at the interface is around the middle of the silicon band gap. In other words,
Eq. (4.16.2) is only correct for ψM around 4.6V, under which condition there is little
interface charge. At any other ψM, there is a dipole at the interface as shown in

7 In a three-dimensional crystal, there are no energy states in the band gap. Not so at the metal–semiconductor
interface.
Hu_ch04v4.fm Page 136 Friday, February 13, 2009 5:54 PM

136 Chapter 4 ● PN and Metal–Semiconductor Junctions

TABLE 4–5 Measured Schottky barrier heights of metal silicide on Si.

Silicide ErSi1.7 HfSi MoSi2 ZrSi2 TiSi2 CoSi2 WSi2 NiSi2 Pd2Si PtSi
φBn (V) 0.28 0.45 0.55 0.55 0.61 0.65 0.67 0.67 0.75 0.87
φBp (V) 0.55 0.55 0.49 0.45 0.43 0.43 0.35 0.23

Fig. 4–35b and it prevents φBn from moving very far from around 0.7 V. This
phenomenon is known as Fermi-level pinning. Table 4–4 can be approximated with
φ Bn = 0.7 V + 0.2 ( ψ M – 4.75 ) (4.16.3)
The factor of 0.2 in Eq. (4.16.3) is determined by the polarizability of Si and the
energy state density at the metal–silicon interface [8].

● Using C–V Data to Determine φ B ●


In Fig. 4–36a, φbi is the built-in potential across the depletion layer.
N
q φ bi = q φ Bn – ( E c – E F ) = q φ Bn – kT ln ------c- (4.16.4)
Nd

The depletion-layer thickness is [see Eq. (4.3.1)]

2 ε s ( φ bi + V )
W dep = ------------------------------- (4.16.5)
qN d

qfbi
qfBn
Ec
EF

Ev
(a)

qfBn q(fbi ⫹ V)

qV
Ec
EF

Ev
(b)

FIGURE 4–36 The potential across the depletion layer at the Schottky junction. (a) No voltage
applied; (b) a negative voltage (reverse bias) is applied to the metal.
Hu_ch04v4.fm Page 137 Friday, February 13, 2009 5:54 PM

4.17 ● Thermionic Emission Theory 137

εs
C = A ------------- (4.16.6)
W dep
1 2 ( φ bi + V )
------ = -------------------------
- (4.16.7)
2 2
C qN d ε s A

Figure 4–37 shows how Eq. (4.16.7) allows us to determine φbi using measured C–V
data. Once φbi is known, φBn can be determined using Eq. (4.16.4).

1/C 2

V
⫺fbi

FIGURE 4–37 φbi (and hence φB) can be extracted from the C–V data as shown.

Much more prevalent in IC technology than metal–Si contacts are the silicide–Si
contacts. Metals react with silicon to form metal like silicides at a moderate
temperature. Silicide–Si interfaces are more stable than the metal–Si interfaces and
free of native silicon dioxide. After the metal is deposited on Si by sputtering or
CVD (Chemical Vapor Deposition) (see Chapter 3), an annealing step is applied to
form a silicide–Si contact. The term metal–silicon contact is understood to include
silicide–silicon contacts. Table 4–5 shows some available data of φBn and φBp of
silicide–silicon contacts.

4.17 ● THERMIONIC EMISSION THEORY ●

Figure 4–38 presents the energy band diagram of a Schottky contact with a bias V
applied to the metal. Let us analyze the current carried by the electrons flowing
from Si over the energy barrier into metal, J S → M. This current can be predicted
quite accurately by the thermionic emission theory.
In the thermionic emission theory, we assume that EFn is flat all the way to the
peak of the barrier, the electron concentration at the interface (using Eqs. (1.8.5)
and (1.8.6)) is
2 π m n kT 3⁄2
–q ( φ B – V) ⁄ kT –q ( φ B – V) ⁄ kT
n = Nce = 2 ---------------------
- e (4.17.1)
2
h
The x-component of the average electron velocity is of course smaller than the total
thermal velocity, 3kT ⁄ m n [Eq. (2.1.3)], and only half of the electrons travel
Hu_ch04v4.fm Page 138 Friday, February 13, 2009 5:54 PM

138 Chapter 4 ● PN and Metal–Semiconductor Junctions

vthx

q(fB⫺V)
Ec
qfB
EFn
N-type EFm qV
V Metal
silicon

Ev

x
FIGURE 4–38 Energy band diagram of a Schottky contact with a forward bias V applied
between the metal and the semiconductor.

toward the left (the metal). It can be shown that the average velocity of the left-
traveling electrons is
v thx = – 2kT ⁄ π m n (4.17.2)
Therefore,
2
1 4 π qm n k 2 –q φB ⁄ kT qV ⁄ k T
-T e
J S → M = – --- qnv thx = ----------------------- e (4.17.3)
2 3
h
qV ⁄ k T
≡ J0e (4.17.4)
–q φ B ⁄ kT
Equation (4.17.4) carries two notable messages. First J0 ≈ 100 e (A/cm2)
is larger if φB is smaller. Second, J S → M is only a function of φB – V (see Fig. 4–38).
The shape of the barrier is immaterial as long as it is narrow compared to the carrier
mean free path. φB – V determines how many electrons possess sufficient energy to
surpass the peak of the energy barrier and enter the metal.

4.18 ● SCHOTTKY DIODES ●

At zero bias (Fig. 4–39a), the net current is zero because equal (and small) numbers
of electrons on the metal side and on the semiconductor side have sufficient energy
to cross the energy barrier and move to the other side. The probability of finding an
–( E – Ec ) ⁄ kT –q φ B ⁄ kT
electron at these high-energy states is e = e on both sides of the
junction, as shown in Fig. 4–39a. Therefore, the net current is zero.8 In other words,
I S → M = I0 and I M → S = –I0, where I S → M and I M → S (see Fig. 4–39a) represent

8 What if the densities of states are different on the two sides of the junction? Assume that the density of
states at E on the metal side is twice that on the silicon side. There would be twice as many electrons on
the metal side attempting to cross the barrier as on the Si side. On the other hand, there would be twice
as many empty states on the metal side to receive the electrons coming from the Si side. Therefore, in a
more detailed analysis, the net current is still zero.
Hu_ch04v4.fm Page 139 Friday, February 13, 2009 5:54 PM

4.18 ● Schottky Diodes 139

IM S ⫽ ⫺I0 IS M ⫽ I0 IM S ⫽ ⫺I0 IS M ⫽ I0 eqV/kT


⫺ ⫺ ⫺ ⫺
⬍qfB
qfB E ⫺ EF ⫽ qfB qfB EFn
qV
EF

(a) V = 0. IS M ⫽ IM S ⫽ I0 (b) Forward bias. Metal is positive wrt


Si. IS M ⬎⬎ IM S ⫽ I0

IM S⫽ ⫺I0 IS M ⬇0

qfB
⬎qfB I
qV

EFn

Reverse bias Forward bias

(c) Reverse bias. Metal is negative wrt Si. (d) Schottky diode IV.
IS M ⬍⬍ IM S ⫽ I0

FIGURE 4–39 Explanation of the rectifying IV characteristics of Schottky diodes. The arrows
in the subscripts indicate the direction of electron flows.

the electron current flowing from Si to metal and from metal to Si, respectively.
According to the thermionic emission theory,
2 –q φ B ⁄ kT
I 0 = AKT e (4.18.1)
A is the diode area and
2
4 π qmn k
K = -----------------------
- (4.18.2)
3
h
K ≈ 100 A/(cm2/K2) is known as the Richardson constant. In Fig. 4–39b, a positive
bias is applied to the metal. I M → S remains unchanged at –I0 because the barrier
against I M → S remains unchanged at φB. I S → M, on the other hand, is enhanced by
qV ⁄ kT because the barrier is now smaller by qV. Therefore,
e
2 –( q φ B – qV) ⁄ kT 2 –q φ B ⁄ kT qV ⁄ kT qV ⁄ kT
IS → M = AKT e = AKT e e = I0e (4.18.3)
qV ⁄ kT qV ⁄ kT
I = IS → M + IM → S = I0e – I0 = I0( e – 1) (4.18.4)
Hu_ch04v4.fm Page 140 Friday, February 13, 2009 5:54 PM

140 Chapter 4 ● PN and Metal–Semiconductor Junctions

In summary,

qV ⁄ kT (4.18.5)
I = I0( e – 1)
2 –q φ B ⁄ kT (4.18.6)
I 0 = AKT e

Equation (4.18.5) is applicable to the V < 0 case (reverse bias, Fig. 4–39c) as well.
For a large negative V, Eq. (4.18.5) predicts I = –I0. Figure 4–39c explains why:
I S → M is suppressed by a large barrier, while I M → S remains unchanged at –I0.
Equation (4.18.5) is qualitatively sketched in Fig. 4–39d. I0 may be extracted using
Eq. (4.18.5) and the IV data. From I0, φB can be determined using Eq. (4.18.6).
The similarity between the Schottky diode IV and the PN junction diode IV is
obvious. The difference will be discussed in Section 4.19.

4.19 ● APPLICATIONS OF SCHOTTKY DIODES ●

Although Schottky and PN diodes follow the same IV expression


qV ⁄ kT
I = I0( e – 1) , (4.19.1)

I0 of a silicon Schottky diode can be 103–108 times larger than a typical PN junction
diode, depending on φB (i.e., the metal employed). A smaller φB leads to a larger I0.
A larger I0 means that a smaller forward bias, V, is required to produce a given
diode current as shown in Fig. 4–40.
This property makes the Schottky diode the preferred rectifier in low-voltage
and high-current applications where even a ~0.8 V forward-voltage drop across a
PN junction diode would produce an undesirably large power loss. Figure 4–41
illustrates the switching power supply as an example. After the utility power is
rectified, a 100 kHz pulse-width modulated (square-wave) AC waveform is
produced so that a small (lightweight and cheap) high-frequency transformer can
down-transform the voltage. This low-voltage AC power is rectified with Schottky
diode (~0.3 V forward voltage drop) and filtered to produce the 50 A, 1 V, 50 W DC

Schottky diode
I

fB
PN
diode

FIGURE 4–40 Schematic IV characteristics of PN and Schottky diodes having the same area.
Hu_ch04v4.fm Page 141 Friday, February 13, 2009 5:54 PM

4.20 ● Quantum Mechanical Tunneling 141

PN junction Schottky
rectifier Transformer rectifier
110/220 V 100 kHz 50 A
Hi-voltage Hi-voltage Lo-voltage
AC 1 V DC
DC DC–AC AC AC
Utility
Inverter
power

Feedback to modulate the pulse width to keep Vout ⫽ 1V


FIGURE 4–41 Block diagram of a switching power supply for electronic equipment such as PCs.

output. If a PN diode with 0.8 V forward voltage drop is used, it would consume
40 W (50 A × 0.8 V) of power and require a larger fan to cool the equipment.
For this application, a Schottky contact with a relatively small φB would be
used to obtain a large I0 and a small forward voltage drop. However, φB cannot be
too small, or else the large I0 will increase the power loss when the diode is reverse
biased and can cause excessive heat generation. The resultant rise in temperature
will further raise I0 [Eq. (4.18.1)] and can lead to thermal runaway.

● The Transistor as a Low Voltage-Drop Rectifier ●


Even a Schottky diode’s forward voltage may be too large when the power-supply
output voltage is, say, 1V. One solution is to replace the diode with a MOSFET
transistor [9]. A MOSFET is essentially an on–off switch as shown in Fig. 6–2. A low-
power circuit monitors the voltage polarity across the transistor and generates a signal
to turn the switch (transistor) on or off. In this way, the transistor, with the control
circuit, functions as a rectifier and is called a synchronous rectifier. The MOSFET in this
application would have a very large channel width in order to conduct large currents.
The important point to note is that a MOSFET is not subjected to the same trade-off
between the reverse leakage current and forward voltage drop as a diode [Eq. (4.18.5)].

The second difference between a Schottky diode and a PN junction diode is


that the basic Schottky diode operation involves only the majority carriers (only
electrons in Fig. 4–39, for example). There can be negligible minority carrier
injection at the Schottky junction (depending on the barrier height). Negligible
injection of minority carriers also means negligible storage of excess minority
carriers (see Section 4.10). Therefore, Schottky diodes can operate at higher
frequencies than PN junctiondiodes.
Schottky junction is also used as a part of a type of GaAs transistor as described
in Section 6.3.2.

4.20 ● QUANTUM MECHANICAL TUNNELING ●

Figure 4–42 illustrates the phenomenon of quantum mechanical tunneling. Electrons,


in quantum mechanics, are represented by traveling waves. When the electrons arrive
at a potential barrier with potential energy (VH) that is higher than the electron
Hu_ch04v4.fm Page 142 Friday, February 13, 2009 5:54 PM

142 Chapter 4 ● PN and Metal–Semiconductor Junctions

E ⬎ VL E ⬍ VH E ⬎ VL
Incident electron VH
wave with energy E
Transmitted
electron wave

T
VL
FIGURE 4–42 Illustration of quantum mechanical tunneling.

energy (E), the electron wave becomes a decaying function. Electron waves will
emerge from the barrier as a traveling wave again but with reduced amplitude. In
other words, there is a finite probability for electrons to tunnel through a potential
barrier. The tunneling probability increases exponentially with decreasing barrier
thickness [10] as

 8π m
2 
P ≈ exp  – 2T -------------- ( V H – E ) (4.20.1)
2
 h 
where m is the effective mass and h is the Planck’s constant. This theory of
tunneling will be used to explain the ohmic contact in the next section.

4.21 ● OHMIC CONTACTS ●

Semiconductor devices are connected to each other in an integrated circuit


through metal. The semiconductor to metal contacts should have sufficiently low
resistance so that they do not overly degrade the device performance. Careful
engineering is required to reach that goal. These low-resistance contacts are
called ohmic contacts. Figure 4–43 shows the cross-section of an ohmic contact. A
surface layer of a heavily doped semiconductor diffusion region is converted into
a silicide such as TiSi2 or NiSi2 and a dielectric (usually SiO2) film is deposited.

Metal 1 (AICu)

W-plug

Oxide
TiN

TiSi2
n⫹ Diffusion region
FIGURE 4–43 A contact structure. A film of metal silicide is formed before the dielectric-
layer deposition and contact-hole etching. (From [11]. © 1999 IEEE.)
Hu_ch04v4.fm Page 143 Friday, February 13, 2009 5:54 PM

4.21 ● Ohmic Contacts 143

Lithography and plasma etching are employed to produce a contact hole through
the dielectric reaching the silicide. A thin conducting layer of titanium nitride
(TiN) is deposited to prevent reaction and interdiffusion between the silicide and
tungsten. Tungsten is deposited by CVD to fill the contact hole. Figure 4–43 also
shows what goes on top of the W plug: another layer of TiN and a layer of AlCu
as the interconnect metal material.
An important feature of all good ohmic contacts is that the semiconductor is
very heavily doped. The depletion layer of the heavily doped Si is only tens of Å
thin because of the high dopant concentration.
When the potential barrier is very thin, the electrons can pass through the
barrier by tunneling with a larger tunneling probability as shown in Fig. 4–44. The
tunneling barrier height, VH – E in Eq. (4.20.1) is simply φ Bn . The barrier thickness
T may be taken as
T ≈ W dep ⁄ 2 = ε s φ Bn ⁄ ( 2qN d ) (4.21.1)

–Hφ Bn ⁄ Nd
P≈e (4.21.2)

H ≡ ------ ( ε s m n ) ⁄ q (4.21.3)
h
At V = 0, J S → M and J M → S in Fig. 4–44a are equal but of opposite signs so
that the net current is zero.
1
J S → M(=–J M → S ) ≈ --- qN d v thx P (4.21.4)
2
Only half of the electrons in the semiconductor, with density Nd/2, are in thermal
motion toward the junction. The other half are moving away from the junction. vthx
may be found in Eq. (4.17.2). Assuming that Nd = 1020 cm3, P would be about
0.1and J S → M ≈ 108 A/cm2. (This is a very large current density.) If a small voltage is
applied across the contact as shown in Fig. 4–44b, the balance between J S → M and
J M → S is broken. The barrier for J M → S is reduced from φBn to (φBn – V).

1 –H( φ Bn – V) ⁄ Nd
J S → M = --- qN d v thx e (4.21.5)
2

Silicide N⫹Si
fBn ⫺ V
fBn
⫺ ⫺
⫺ ⫺ Ec, EFn
EFm V
Ec, EF

Ev
Ev
x x
(a) (b)
+
FIGURE 4–44 (a) Energy band diagram of metal–N Si contact with no voltage applied and
(b) the same contact with a voltage, V, applied to the contact.
Hu_ch04v4.fm Page 144 Friday, February 13, 2009 5:54 PM

144 Chapter 4 ● PN and Metal–Semiconductor Junctions

At small V, the net current density is


d JS → M 1 –Hφ Bn ⁄ Nd
J ≈ -------------------
- ⋅ V = V ⋅ --- qv thx H N d e (4.21.6)
dV V = 0 2
Hφ ⁄ N
2⋅e
Bn d
V
R c ≡ ---- = -------------------------------- (4.21.7)
J qv thx H N d
Hφ Bn ⁄ Nd
∝e (4.21.8)
2 2
Rc is the specific contact resistance (Ω cm ), the resistance of a 1 cm contact. Of course,
Eq. (4.21.8) is applicable to P+ semiconductor contacts if φBn, mn, and Nd are replaced
by φBp, mp, and Na. Figure 4–45 shows the IV characteristics of a silicide–Si contact. The
IV relationship is approximately linear, or ohmic in agreement with Eq. (4.21.6). The
resistance decreases with increasing temperature in qualitative agreement with Eq.
(4.21.7), due to increasing thermal velocity, vthx. The contact resistance is 140 Ω and
Rc ≈ 107 Ω cm2. The Rc model embodied in Eq. (4.2.7) is qualitatively accurate, but B
and H are usually determined experimentally9 [11]. Rc calculated from a more complex
model is plotted in Fig. 4–46. If we want to keep the resistance of a 30 nm diameter
contact below 1 kΩ, Rc should be less than 7 × 10–10 Ω cm2. This will require a very high
doping concentration and a low φB. Perhaps two different silicides will be used for N+
and P+ contacts, since a single metal cannot provide a low φBn and a low φBp.
1.2

1
300 ⬚C

0.8 25 ⬚C
Current (mA)

0.6
n ⫹ Si
0.4 p ⫹ Si

0.2
0.3 ␮m
0
0 0.05 0.1 0.15 0.2
Voltage (V)
FIGURE 4–45 The IV characteristics of a 0.3 µm (diameter) TiSi2 contact on N+-Si and P+-Si.
(From [11] ©1999 IEEE.)

● Boundary Condition at an Ohmic Contact ●


The voltage across an ideal ohmic contact is zero. This means that the Fermi level cannot
deviate from its equilibrium position, and therefore n' = p' = 0 at an ideal ohmic contact.

9 The electron effective mass in Eq. (4.21.2) is not equal to m (effective mass of electron in the conduc-
n
tion band) while it is tunneling under the barrier (in the band gap). Also Eq. (4.17.2) overestimates vthx
for a heavily doped semiconductor, for which the Boltzmann approximation is not valid.
Hu_ch04v4.fm Page 145 Friday, February 13, 2009 5:54 PM

4.22 ● Chapter Summary 145

1.E⫺06

0.6 V
0.5 V
fB
0.4 V
0.3 V
Specific contact resistance (⍀-cm2)

1.E⫺07

1.E⫺08

1.E⫺09
5E⫹19 1.5E⫹20 2.5E⫹20 3.5E⫹20 4.5E⫹20 5.5E⫹20
Surface doping density (1/cm3)

FIGURE 4–46 Theoretical specific contact resistance. (After [12].)

4.22 ● CHAPTER SUMMARY ●

PART I: PN JUNCTION

It is important to know how to draw the energy band diagram of a PN junction. At


zero bias, the potential barrier at the junction is the built-in potential,

kT N d N a
φ bi = ------- ln -------------- (4.1.2)
q 2
n i

The potential barrier increases beyond φbi by 1V if a 1V reverse bias is applied and
decreases by 0.1V if a 0.1V forward bias is applied.
The width of the depletion layer is

2 ε s × potential barrier
W dep = --------------------------------------------------------- (4.3.1)
qN
N is basically the smaller of the two doping concentrations. The main significance of
Wdep is that it determines the junction capacitance.
εs
C dep = A ------------- (4.4.1)
W dep

You might also like