1 Construction and Mechanism of Operation

advertisement
1.
1.1
Construction and Mechanism of Operation
Physical Structure
The physical structure of an n-type MOS Transistor is shown in
Fig. 1.1. This can be seen to be composed of two heavily doped n-type
regions, referred to as the source and drain, diffused into a lightly
doped p-type substrate, also known as the bulk or body. These are
spaced a distance L m apart and extend across the geometry for a
width W m. Between these two regions, on the surface of the
substrate, is deposited an insulating layer of Silicon Dioxide, SiO2, of
thickness tox, usually between 10 and 50 nm. Finally, this oxide layer is
covered with a metal layer (or in more recent years polysilicon)
referred to as the gate. Metal connections are then made to the gate,
source and drain as terminals. There is normally a connection to the
body also and this can be made for individual transistors when
required in today’s technologies.
The three central layers essentially form a parallel plate
capacitor and it is this layered structure which gives the device its
name, the Metal-Oxide-Semiconductor or MOS Transistor. This
capacitor has a capacitance per unit area of Cox = ox/tox. It is the
action of creating an electric field across the capacitor to control a
conducting channel in the silicon which forms the basis of operation of
the device and gives it its full name Metal-Oxide-Semiconductor-FieldEffect-Transistor or MOSFET.
The n-type, source and drain regions are reasonably highly doped
at a concentration of between 1016 and 1018 cm-3 while the p-type
substrate is lightly doped at a concentration of the order of 10 13 cm-3.
The device shown in Fig. 1.1 is referred to as an n-type MOSFET
because, as will be seen, the conducting channel formed is n-type. A ptype MOSFET can be fabricated by diffusing p-type source and drain
regions into an n-type substrate.
1.2
Device Operation
Voltage sources are normally connected to the n-type MOS
transistor as shown in Fig. 1.2(a), which presents a vertical section
view through the centre of the device. The voltage connected between
gate and source is known as VGS while that connected between drain
and source is known as VDS. A voltage may also be connected between
the source and the body as shown, but very often the source and body
terminals are connected together.
1
Body B
Gate G
Source S
Drain D
W
source
n+
drain
n+
L
substrate
p-
Silicon Dioxide SiO2
Fig.1.1
Physical Structure of the MOSFET
2
(i) All Terminals Grounded, VGS = 0 and VDS = 0
Consider initially the situation where all terminals of the MOSFET,
including the body or substrate are connected to ground. Under this
condition, p-n junctions will be formed between the source-substrate
and drain-substrate regions, which appear essentially back-to-back.
Consequently, depletion regions are formed around the source and
drain as shown in Fig.1.2(a) and there is an extremely high resistance
to current flow between these two terminals. The depletion regions
contain positive ions on the n-type side and negative ions on the ptype side of the junctions. As the substrate is more lightly doped, the
depletion regions will extend further into the substrate than into the
source or drain regions.
The energy band diagram for the metal, insulator and semiconductor
sandwich is shown in Fig. 1.2(b). When these three materials are
brought together, their individual energy bands must realign so that
the Fermi level is constant throughout all three materials. This
realignment takes place in a different manner to the junction formed
in the bipolar transistor, as carriers cannot pass through the insulating
dielectric. Instead an electric field is built up across the insulating
oxide layer and at its boundary with the substrate. This field is
dependent on the difference in the work function of the metal, qM and
that of the semiconductor, qS. The work function is the energy
required to remove an electron from the Fermi level in a specific
material and is quoted for the material in isolation. The electric field
gives rise to the gradient in the energy profile of the oxide from the
metal towards the semiconductor as shown. It also gives rise to a
bending of the energy levels in the surface region of the
semiconductor just below the oxide. Notice that this has the effect of
bringing the Fermi level at the doped semiconductor surface closer to
the Fermi level of intrinsic silicon. The Fermi potential, F, arises from
the difference between the Fermi level of the p-type substrate in the
MOS structure and that of intrinsic silicon, which is dependent on the
doping concentration NA and is given as:
F

EF  Ei kT  ni

ln 
q
q  NA



This is a negative quantity for p-type silicon. At the surface of the
semiconductor this potential is known as the surface potential, S,
which can be seen to be less than the Fermi potential because of the
bending of the energy bands in this region.
3
VDS = 0V
VSB = 0V
B
VGS = 0V
insulating
layer SiO2
depletion
region
G
D
S
n+
n+
pFig. 1.2(a)
Conditions in the MOSFET with all Terminals Grounded
Metal
Semiconductor
Oxide
+ E Field
-
EC
Ei
EF
qS
S
M
qF
EFS
EV
Fig. 1.2(b) Energy Band Diagram MOSFET with all Terminals Grounded
4
(ii) Small Applied Gate Bias, VGS > 0 and VDS = 0
Next, consider a small positive bias voltage applied to the gate,
while the drain-to-source voltage is kept at zero. This will cause
positive charge to be accumulated on the gate electrode and will
increase the electric field across the insulator. This has the effect of
repelling the mobile holes back into the substrate material leaving
behind negatively ionised dopant atoms at the surface of the
semiconductor to balance the charge accumulated on the gate
electrode. In reality, thermally available electrons occupy the vacant
energy levels in the dopant acceptor atoms under the influence of the
voltage applied to the gate, which holds them in place. This now
causes the depletion region to extend along the surface of the
semiconductor between the source and drain as shown in Fig. 1.3(a).
The energy band diagram under this condition is shown in Fig.
1.3(b). It can be seen that the Fermi levels in the metal and the
semiconductor now separate under the influence of the external
potential applied to the gate, VGS, by an amount equal to qVGS. It can
also be seen that increased bending of the energy bands at the surface
of the semiconductor brings the Fermi level of the substrate nearer to
the intrinsic Fermi level. With further increase in the gate voltage, the
width of the depletion region at the semiconductor surface increases.
Simultaneously, the energy bands continue to bend until the Fermi
level in the semiconductor becomes equal to the intrinsic level when S
= 0. At this point the depletion region under the gate is almost devoid
of free charge carriers.
(iii) Increased Gate Bias, Weak Inversion, VGS < VT, VDS = 0
Fig. 1.4(a) shows the situation as the gate voltage is increased
further and Fig. 1.4(b) shows how the energy bands in the
semiconductor bend further so that the Fermi level at the surface now
becomes greater than the intrinsic Fermi level. Under this condition,
more electrons are attracted into the region of semiconductor
underneath the oxide than are required to ionise the p-type dopant
atoms. These electrons are readily supplied by the heavily doped ntype source and drain regions to form a thin conducting layer directly
underneath the oxide known as the channel. This process is known as
inversion whereby the originally p-type semiconductor now becomes
essentially n-type in the channel region, under the influence of the
electric field resulting from the applied gate voltage. If a voltage is
applied between the drain and the source at this point a current will
flow between them. Initially, a channel of very low conductivity is
formed and the resulting current is very small and this state is referred
to as weak inversion.
5
VDS = 0V
VSB = 0V
VGS > 0V
depletion
region
G
B
D
S
n+
n+
pFig. 1.3(a)
Conditions in MOSFET, Small Gate Bias VGS >0 and VDS = 0
Semiconductor
Oxide
Metal
+ E Field
-
EC
Ei
qS
EF
qVGS
qF
EFS
EV
M
Fig. 1.3(b) Energy Band Diagram for Small Gate Bias VGS >0, VDS = 0
6
VDS = 0V
VSB = 0V
VGS > 0V
conducting
channel
G
B
D
S
n+
n+
pFig. 1.4(a) Conditions at Weak Inversion, 0 <VGS <VT and VDS = 0
Semiconductor
Oxide
Metal
+ E Field
-
EC
Ei
qF
qS
EFS
EV
qVGS
EF
M
Fig. 1.4(b) Energy Band Diagram at Weak Inversion, 0<VGS< VT,VDS=0
7
(iv) Increased Gate Bias, Strong Inversion, VGS > VT
If the gate voltage is increased even further, the n-type
conducting channel at the surface of the semiconductor becomes more
heavily impregnated with free electrons and therefore becomes more
conducting as shown in Fig. 1.5(a). This allows a much greater current
to flow between drain and source if a voltage is applied between these
two terminals. This state is known as strong inversion. In practice
there is a gradual change from the weak inversion state to the strong
inversion state, although channel conductivity does rise notably once
strong inversion is reached. Further increase in the gate voltage above
this value simply leads to increased conductivity of the induced
channel and higher levels of current flow between drain and source for
a given potential difference between them.
1.3 Threshold Voltage VT
For the purposes of device modelling and circuit analysis it is
highly desirable to have a definite point marking the transition from
weak to strong inversion. The point used is known as the threshold
voltage, VT, of the MOSFET.
The threshold voltage is defined, for an n-channel device, as the gatesource voltage which causes the n-type conducting channel to have
the same concentration of electrons as the p-type substrate has of
holes.
The energy band diagram of Fig. 1.5(b) shows that at the
threshold voltage, the energy bands have bended to bring the Fermi
level above the intrinsic Fermi level at the surface of the
semiconductor by the same amount as it is below it in the p-type
region away from the surface. At this point S = -F. An expression for
the threshold voltage will be derived later. At this point also the
depletion region reaches its maximum depth. Any further increase in
the gate-source voltage above the threshold value gives negligible
increase in the depth of the depletion region.
1.4 CMOS Structures
The MOS transistor shown in Fig. 1.1 is an n-channel device, so
called because the induced conducting channel is n-type and the
charge carriers are electrons. Note that current flow is due entirely to
electrons and so this device can be thought of as a unipolar transistor.
A p-channel MOS transistor can be fabricated by diffusing heavily
doped p-type source and drain regions into a lightly doped n-type
substrate. In such a device the induced channel will be p-type and
current flow will be entirely due to holes.
8
VDS = 0V
VSB = 0V
VGS =VT
G
B
D
S
n+
n+
pFig. 1.5(a)
Conditions at Onset of Strong Inversion, VGS =VT, VDS = 0
Semiconductor
Oxide
Metal
+ E Field
-
EC
Ei
qF
qS = -qF
EFS
EV
qVGS
EF
M
Fig. 1.5(b) Energy Bands at Onset of Strong Inversion, VGS = VT, VDS=0
9
CMOS or Complementary-Metal-Oxide-Semiconductor technology
combines both p-type and n-type devices on the same chip. Fig. 1.6(a)
shows the geometry of a CMOS pair of transistors where the main
substrate is p-type and this is used for the n-channel transistors. A
lightly doped n-type region, known as an ‘n-well’ is then diffused
sufficiently deeply into the p-type substrate so as to act as a
secondary substrate for the p-channel devices. This is known as n-well
CMOS technology and its complement p-well technology also exists.
Another well-established technology to emerge is ‘twin-tub’
CMOS technology, shown in Fig. 1.6(b), which uses both p-wells and nwells diffused into a main substrate that may be of either p-type or ntype material.
10
n-channel MOSFET
B
G
S
p-channel MOSFET
D
D
n+
n+
G
p+
S
B
p+
n--well
p- substrate
Fig. 1.6(a)
A CMOS Transistor Pair using n-Well Technology
n-channel MOSFET
B
S
G
p-channel MOSFET
D
D
p+
n+
n+
G
S
B
p+
n-well
p-well
substrate
Fig. 1.6(b)
A CMOS Transistor Pair using Twin-Tub Technology
11
Download