You are on page 1of 13

Chapter 4.

Electric Power Generation and Transmission Systems


(H Thng Pht v Truyn Ti in)

The electrical power system is a network of interconnected components which generate


electricity by converting different forms of energy, (thermal, hydro, and nuclear are the
most common forms of energy converted) to electrical energy; and transmit the electrical
energy to load centers to be used by the consumer. Electricity is generated at the
generating station by converting a primary source of energy to electrical energy. The
output of the generators is then stepped-up to appropriate transmission levels using a
step-up transformer. The transmission subsystem then transmits the power close to the
load centers.

4.1 Electric Power Generation


Almost all energy conversion processes take advantage of the synchronous AC
generator coupled to a steam, gas, or hydro turbine such that the turbine converts steam,
gas, or water flow into rotational energy, and the synchronous generator then converts the
rotational energy of the turbine into electrical energy. It is the turbine-generator
conversion process that is by far most economical and consequently most common in the
industry today. In this section, we will briefly look at this conversion process with
particular emphasis on the synchronous machine and the controls that are used to govern
its behavior. A basic turbine-generator can be illustrated in Figure 4.1.

Torque at
Steam speed Power at voltage Vt
Generator
Turbine

_ _
Governor Excitation
System
Vt
+ +
ref Vref

Figure 4.1 Block Diagram for Turbine-Generator System

The governor and excitation systems are known as feedback control systems because it is
the feedback loops which provide for good control of certain parameters. The governor
and excitation systems are typical feedback controllers in that the quantities to be
controlled (speed and voltage, respectively) are also providing the feedback signal.
The generator is classified as a synchronous machine because it is only at synchronous
speed that it can develop electromagnetic torque. If the nominal system frequency is f ,
synchronous speed is computed as
m
2
e
p
where e 2f is the frequency in rad/sec and p is the number of poles on the rotor of
120
the machine. The machine speed in RPM can be computed as N s f.
p

The synchronous generator has two iron structures. The rotor is the revolving part of the
machine, and is located inside the stator, which is the stationary part of the machine.
Because hydro-turbines are relatively slow, the number of poles must be high in order to
produce 50 Hz voltages (60 Hz in North America). Salient pole construction is simpler
and more economical when a large number of poles is required. Steam plants, on the
other hand, operate at high speeds (1500 and 3000 RPM steam-turbine-generators are
typical), and saliency would create significant mechanical stress at these speeds.
Therefore, smooth or round rotor construction is employed for these generators. The two
types of rotor construction are illustrated in Figure 4.2.

(a) (b)

Figure 4.2 (a) Salient Pole and (b) Round Rotor Construction

A magnetic field is produced by the DC-current carrying field winding on the rotor,
which induces the desired AC voltage in the armature winding. The field winding is
always located on the rotor where it is connected to an external DC source via slip rings
and brushes or to a revolving DC source via a special brushless configuration. The
armature consists of three windings, all of which are wound on the stator, physically
displaced from each other by 120 degrees. It is through these windings that the electrical
energy is produced and distributed. A typical layout for a 2 pole, round rotor machine
would appear as in Figure 4.3.
A

C B

B C

Figure 4.3 Winding Layout for Two-Pole Round Rotor Synchronous Machine

The DC current in the revolving field windings on the rotor produces a revolving
magnetic field. We denote the flux associated with this field that links the armature
windings as f (the subscript f indicates field windings). By Faradays Law of
Induction, this rotating magnetic field will induce voltages in the three armature
windings. Because these three windings are physically displaced by 120 degrees (for a
two-pole machine), the induced voltages will be phase displaced in time by 120 degrees.

If each of the three armature windings is connected across equal impedances, balanced
three phase currents will flow in them. These currents will in turn produce their own
magnetic fields. We denote the flux associated with each field as a , b , and c . The
resultant field with associated flux obtained as the sum of the three component fluxes a ,
b , and c is the field of armature reaction. We designate the associated flux as ar .
Using electromagnetic field theory and a trigonometric identity, one can show that ar
revolves at the same velocity as the rotor. Therefore the two fields represented by f
and ar are stationary with respect to each other. The armature field is effectively
locked in with the rotor field and the two fields are said to be rotating in synchronism.
The total resultant field is the sum of the field from the rotor windings and that associated
with armature reaction: r ar f .
d r
A voltage is induced in each of the three armature windings according to v N
dt
where is the number of winding turns. Because r is a sinusoidal function of time, the
negative sign captures the fact that the induced voltage will lag the flux by 90 degrees.
Letting E r , E ar , and E f be the voltages induced in winding a by the fluxes r , ar , and
f , respectively, we can represent the relationships in time among the various quantities
using the phasor diagram, illustrated in Figure 4.4.
ar
f

r Ef

Ear

Er
Ia

Figure 4.4 Phasor Diagram for Synchronous Generator

Regarding Figure 4.4, take note that


All voltages lag their corresponding fluxes by 90 degrees.
The current in winding a, denoted by I a , is in phase with the flux it produces ar
If I a 0 (no load conditions), then ar 0 , and in this case, r F , and E r E F .
All resistances have been neglected.

A Generator Equivalent Circuit can be illustrated in Figure 4.5.

jXs
Ia
jX ar + jX l +
Ef Er Vt Load
_ _

Figure 4.5 Equivalent Circuit Model of Synchronous Generator

Defining X s X l X ar as the synchronous reactance, we have that


Vt E f jX s I a

The phasor diagram corresponding to the equivalent circuit, when the load is inductive, is
shown in Figure 4.6.
Ef

-jXSIa


Vt
Ia

Figure 4.6 Phasor Diagram for Equivalent Circuit Inductive Load

The phasor diagram corresponding to the equivalent circuit, when the load is capacitive,
is shown in Figure 4.7.

Ia
Ef
-jXSIa


Vt
Figure 4.7 Phasor Diagram for Equivalent Circuit Capacitive Load

When the load is inductive, the current I a lags the voltage Vt ; the generator is said to be
operating lagging. When the load is capacitive, the current I a leads the voltage Vt ; the
generator is said to be operating leading. The angle between I a and Vt is i, i.e.,
I a | I a | I if Vt | Vt | 0. This implies that
i 0 lagging, i 0 leading
The excitation voltage magnitude | E f | is higher in the lagging case. We sometimes refer
to the lagging case as overexcited operation; here we have that | E f | cos | Vt | , where
is the angle between E f and Vt , it is also referred to as the load angle or power angle.
The leading case results in under-excited operation; in this case we have E f cos Vt .

Power Relationships:
From our equivalent circuit in Figure 4.5, we write that Vt E f jX s I a . Solving for I a
yields

E f Vt
Ia
jX s

If is the angle at which the excitation voltage leads the terminal voltage. Therefore,
| E f | Vt 0 | E f | cos j | E f | sin Vt | E f | cos Vt j | E f | sin
Ia
jX s jX s jX s jX s

| E f | sin | E f | cos Vt
Ia j
Xs Xs

E f sin
Equating real and imaginary parts of the above equation, we have I a cos
Xs
E f cos Vt
and I a sin . Multiplying both sides of the previous equations by 3Vt
Xs
yields
3Vt E f sin
Pout 3Vt I a cos
Xs

3Vt E f cos 3V
2

Qout 3Vt I a sin t


Xs Xs

We can note that, reactive power out of the machine is positive when the machine is
operated overexcited, i.e., when it is lagging implying 0 . The above power
relationships are based on the assumption that stator winding resistance is zero.

The electrical power output Pout can be plotted against the power angle , resulting in
sinusoidal variation as shown in Figure 4.8.

Pout
Pmax


0 90o 180o
Figure 4.8 Power Angle Curve

For simplicity, and without loss of generality, we neglect all real power losses associated
with windage and heat loss in the turbine and friction in turbine and generator bearings.
Continuing with the assumption that stator winding resistances are zero, in steady-state
operation, the mechanical power input to the machine is equal to the electrical power:
Pmech Pout . (In reality, Ploss 0 in steady-state operation so that Pmech Pout Ploss ).
Consider what happens to this lossless machine operating at Pout Pmax ( 90 ) when
the steam valve opening is increased so that Pmech becomes slightly larger. In this case,
the power angle increases beyond 90 , and the electrical power begins to decrease.
However, the mechanical power is only dependent on the steam valve opening, i.e., it is
unaffected by the decrease in Pout . This can only mean that Pmech Pout . The difference
Pmech Pout causes the machine to accelerate beyond its synchronous speed. When this
happens, we say that the machine has pulled out, gone out of step, or lost
synchronism. The generation level at which this happens is called the pull out power. It
is given by
3Vt E f
Pmax
Xs

This limit is lower when the generator is under-excited (leading current) because E f is
lower. Note that the excitation control system is used on synchronous machines to
regulate terminal voltage, and the turbine-governor system is used to regulate the speed
of the machine.

4.2 Electric Power Transmission


Electric power transmission, a process in the delivery of electricity to consumers, is the
bulk transfer of electrical power. Typically, power transmission is between the power
plant and a substation near a populated area. Electricity distribution is the delivery from
the substation to the consumers. Electric power transmission allows distant energy
sources (such as hydroelectric power plants) to be connected to consumers in population
centers. We may think of the transmission system as providing the medium of
transportation for electric energy, but one must realize that this transportation system is
unlike most in that the transportation takes place almost instantaneously. In addition, the
transmission system is a highly integrated system; that is, a change in the status of any
one component can significantly affect the operation of the entire system.

Due to the large amount of power involved, and in order to enable long distance power
transfer at lower current levels and therefore minimize | I 2 R | losses, transmission
normally takes place at high voltage (110 kV or above). Electricity is usually transmitted
over long distance through overhead power transmission lines. Underground cables are
used predominantly in densely populated areas where right-of-way costs for overhead
lines are prohibitively high, and they are also often used for transmission under rivers,
lakes, and bays. A power transmission system is sometimes referred to colloquially as a
"grid"; however, for reasons of economy, the network is not a mathematical grid.
Redundant paths and lines are provided so that power can be routed from any power plant
to any load center, through a variety of routes, based on the economics of the
transmission path and the cost of power. In this section, we will focus on transmission
lines; substation equipment such as transformers, relays, and circuit breakers, will not be
discussed.
4.2.1 Transmission Line Components

The basic components of a transmission line are the supports (towers), insulators, and the
conductors. Operation of a transmission line is also dependent on fault detection
equipment, voltage control equipment, and the bus arrangement at the terminals. In this
chapter, we will focus on conductor characteristics.

A single transmission circuit is comprised of three phases. Each phase may consist of a
single conductor, or each phase may be bundled in that it consists of two or more
conductors suspended from the same insulator string. The bundled design is usually used
at the high voltage levels, because it minimizes power loss and radio interference due to
corona, which occurs when the surface voltage gradient of a conductor gets so high, that
it causes partial dielectric breakdown in the surrounding air, and ionization occurs.
Conductors are always bare in order to allow maximum heat dissipation. In addition to
the phase conductors, one or two grounded shield wires are also strung along the top of
the tower in order to protect the phase conductors from lightning strokes. Shield wires on
transmission lines may include optical fibers (OPGW Optical Ground Wire), used for
communication and control of the power system. Today, almost all conductors utilize
aluminum in their construction because aluminum is relatively inexpensive, and
lightweight. The most common types of conductors are commonly referred to by the
following acronyms: AAC (all-aluminum conductor), AAAC (all-aluminum-alloy
conductor), and ACSR (aluminum conductor, steel-reinforced).

There are three main parameters of a conductor that are of concern to us. These are the
resistance, inductance, and the shunt capacitance. We will discuss some of the
influencing effects regarding these characteristics. Because we consider only balanced
three-phase operation, discussion is limited to positive sequence quantities only.

Conductor Resistance:
Conductors of any material have resistance. Although conductor resistance is small
enough so that it does not appreciably contribute to voltage drop, it is of considerable
interest in systems analysis because it causes | I | 2 R losses. The conductor resistance to
direct current is given by R l / A , where is the resistivity, | l | is the length, and A
is the cross-sectional area of the conductor. However, there are two other important
considerations regarding the DC resistance: (1) Values of resistivity are given for a
specified temperature (e.g. 20 degrees C), and resistivity increases approximately linearly
with temperature; (2) Because conductors are actually made with strands of material that
are spiraled around a central core, the length used to compute resistance should be greater
than the length of the conductor itself.

Due to the skin effect, resistance to AC is usually higher than the resistance to DC
because AC causes current distribution in the conductor to be non-uniform; typically,
more current tends to flow at the surface of the conductor than in the interior.

Conductor Inductance:
Current flowing in a single conductor generates a magnetic field surrounding the
conductor. We may denote the alternating current as i (t ) because it varies with time;
consequently, so will the magnetic field. We characterize this magnetic field with
magnetic flux (t ) in Webers. The amount of flux which links this conductor is the flux
linkage (t ) N (t ) in Weber-turns, where N is the number of turns linked. In the case
of a single conductor, N 1 . We assume here that (t ) is given on a per unit length
basis. According to Faradays Law, the magnitude of the induced voltage will be
d (t )
vL (t )
dt

in units of induced volts per unit length of conductor. If the magnetic field is set up in a
medium of constant permeability, then
(t ) Li(t )
where L is a constant. Substitute (t ) into equation of the induced voltage, we have
di (t )
v L (t ) L
dt
The constant L is defined as the inductance of the conductor, and it relates the voltage
induced by the changing magnetic field to the rate of change of current:
v (t )
L L
di (t ) / dt
Here, L is given in units of henries per unit length. If we consider sinusoidal steady-
state quantities, we have
VL jLI
where the quantity X L L is defined as the inductive reactance per unit length of the
conductor, and V L represents the voltage drop per unit length across the conductor
carrying current I. For a power transmission line, the inductive reactance is higher than
the resistance by a factor of about 2 to 3 for lower voltage transmission lines, increasing
to a factor of about 20 to 30 for the highest voltage transmission line. Consequently,
inductive reactance is the dominant factor in computing voltage drop and power flow
across a line.

Conductor Shunt Capacitance:


Let us consider again a single conductor such that the return path is located far away from
this conductor, that there is a charge on the peripheral of the conductor that is uniformly
distributed throughout its length, and that an equal and opposite charge is distributed
along the earth below the conductor. This charge generates an electric field emanating
from the conductor directed towards the ground. If the conductor voltage is alternating,
then we may denote it as v(t ) because it varies with time; consequently, so will the
electric field. We characterize the electric field with the charge q (t ) in coulombs per unit
length of conductor. The flow of charge per unit length caused by the changing voltage
is current, given by
dq(t )
iC (t )
dt
If the electric field is set up in a medium of constant permittivity, then q(t ) Cv(t ) ,
where C is a constant. Substitution of q (t ) into iC(t) yields

dv(t )
iC (t ) C
dt
The constant C is defined as the capacitance per unit length of the conductor, and it
relates the current resulting from the changing electric field to the rate of change of
voltage:

iC (t )
C
dv(t ) / dt

If we consider sinusoidal steady-state quantities, then we can write I C jCV , where


the quantity BC C is the capacitive susceptance per unit length of the conductor to
ground. For three phase transmission lines, the capacitance between the phases is
normally much larger than the capacitance to ground.

Note the j in the above equation causes I C to lead V by 90 degrees. Therefore line
capacitance, also called line charging, contributes leading current. In a power system,
when the loads are heavy, currents of high magnitude in the lines result in heavy reactive
losses ( | I | 2 X L ). The current in this case is lagging (since I (V S V R ) / jX L ), and the
effect of the leading current from the line capacitance is to bring the total current angle
closer to zero degrees. This effect is desirable since the same real power flow can be
obtained for a smaller current magnitude: P 3 Re(VI * ) 3V | I | cos , where is the
angle between the current and voltage phasors; if decreases, then | I | can decrease for
constant P , assuming that V remains approximately constant. Reducing current
magnitude will decrease both real losses( | I | 2 R ) and reactive losses ( | I | 2 X L ). Another
equally valid way to think about this is that the line inductance absorbs reactive power (-
MVAr) whereas the line capacitance produces reactive power (+MVAr).

There is also a real part, the conductance, caused mainly by insulator leakage. However,
this component is usually very small and negligible. It is rarely modeled in power system
studies.

Lumped-Circuit Equivalent:
The pi-equivalent lumped parameter model is used for most power flow analysis
applications. This model represents the distributed effects of the series resistance and
inductance and the shunt capacitance with composite or lumped values. Figure 4.9
illustrates the model.
R jXL
p q

Y/2 Y/2

Figure 4.9 Pi-Equivalent Model of a Transmission Line

For transmission lines less than about 200 km in length, the lumped parameters are
computed as simply the product of the per-unit length parameter and the line length. For
lines that exceed 200 km, the model is still used but one needs to more rigorously derive
appropriate expressions based on equivalent terminal characteristics.

4.2.2 Complex Power Transmission

In this section, we will develop equations for computing real and reactive power flow in a
transmission line. We consider a transmission line that interconnects two buses p and q
in the fig. 4.9. Therefore we may denote the series impedance as Z pq R jX L and the
shunt admittance at each end of the line as Y pp Yqq Y / 2 where Y is the total line
charging given by Y jBC . In addition, we may represent the series impedance as
1 R jX L
admittance; thus we have Y pq G jB ( Y pq G jB 2 ). We will
R jX L R X L 2
investigate the power flow into the transmission line from bus p in terms of two
components: the flow into the p-side shunt capacitance, and the flow into the p to q series
impedance. The total flow into the line from the p bus will then be the sum of these two
components. Similar analysis will apply for the flow from bus q into the transmission
line (which is just the negative of the flow from the transmission line into the q bus).

Using phasor notation, we denote the per unit voltages at the p and q buses as V p p and
Vq q , respectively.

Flow Into Charging Capacitance:


Let the current flowing into the p-side charging capacitance be | I Sp | . This current is
given by
I Sp V p p (Y pp ) V p p ( jBC / 2) (4.1)

The per phase complex power flowing into the p-side charging capacitance is then given
by
S Sp V p p ( I Sp ) * V p p [V p p ( jBC / 2)]* jV p ( BC / 2)
2
(4.2)
where the asterisk indicates complex conjugation. Therefore the power flow into the
charging capacitance is purely reactive, and the negative sign indicates that vars are being
supplied to the network, not absorbed from it.

Flow Into Series Impedance:


The per phase complex power flowing into the series impedance from the p bus is given
by

S pq V p p ( I pq ) * (4.3)
where | I pq | is the current flowing into the series impedance, computed as

I pq [V p p V q q ]Y pq [V p p V q q ](G jB) (4.4)

However, from eqn. 4.3, we see that | I pq | must be conjugated. Recall that if x ab
where a and b are two complex numbers, then x * a * b * . Applying this relation to
equation 4.4, we have that

I pq [V p p V q q ]* (G jB) * [V p p V q q ](G jB)


*

Substitution into eqn. 4.3 yields

S pq V p p [V p p V q q ](G jB) [V p V pV q ( p q )](G jB)


2

Recalling that V pV q ( p q ) V pV q cos( p q ) jV pV q sin( p q ) , we may rewrite


the last expression as

S pq [V p V pV q cos( p q ) jV pV q sin( p q )](G jB)


2

Denoting p q p q , carrying out the indicated multiplication, and then collecting


real and imaginary parts, we have that

S pq V p G V pV q G cos p q V pV q B sin p q
2

j[V p B V pV q B cos p q V pV q G sin p q ]


2
(4.5)

The real and reactive power flow from bus p to bus q, measured at bus p, are then given
by
Ppq V p G V pV q G cos p q V pV q B sin p q
2
(4.6)
Q pq V p B V pV q B cos p q V pV q G sin p q
2
(4.7)

These equations are appropriate for computing real and reactive power flow across a
transmission line. If voltages are given as phase to neutral, in volts, and impedances as
per phase, in ohms, then the computed power quantities are per phase, in watts and vars.
These equations can also be used if voltages and impedance values are given in per unit;
in this case, the computed power quantities are also per unit, and multiplication by the
system 3-phase base gives the three phase power flowing across the transmission line.

Normally, R X L so that we can assume R=0, and this implies G=0; then eqns. 4.6
and 4.7 become

V pV q
Ppq sin p q (4.8)
XL
V p V pVq cos p q V p (V p Vq cos p q )
2

Q pq (4.9)
XL XL

These equations are very frequently used to analyse how certain flows are affected by
design or operational actions that might be under consideration.

To avoid system stability problems, the angular difference p q p q across a


transmission line is rarely allowed to exceed about 40 degrees; typically, angle
differences are less than 20 to 30 degrees. This means that the arguments of the
trigonometric functions in eqns. 4.8 and 4.9 are small angles. For small angles (given in
radians) these trigonometric functions may be approximated with sin p q p q and
cos p q 1.0 , causing eqns. 4.8 and 4.9 to simplify to

V pV q p q
Ppq
XL
V p (V P V q )
Q pq
XL
From these two equations, we see that real and reactive flows are heavily influenced by
the the series reactance. Real power flow is closely related to differences between bus
angles; this is especially apparent if we realize that normally, bus voltage magnitudes do
not deviate substantially from 1.0 per unit. Reactive power flow is closely related to
differences in bus voltage magnitudes. These concepts are fundamental to understanding
control of real and reactive power flow in a transmission system.

You might also like