Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

TOPIC 3: LOAD-FLOW ANALYSIS IN POWER SYSTEMS

TOPIC OUTLINE
3.1. Meaning and Objectives of Load-Flow Study
3.2. System Specification for Load-Flow Analysis
3.3. Mathematical Formulation of Load-Flow Problem
3.4. Numerical Solutions of the Load-Flow Problem
3.4.1. Gauss-Seidel (G-S) load-flow solution method
3.4.2. Newton-Raphson (N-R) load-flow solution method
3.4.3. Pθ-QV Decoupling method
3.4.4. DC Load-flow model
3.5. Methods of Load-Flow Control/Regulation

3.1. MEANING & OBJECTIVES OF LOAD-FLOW ANALYSIS


A load flow analysis is power system term for the steady-state performance analysis of a network. That
is, the calculation or assessment of how a given system responds to certain load-demand and generation
schedules under steady-state conditions. Steady-state system response is normally measured by its
busbar voltages (both magnitude and phase angles) and the magnitudes of active and reactive power
flow in its branches. This system response analysis does not essentially differ from the solution of
electric circuits outlined in Topic #2, except that certain constraints are peculiar to power supply
systems. For example, successful power supply system operation under normal balanced three-phase
steady state conditions requires (i.e., key assumptions in load-flow analysis/studies):
(i) The individual generator outputs must be maintained at predetermined set points to efficiently
match the total consumer demand placed on the system and the total system losses (i.e.,
adequacy requirement/constraint).
(ii) The busbar voltage magnitudes and phase must remain close to rated values (i.e., voltage quality
& stability constraints).
(iii) The generators must be operated within specified real and reactive power limits (i.e., generator
safety constraints).
(iv) The transmission lines, transformers, and other system network components must not be
overloaded or be operated too close to their stability or thermal limits (i.e., network safety or
thermal overload constraints).
Load flow study is performed to investigate the above requirements or assumptions by computing the
voltage magnitude and phase angle at each busbar in the system, and the flow of real and reactive power
in the branches of the network. As a by-product of this calculation, network power losses can be
computed and/or optimised.
Load flow studies are normally performed on both existing power systems and on power systems
incorporating new generation and transmission facilities to meet projected load growth for various
operating conditions, which include both normal and contingency conditions. Power systems operating
under outage or contingency conditions (i.e., when one or several elements normally in service is lost)
are also of interest, especially the effect of such outages on the loading of the system (i.e., the power
flow in the elements remaining in service and bus voltages).

EE 307_Topic #3 (Sem. I, 2014/2015) 1 of 20


The results (outputs) of load-flow studies are used in a number of areas, both during long term planning
and short term operation:
(a) Assessment of alternating plans of generation and/or network expansion to meeting projected load
growth (incl. change of conductor size and/or change of level of service voltage)
(b) Management of contingency situations resulting from loss of a large generating unit or a major
transmission line outage due to thermal overloading of the line (i.e., managing the temporary loss
of generation and network components)
(c) Determining the optimal size and favourable location/site for capacitor banks for improving the
power factor as well as the voltage profile of a line
(d) Determining the reactive power compensation needed to maintain bus voltages at some desired
levels
(e) Establishing the incremental line losses associated with changing the output of a generator that is
useful in deciding optimal generation allocations to the generating stations so that the cost of
generation is minimal
(f) Management the incremental loading effect associated with rearrangement or reconfiguration of
power networks and/or incorporating/adding new network elements, and/or injection of in-phase
or quadrature boost voltages (incl. optimising system losses)
If unfavourable conditions are detected from a load-flow analysis (i.e., any one of the constraints of
adequacy, quality, stability, and safety is violated), some corrective (preventive) measures (actions)
should be taken (initiated) to mitigate the potential problems, especially by changing the operating
condition of the system. This can be effected by changing where real and reactive powers are generated
and altering their paths from the point of generation to the utilisation point using any of these methods
or their combination: synchronous generators, capacitor banks, and transformers with tap settings. These
and many other methods of system load-flow control are discussed in Topic #5.

3.2. SYSTEM SPECIFICATION FOR LOAD-FLOW STUDY


The starting point for a load-flow study problem is a schematic (a single-line) diagram of the power
system, from which the input data for its solution (performance analysis) can be obtained. Input data
consist of bus data (types and parameters) and branch data (i.e., transmission line and transformer
data).
The data associated with each bus (k) are the active and reactive load demands (Pd,k, Qd,k), generation
power outputs (Pg,k, Qg,k), shunt capacitor output (QC), and voltage magnitude (Vk) and relative phase
angle (θk). The net injected power (Pk, Qk) into any bus “k” is given as a difference between power
generated at the bus and power of any loads at that bus. Of all these bus data, only the load demands and
generation outputs are specified as input data at each bus, and the other two are unknowns to be
computed.

The input data for each transmission line includes the per-unit equivalent π circuit series impedance and
shunt admittance, the two buses to which the line is connected and maximum MVA ratings. Similarly,
input data for each transformer include per-unit winding impedances, the per-unit exciting branch
admittance, the buses to which the windings are connected, and maximum MVA ratings.

EE 307_Topic #3 (Sem. I, 2014/2015) 2 of 20


Branch Buses Branch Equiv. Circuit Parameters Tap
Start Bus End Bus Series Resistance Series Reactance Shunt Admittance Ratio*

For the purposes of load-flow analysis and depending on the system performance variable specified
apriori, power system buses can be categorised into the following types:
• swing (slack, or floating) bus;
• P-Q bus; and
• P-V bus.
Table 3.1 presents a summary of the characteristics or parameters normally specified for these principal
bus types.
Table 3.1. Summary of principal bus types
Bus Bar Electrical Quantities
Voltage Generation Demanded Power Shunt Capacitor
Bus Type Power Power
V δ Pg Qg Pd Qd Qc
Load (P-Q) Bus x x n.a. n.a. √ √ √
Generator (P-V) Bus √ x √ x n.a. n.a. n.a.
Slack Bus √ √ x x n.a. n.a. n.a.
The detailed nature of the above types of buses is outline below.
(a) Swing (aka slack or reference) bus: This is typically as a generator bus and is only one in a
power system. For this type of bus, we know the voltage magnitude and phase angle. Physically,
there is nothing special about the swing bus; in fact, it is a mathematical artefact of the solution
procedure. The generation must supply both the load and the losses in the network. Before solving
the load flow problem, we will know all injections at P-Q buses, but we will not know what losses
will be as losses are function of the flows which are yet to be computed. So we may set the real
power injections for, at most, all but one of the generators. The one generator for which we do not
set the real power injection is the one modelled as the swing bus. Thus, this generator “swings” to
compensate for the network losses (i.e., compensate for the real power mismatches of the entire
system), or, one may say that it “takes up the slack.” The voltage magnitude of the swing bus is
chosen to correspond to the typical voltage setting of this generator, while the voltage angle may
be designated to be any angle, but normally it is set equal to zero.
(b) P-V Buses: For this type of buses, we know the net real power injection (i.e., Pg – Pd) and the
voltage magnitude. These buses fall under the category of voltage-controlled buses because of the
ability to specify the voltage magnitude of this bus. Most generator buses (i.e., buses at which
generators are connected) fall into this category, independent of whether it also has load;
exceptions are buses that have reactive power injection at either the generator’s upper limit or its
lower limit. P-V buses where generators operate at their limiting reactive power values are treated
as P-Q buses.
(c) P-Q Buses: For this type of buses, we know the net real and reactive power injections, but not the
voltage magnitude or angle. All load buses fall under this category, including buses that have not
either load or generation (i.e., null-injection buses).
Once the power system whose steady-state performance analysis is to be conducted is specified as
detailed above, the next step needed is formulation of a mathematical model that describe the

EE 307_Topic #3 (Sem. I, 2014/2015) 3 of 20


relationship between voltages and powers in the system. For complete solution of the system, the
formulated mathematical model must be solved for the values of voltage magnitude and phase angle of
each bus under balanced three-phase steady-state conditions. As an offshoot of this calculation, the real
and reactive power flows in transmission lines, transformers, as well as the equipment and system losses
can be computed. These two sub-problems of the load-flow study problem are discussed the following
sections.

3.3. MATHEMATICAL FORMULATION OF THE LOAD-FLOW PROBLEM


Having established the basic principles in load-flow analysis, let us now formulate the equation(s),
which facilitates this kind of analysis. For this purpose, we will consider a simple power system
consisting of two busbars connected together by a single line (Fig. 3.1).

G1 G2

Sg1=Pg1+jQg1 Sg2=Pg2+jQg2
2
1 V1ejδ1 V2ejδ2

Transm. Line
Sd1=Pd1+jQd1 Sd2=Pd2+jQd2

I1 1 xs rs 2 I2
zs=rs+jxs

yp S2
S1 yp

Figure 3.1. Single-line diagram and equivalent circuit of a two-bus system


The two busbars are denoted as “1” and “2”, respectively. If busbar “1” is assumed to be a net
generation (or source) busbar and busbar “2” – a net load (sink) busbar, then the complex power flow in
the interconnecting line (S12) is equal to the net complex power injected at busbar “1”, which is given by
the expression:

S1 = S12 = P12 + jQ12 = V1I12* (3.1)

or

S1 = S12 = P12 − jQ12 = V1* I12 (3.2)

where V1 = |V1|∠θ1 and I12 = current flow in the transmission line or the net current injected at busbar
“1” from some source. The expression for the line or injected current can be obtained by application of
the bus admittance equation to the appropriate bus. For our case, the application of this equation at bus
“1” gives the following expression
2
I12 = I1 = ∑ Y1iVi = Y11V1 + Y12V2 (3.3)
i =1

EE 307_Topic #3 (Sem. I, 2014/2015) 4 of 20


where Y11 is the self-admittance of busbar “1” and Y12 is the mutual admittance between busbar “1” and
bus “2”.
By substituting Eq. (3.3) into Eq. (3.1), we obtain

S1* P1 − jQ1
= = Y11V1 + Y12V2 (3.4)
V1* V1*
or

S1* = P1 − jQ1 = V1* (Y11V1 + Y12V2 ) (3.5)

By equating the real and the imaginary parts in the equation (3.4) or (3.5), we obtain

P1 = Re[V1* (Y11V1 + Y12V2 )] (3.6)

Q1 = − Im[V1* (Y11V1 + Y12V2 )] (3.7)

The power flow expression given by any of the equations (3.5 – 3.7) constitutes the mathematical
model/statement/formulation of the load-flow problem for the above single-circuit two-bus system.
Using the same procedure outlined above, we can obtain similar models for a generalised bus “k” of an
“N” bus system. That is,

PK − jQK N N

VK*
= ∑
i =1
YKiVi or PK − jQK = VK* ∑ YKiVi
i =1
(3.8)

Consequently,

 * N 
Pk = Re Vk ∑ YkiVi  (3.9)
 i =1 
 N 
Qk = − ImVk* ∑ YkiVi  (3.10)
 i =1 
If we denote Vk = Vk e jδ k , Vi = Vi e jδ i , and Yk ,i = Yk ,i e jδ ki , then:

N
Pk = V k ∑Y
i =1
ki Vi cos(θ ki + δ i − δ k )
N (3.11)
Qk = − Vk ∑ Yki Vi sin(θ ki + δ i − δ k )
i =1

The equations [3.8], [3.9], [3.10] and [3.11] are the various versions of the mathematical formulations
(models) of a load-flow problem. They are often referred to as the static load-flow equation (SLFE).
These equations are non-linear (because of the non-linear relation between current and voltage
variables) and therefore explicit or direct solution by inversion of the bus admittance matrix is not
possible. Solution can only be obtained by iterative numerical techniques. Some of the commonly used
numerical solution methods are discussed in the next section.
Before going to the solution of the formulated power flow problem, it is important to point out some of
the most important characteristics of these equations:

EE 307_Topic #3 (Sem. I, 2014/2015) 5 of 20


(i) The load flow equations are algebraic because they represent a static system or a system
operating in steady state
(ii) The equations are non-linear because of they have terms containing products of some of the
unknowns and also terms containing trigonometric functions of some of the unknowns. These
nonlinearities will generally make it impossible to obtain analytical solutions. However, we shall
be content to obtain numerical solutions.
(iii) The load flow equations satisfy the power balance requirements. That is, the sum of the net power
injections at all the buses is equal to the total power losses in the system. The latter is given by the
sum of the expressions on the RHS of the load flow equations. This fact holds for both real and
reactive power balances, although for the latter, the actual total reactive power loss is less than the
value calculated from the power balance by the reactive power generated in the line shunt
capacitance.
(iv) Because the RHS of the load flow equations are functions of the voltage variables only, the same
will apply to the system power losses.
(v) In the load flow equations, the voltage phase angles appear in difference form. This characteristic
is often employed to derive linear approximations or models of the load flow equations (as will be
seen later).
For the above SLFE solution to have practical significance, all the state and control variables must be
within specified practical safety (load-carrying capability), quality, and stability limits. State or
dependent variables are those affected by other system variables (e.g., busbar voltage magnitude and
phase angle), while control or independent variables are those, which are easily controlled by the
electric utility (e.g., generated real and reactive powers). System variables, which are determined by the
consumers and hence beyond the control of the electric utility are called disturbance or uncontrollable
variables. Examples of disturbance variables include real and reactive load demand. The magnitude of
the busbar voltage is affected by the generated reactive power, while the phase angle of the busbar
voltage – by the generated real power. A typical SLFE will, thus, have six variables.
The problem of the load-flow analysis can, therefore, be descriptively formulated as follows: “Solve for
the state variables in the SLFE, taking as given the control variables for a system of known bus
admittance matrix with known operating safety, quality and stability limits on generators and regulating
transformers at a specified load condition. Additionally, solve for the real and reactive power flows in
all the branches or lines.”

3.4. THE LOAD-FLOW SOLUTION METHODS


As observed above, the formulated system of load-flow equations are non-linear and direct solutions
using matrix inversion and Gauss elimination methods is not possible. Special techniques are, therefore,
required for their solution. A large number of methods for solving systems of non-linear algebraic or
transcendental equations have been developed. Most of these methods are iterative in nature. That is, we
guess an approximate value of the solution and substitute it into some formula involving the equation to
obtain the next approximation, which is hopefully closer to the solution. This approximation is in turn
used to obtain a third one and so on. The iteration process is stopped when we think it has converged to
the solution with sufficient accuracy (i.e., criterion for acceptance of a solution).
In this section, we will discuss two of these methods, which have gained wide adoption in power
system-related problems. These are the Gauss-Seidel method and Newton-Raphson method. The former
has been widely used for many years and is simple in approach. The latter is, however, more complex
but has certain advantages. We will also look at the following approximate methods: decoupling method
and DC load flow method.

EE 307_Topic #3 (Sem. I, 2014/2015) 6 of 20


3.4.1. THE GAUSS-SEIDEL (G-S) ITERATIVE TCHNIQUE

(a) The underlying mathematical theory of the technique


The principal idea behind this method is that in computing the (m+1)-th approximation of the unknown
Xi for i > 1, the earlier computed (m+1)-th approximations of the unknowns X1, X2, ….., Xi – 1 are taken
into account (cf. Simple iteration). Thus in this method, as soon as a new value of a variable is available,
it is used in all the subsequent calculations.
To use this method to solve a system of equations

f i ( x1 , x2 , x3 ,...., xn ) = 0 (3.12)

we re-write it in the form

xi = ϕ i ( x1 , x2 , x3 ,........., xn ) (3.13)

By using the method of successive approximation to solve Equation (3.13), we get

(
xi( m +1) = ϕ i x1m +1 , x2m +1 , x3m +1 ,....., xim−1+1 , xim , xim+1 ,....., xnm ) (3.14)

where the superscripts (m+1) and (m) indicates the new (improved) and initial (old) iteration.
Calculation using Equation (3.14) is continued until two consecutive iterations of xi are within some
predefined tolerance, ε, of each other. That is,

xim +1 − xim ≤ ε (3.15)

Once the test in Equation (3.15) is satisfied, the iterative process is stopped. Note that the above
criterion for acceptance of solution can also be written as

ϕ ( xim ) − xim ≤ ε (3.16)

(b) Application of G-S iteration technique to power flow solution


Let us now consider the application of this method to the static load-flow equation. We already know
that the static load-flow equation can be written in the form
N
Si*

k =1
yi , kVk = * = yi1V1 + yi 2V2 + .... + yiiVi + ... + yinVn
Vi
(3.17)

From this equation,

1  Si* N 
Vi =  * − ∑ yikVk , for k ≠ i (3.18)
yii Vi k =1 
Equation (3.18) can now be solved using the method of successive approximation. Thus,

1  Si* i −1 N

 *( m ) ∑ ik k ∑
m +1 m +1
Vi = − y V − yikVkm  (3.19)
yii Vi k =1 k = i +1 
The superscript (m) denotes the iteration number “m” and superscript (m+1) denotes the subsequent
iteration. The equation for Vk for k < i would have been solved prior to bus i. Hence, the voltage for

EE 307_Topic #3 (Sem. I, 2014/2015) 7 of 20


iteration “m+1” is available. The Vk for k > i have not been determined, so Equation (3.19) must use the
Vk from the previous iteration, i.e., m-th iteration.
In order to terminate the iteration using equation [3.19], a test equation similar to that of equation [3.15]
or [3.16] is used. Thus

Vi ( m +1) − Vi ( m ) ≤ ε (3.20)

That is, the iteration process is continued until the difference between two consecutive estimates of the
voltage becomes less than or equal to the indicated (predetermined) tolerance or error of accuracy.
Convergence upon an erroneous solution may occur if the initial voltages are widely different from the
correct value. Such problems can be avoided if the initial estimates are of reasonable magnitude and do
not differ too widely in phase. Since the nominal operating voltages of a power system are usually
chosen as per unit bases, a good estimate of the magnitudes of the bus voltages is 1.0 p.u. (or nominal
rated voltage). The voltage phase angles throughout a system usually vary by a few degrees. By setting
the floating busbar voltage angle to zero, the other busbar voltage angles will probably be near zero also.
Therefore, zero degrees is a natural first approximation for all voltage angles.
Equation [3.19] refers to a busbar with P and Q specified. At a generator bus, however, voltage
magnitude and real power are specified with perhaps upper and lower limits to the reactive power.
The magnitude of the busbar voltage is fixed, but its phase depends on the reactive power at the busbar.
The values of complex voltage and reactive power from the previous iteration are not related by [3.19]
because complex voltage has been modified to give a constant value of magnitude. It is necessary at the
next iteration to calculate the value of reactive power corresponding to the complex voltage from
equation [3.19]. This value of reactive power holds for the existing value of the complex voltage and is
then substituted into equation [3.19] to obtain the new or corrected value of the voltage. Thus, for a PV
bus,

 Pi − jQi( m +1) i −1 N 
∑ ∑
( m +1) 1 ( m +1)
Vi =  *( m )
− y ik V k − y ik V k
( m)
 (3.21)
y ii  Vi k =1 k = i +1 

The changing value of the reactive power appearing in the above equation is determined using the
equation (3.10). That is:

 * N 
Qi = − Im Vi ∑YikVk 
 k =1 
"
For evaluating the value of this reactive power, the best estimate obtained so far for the voltage and
should be used.
The real and imaginary components of the new complex voltage are then multiplied by the ratio of the
fixed (regulated) voltage and the magnitude of the new voltage, thus complying with the constant
voltage magnitude constraint. The phase of the voltage is thus obtained and the iteration can proceed to
the next busbar.
If at any iteration the required reactive power to maintain the voltage at the PV bus constant is outside
the specified limits, equation (3.19) is introduced for that bus, using the limit as the specified reactive
power value. That is, that bus is converted into a PQ (load) bus.
For a swing busbar, both the voltage magnitude and phase are specified. Therefore, the equation
corresponding to this busbar can be eliminated from the bus equations. So, for an N busbar system, only

EE 307_Topic #3 (Sem. I, 2014/2015) 8 of 20


(N – 1) equations must be solved to determine all the system bus voltages. The net real and reactive
power being injected at the swing bus can be determined after the other system voltages have been
found. The sum of real and reactive powers leaving the swing bus on all system elements connected to
it, other than generators and loads, is the net injected power at the swing bus.
Note: Often, faster convergence of the G-S iteration within the specified tolerance limit can be achieved
with the use of an acceleration factor, α, at the end of each iteration. This accelerated G-S scheme is
given as:
= + − = + 1−
or

= ∗ − + 1−

The selection of the acceleration factor is based on conjecture and is often governed by experience. A
scalar acceleration factor in the range 0 < ≤ 2 will improve the convergence but cannot change a
divergent problem into one that converges. When α = 1, this becomes the basic G-S iteration. An 0<α<1
is called under-relaxation and 1<α<2 is over-relaxation. The acceleration factor chosen depends upon
the system and its value normally lies within 1.4 to 1.6.
After the power injected at the swing bus is calculated and voltage magnitudes and phases are known at
every bus in the system, the line power flows and line losses can then be easily be calculated. For
instance, the complex power flow from any bus k to m on the element connecting these two buses is:

Skm = Pkm + jQkm = Vk I km = Vk [(Vk − Vm ) ykm ]


* *
(3.22)

Similarly, the line flow from bus m to k is given by:

S mk = Pmk + jQmk = Vm I mk = Vm [(Vm − V k ) y mk ]


* *

The algebraic sum of the two equations for the line flows is the line losses in the element
interconnecting buses k and m.
Note that if we assume the π representation of transmission line, the above line flows will be expressed
as:
#&,

= − = ∗
= ∗
! − . #$, + . '
2
and
#&,

= − = ∗
= ∗
! − . #$, + . '
2
(c) The general algorithm for load-flow solution using G-S method
(i) Define the bus types and show the variables specified apriori and the unknown variables.
(ii) Assemble the bus admittance matrix using either the rule of self and mutual admittances, or using
the singular transformation technique.
(iii) Carry out iterative computation of unknown bus voltages (both in magnitude & phase angle)
according to the following sequence of steps:
Assume initial values of voltages for load buses and phase angles (except swing bus):

EE 307_Topic #3 (Sem. I, 2014/2015) 9 of 20


Determine first estimates of voltage at all load and generation buses in terms of the initially
assumed voltages using the standard iteration equation
Insert the obtained value of voltages again in the iteration equation in order to obtain the new
estimates of voltages. Continue this process for a number of iterations
Repeat the above steps for all the buses until the difference between the estimated voltage
and the previous estimated values lies within predetermined precision index (i.e., continue the
iteration until the changes in magnitude of bus voltages between two successive iterations is
less than a certain tolerance for all bus voltages).
(iv) Compute the swing or floating bus power injections (NB: real and reactive power injections at
the swing bus give total system losses).
(v) Compute the power flows and losses in the branches or lines, together with the total system
losses and efficiency.

Sample Problems on G-S Iteration Solution

Example #1. Perform one iteration of a G-S load flow solution of the system shown below. Buses #2
and #3 are load buses with per-unit loads of 0.25 + 0.12 *+ and 0.3 + 0.15 *+, respectively. Bus #1
is the swing bus with = 1.04. /0 *+. Perform the bus calculations in numerical order. The figures
shown on the diagram are per-unit impedances.

1 0.05+j0.3 2 3

G
0.1+j0.5
0.05+j0.3

-j12.5
-j25 -j12.5
-j25

Example #2. Perform one iteration of a G-S load flow solution of the system shown below. Buses #3
and #4 are load buses with per unit loads of 0.23 + 0.19 and 0.42 + 0.13, respectively. Bus #1 is the
swing bus with = 1.03. /0 *+. Bus #2 is a voltage controlled bus with | 3 | = 1.02 *+ and a
generated power of 0.24 pu. Perform the bus voltage calculations in numerical order. The figures shown
on the diagram are per-unit impedances.
1 4 j0.5 2
j0.2
j0.5

G G
3
j0.4 j0.4

Example #3: Using the G-S method, calculate the complex bus voltages after the first iteration for the
three-bus network shown below. Busbar #2 is a load bus consuming 200 MW, 120 MVAr, busbar #3 is
a generator bus set at 70 MW and 228 kV. Busbar #1 is a swing bus with a voltage magnitude of 230
kV. The three interconnecting lines have equivalent 4-circuit parameters as:

EE 307_Topic #3 (Sem. I, 2014/2015) 10 of 20


5 = 53 = 56 = 6.912 + 34.56 Ω
= 3 = 6 = 8.9:;9;<:.

G G

1 3

Line 1

Line 2 Line 3

3.4.2. THE NEWTON-RAPHSON (N-R) METHOD

(a) The underlying mathematical theory of the method


The mathematical basis of this method is Taylor series, which is used to obtain a system of linear
equations from the set of non-linear load-flow equations derived earlier. The Taylor expansion of a
multivariable function about a point gives the change in value of the function caused by changes in its
variables in terms of the function gradients at the chosen point and the variable increments, assuming
that the function can be differentiated. Expanding about the nominal value of a function y, the increment
∆y is given by
N
∂y 1 N N ∂2 y 1 N N N ∂3 y
∆y = ∑ ∆xi + ∑ ∑ ∆xi ∆x j + ∑∑∑ ∆xi ∆x j ∆xk + ..... (3.23)
i =1 ∂xi 2! i =1 j =1 ∂xi ∂x j 3! i =1 j =1 k =1 ∂xi ∂x j ∂xk

where ∆y = y – y0 — function increment, and ∆xi = x – x0 — variable increment, N! – ‘N’ - factorial


A linear form of the above function is obtained when all the terms involving derivatives of the second
and higher order are neglected. This gives the following approximate relationship:
N
∂y
∆y = ∑ ∆xi (3.24)
i =1 ∂xi

The value of the function at any point can then be expressed in terms of its variables and the nominal
values as:

∂y
(xi − x0,i )
N
y = y0 + ∑ (3.25)
i =1 ∂ x0,i

For an equation written in the general form y = C , where C is a constant, linearisation by the Taylor
series gives the equation

∂y
(xi − x0,i )
N
C = y0 + ∑ (3.26)
i =1 ∂x0,i
By the method of successive approximation, any value for the function parameter can be obtained from
its nominal value and the incremental value. That is,

xi( m +1) = xi( m ) + ∆xi( m ) (3.27)

EE 307_Topic #3 (Sem. I, 2014/2015) 11 of 20


The incremental value of the variable is obtained from the linear approximation if the constant C and the
partial derivatives at the nominal point are known.
The method requires selecting a starting (nominal) value of the variable from which their incremental
values and the values of the function are calculated. If the incremental values of the variables are greater
than some small error tolerance value, new values of the variables are calculated. Using these new
values as nominal, incremental values of the variables and new value of the function are again
calculated. This procedure is repeated until two consecutive iterations of the parameters are within the
predefined tolerance.
For a system of N non-linear equations with N variables, the system of linear equations obtained using
the Taylor series will be:

∂y1 m ∂y1 m ∂y
∆xi + m ∆x2 + ...... + m1 ∆xnm = C1 − y1m
∂x1
m
∂x2 ∂xn
∂y2 m ∂y2 m ∂y
∆x1 + m ∆x2 + ...... + m2 ∆xnm = C2 − y2m
∂x1
m
∂x2 ∂xn
(3.28)
....................................................................
∂yn m ∂yn m ∂y
∆x1 + m ∆x2 + ....... + mn ∆xnm = Cn − ynm
∂x1 m
∂x2 ∂xn
The above system of linear equations can be written in matrix form, in which the main matrix will be
that whose elements are the partial derivatives of the functions, and two column matrices: a column
matrix of increments of the variables =∆?@, and a column matrix of the incremental values of the
functions about the constants, =∆ @. The matrix of partial differentials is called the Jacobian matrix, and
denoted, =A@,

 ∂y1 ∂y1 ∂y1 


 ∂x ..
∂x2 ∂xn   ∆x1   ∆y1 
 1 
 ∂y2 ∂y2
..
∂y2  ∆x  ∆y 
 2  2
 ∂x1 ∂x2 ∂xn  ×   =  (3.29)
 .. .. .. 
.. .. ..     
 ∂y ∂yn ∂yn  ∆xn  ∆yn 
 n 
 ∂x1 ∂x2 ∂xn 
or in a more compact form as:
=A@=∆?@ = =∆ @
By finding the inverse of the Jacobian matrix and the nominal (initial) values of the functions, we can
determine the incremental values of the variables. Then using the method of successive approximations,
we get the new or corrected values of these variables. That is,
=∆?@ = =A@B =∆ @
and
=?@ = =?@ + =∆?@

(b) Application of N-R method to the SLPE solution


In order to use the above method in the solution of the load-flow problem, we use the SLFE written in
trigonometric form (i.e., Eq. 3.11), in which the functions are the net real and reactive power injections

EE 307_Topic #3 (Sem. I, 2014/2015) 12 of 20


at any bus and the state variables are the voltage angles and magnitudes at the various buses for which
the load-flow equations are written. The constants correspond to the specified or known values of the
net real and reactive power injections at the various buses. Further, the state vector of unknown voltage
angles and magnitudes is ordered such that

 δ1 
δ 
 2 
 .. 
 
[x] =  N
δ  =  [δ ] 
V1  [V ]
   
 V2 
 .. 
 
 V N 
and the nonlinear function increment [∆y] ordered so that the first components correspond to real power
and the last ones to reactive power

[∆P([x])]
[∆y ] = [∆f ( x)] =  
[∆Q([x ])]
The functions [∆P([x ])] and [∆Q([x])] are called the real and reactive power mismatches since they are
given by the difference between the values of the specified and the calculated net real power and
reactive power injections, respectively.
We can now write equation (3.28 or 3.29) using power system variables (for all buses except the swing
bus, which is denoted zero) to obtain equation (3.30) in, which the Jacobian matrix is partitioned into
four submatrices (JA, JB, JC, JD), the column matrix of the variables into two submatrices (∆δ, ∆V), and
column matrix of function increments into two submatrices (∆P, ∆Q).

[J A ] [J B ] [∆δ ] [∆P ]


[J ] × = (3.30)
 C [J D ] [∆V ] [∆Q]
In Eq. (3.30), the elements of sub-matrices of the Jacobian are as follows:

[J A ] =  ∂P  — Partial derivatives of real power with respect to voltage angles


 ∂δ 

 
[J B ] =  ∂P  — Partial derivatives of real power with respect to voltage magnitudes
 ∂ V 

[J C ] =  ∂Q  — Partial derivatives of reactive power with respect to voltage angles


 ∂δ 
 
[J D ] =  ∂Q  — Partial derivatives of reactive power with respect to voltage magnitudes
 ∂ V 
In the iterative process, the elements of the Jacobian are found by taking the partial derivatives of the
expressions for the net injected real and reactive powers at the various buses and substituting therein the
voltages assumed for the first iteration or calculated in the last previous iteration.

EE 307_Topic #3 (Sem. I, 2014/2015) 13 of 20


The equation (3.30) is solved by inverting the Jacobian. The values found for the increments of voltage
magnitude and phase are added to the previous values of voltage magnitude and phase to obtain the new
values for the calculated real and reactive powers injected at the various buses for starting the next
iteration. The process is repeated until the precision index applied to the power incremental and/or
variable incremental is satisfied. To achieve convergence, however, the initial estimation of voltage
must be reasonable.
Voltage-controlled buses are taken into account easily. Since the voltage magnitude is constant at such a
bus we omit in the Jacobian the column of partial derivatives with respect to voltage magnitude of the
bus. We are not interested at this point in the value of reactive power at the bus so we omit the row of
partial derivatives of the reactive power for the voltage-controlled bus. The value of reactive power at
this bus can be determined after convergence by Eq. (3.11).
(c) General algorithm of the N-R method
The following is the general algorithm for load flow solution by the N-R iteration:
1. For an N bus system in N unknowns, assemble the set of 2N independent equations.
2. Assume an initial solution for busbar voltages (magnitude and phase angles), except the swing bus
and voltage-controlled bus (for magnitude).
3. Determine the calculated real and reactive power injections at every bus at this solution (i.e., for the
first iteration or the most recently determined voltages for subsequent iterations) from Eq. (3.11).
4. Calculate the real and reactive power mismatches (i.e., power changes or increments) at every bus.
If the mismatches are very small, stop the iteration and take the recently determined voltages as the
true solution.
5. Form the linear equations (3.30) (i.e., calculate the Jacobian matrix at the most recently determined
solution) and solve for voltage magnitude and angle increments (corrections or changes) by
inversion of the Jacobian matrix. Gaussian elimination method may also be used.
6. Calculate an improved voltage solution by adding the corrections obtained in step #5 to the initially
assumed solutions of step #2.
7. Return to step #3 and repeat the process using the most recently calculated values of voltage
magnitudes and phase angles until either values of power mismatches or all values of voltage
corrections are sufficiently small.
8. Determine the active and reactive power injections at the swing bus and reactive power at the
voltage-controlled bus after convergence using the appropriate equations. Also, determine the power
flows on the various branches.

3.4.3. DE-COUPLED POWER-FLOW SOLUTION METHOD


The de-coupled method is basically an extension of the Newton-Raphson method and exploits the
property of electrical power networks under steady state conditions, where a strong interdependence
(strong coupling) exists between real power injections (flows) and bus voltage phase angles, as well as
between reactive power injections (flows) and voltage magnitudes. In other words for a small change in
the magnitude of the bus voltage, the real power injected at the bus does not change significantly (i.e.,
∂P
= 0 ) and similarly for a small change in phase angle of the bus voltage, the reactive power does not
∂V

change significantly (i.e., ∂Q = 0 ). It amounts to neglecting [JB] and [JC] sub-matrices of the exact
∂δ
Newton-Raphson method.

EE 307_Topic #3 (Sem. I, 2014/2015) 14 of 20


The obvious advantage of this method is that real power flow and reactive power flow equations may be
solved independently (as two systems of uncoupled linear equations), reducing the storage requirements
and possibly the computation per iteration.

 ∂P 
 ∂δ [∆δ ] = [∆P]
 
 ∂Q 
 [ ∆V ] = [∆Q]
 ∂ V 
There exists also a modification of this method, which is often referred to as fast de-coupled method. In
this method the Jacobian matrix values of the de-coupled method are made constant in value throughout
the iteration (i.e., they are not updated with the progression of iteration).
3.4.4. DC LOAD FLOW MODEL
The linearised DC power flow method greatly simplifies the power flow solution by making a number
of approximations, including:
(i) completely ignoring the reactive power balance equations;
(ii) assuming all bus voltage magnitudes are identically 1.0 per unit;

(iii) ignoring the line losses (i.e., assuming x Line >> rLine ); and
(iv) ignoring tap dependence in the transformer reactances.
The implication of these assumptions is that the DC power reduces the actual power flow problem to a
set of linear equations:

[ ]
N N
Pi = ∑ Bik Vk Vi cos 90 − (δ i − δ k ) = ∑ Bik sin(δ i − δ k )
0

k =1 k =1

where Bik is the series susceptance of the line (i.e., the inverse of the series reactance of the line, Xik), θik
= 900 and Vi = V k = 1.0 .

If the difference between the voltage angles is negligible, then the sine function will tend to the
difference itself and the above load flow equation may be written in the following linear form:
N N N
Pi = ∑ Bik ϕ ik = ∑ Bik δ i + ∑ (− Bik δ k )
k =1 k =1 k =1

= Bi1 (δ i − δ 1 ) + Bi 2 (δ i − δ 2 ) + ...... + Bi 0 δ i + ...... + BiN (δ i − δ N )


= − Bi1δ 1 − Bi 2 δ 2 ........ + ( Bi 0 + Bi1 + .... + BiN )δ i ....... − BiN δ N
This set of N equations, for an (N+1) bus system (bus 0 being the reference bus) can be written in the
form:
[P ] = [B ][δ ]
where:

EE 307_Topic #3 (Sem. I, 2014/2015) 15 of 20


[P] = [P1 P2 .....PN ]T - vector of net real power injections
[δ ] = [δ 1δ 2 ....δ N ]T - vector of bus voltage angles
 B11 B12 .. B1N 
B B22 .. B2 N 
[B] =  21 - bus susceptance matrix
 .. .. .. .. 
 
 B N1 BN 2 .. B NN 
Since the above set of equations is linear, it always has a single solution, which can be directly
calculated by solving:

[δ ] = [B]−1 [P] = [X ][P]


for δ using matrix techniques, thus eliminating the need for iterations.
The above linearised load flow model can be interpreted as the model for a network of resistors fed by
DC current sources where [P] is the vector of nodal current injections, [δ] is the nodal vector of DC
voltages, and [B] is the nodal conductance matrix.
Computationally, the formulated DC model has at least three advantages over the standard N-R power
flow formulation. First, by just solving the real power balance equations its equation set is about half the
size of the full AC problem. Secondly, the DC power flow is non-iterative, requiring just one single
solution. Third, because the susceptance matrix is state-independent provided the system topology does
not change, it need only be factored once.
3.4.5. Comparison of Load-Flow Solution Methods
The following is a brief summary of the salient characteristics of the load-flow methods discussed above
and hence their relative strengths and weaknesses.
(a) G-S method works well when programmed using rectangular coordinates, whereas N-R requires
more memory when rectangular coordinates are used. Hence, polar coordinates are preferred for N-
R solution method.
(b) The G-S method requires the fewest number of arithmetic operations to complete an iteration. This
is because of the sparsity of the network matrix and the simplicity of the solution technique. With N-
R, the elements of the Jacobian are to be computed in each iteration, so that the time is considerably
longer. For typical large systems the time per iteration in N-R is roughly equivalent to seven times
that of the G-S method. However, for both methods, the time per iteration increases almost directly
as the number of buses of the network.
(c) The rate of convergence of the G-S method is slow (because of its linear convergence
characteristic), requiring a considerably greater number of iterations to obtain a solution than the N-
R method, which has a quadratic convergence characteristics and is the best among all methods
from the standpoint of convergence. The fast-decoupled load-flow solution method has a geometric
convergence characteristic.
(d) The number of iterations for the G-S method increases directly as the number of buses in the
network, whereas the number of iterations for N-R method remains practically constant,
independent of system size.
(e) Convergence of the G-S solution method is affected by the choice of the swing bus and the presence
of series capacitor, while the sensitivity of N-R is minimal to these factors.

EE 307_Topic #3 (Sem. I, 2014/2015) 16 of 20


Solution Method
S.No. Problem Type
G-S Method N-R Method
1 Heavily loaded systems Usually cannot solve Solves systems with
systems with more than 70 shifts up to 90 degrees
degree phase shift
2 Systems containing negative Unable to solve Solves with ease
reactance such as series line
capacitor
3 Systems with slack bus at a desired Often requires trial and More tolerant of slack
location error to find a slack bus bus location
location that will yield a
solution
4 Long and short lines terminating on Usually cannot solve if Can solve a system with
the same bus long-to-short ratio is above a long-to-short ratio at
1000:1 any bus of 1,000,000:1
5 Long radial type of system Difficulty in solving Solves a wide range of
such problems
6 Acceleration factors Number of iterations None required
depends on choice of factor

3.5. METHODS OF LOAD-FLOW CONTROL


The results obtained from a load-flow study may not meet or satisfy all of the system’s operating
constraints. That is, load-flow analysis may reveal a violation of the thermal limits due to excessive I2R
heating on the line; voltage limits at the receiving buses; stability limits as measured by the phase angle
difference between sending and receiving bus voltages. If these unfavourable conditions are detected,
some corrective actions which influence the flow of power through the network should be taken. This is
what we refer to in this subsection as load-flow control.
The different load-flow control measures may be divided into two broad groups: (i) design measures,
and (ii) operational measures. The design measures involve constructing new facilities (i.e.,
transmission lines or generating stations) to address the above reasons. Such measures are generally
more involving and expensive. Operational techniques, on the other hand, achieve network load-flow
control by changing the operating characteristics of the system or system components, which are
already in service. These measures are relatively much simpler and less expensive.
The basis of practically all of the operational load-flow control measures is the power transfer equation
given below; with the only exceptions being the operational measures involving control of generator
output power and use of traditional capacitor banks to inject reactive power into the network (hence
changing the reactive power output of the generators).
3 3
= C;8 D − D3 = C;8D
? ?
This equation contains five system parameters or quantities that influence the active power flow in a line
or line section: (i) V1 – voltage magnitude at the sending end of a line; (ii) V2 – voltage magnitude at the
receiving end of a line; (iii) X – the line’s reactance or impedance; (iv) δ1 – voltage phase angle at the
sending-end; and (v) δ2 – voltage phase angle at the receiving end.

V1, δ1 V2, δ2
System X System
#1 #2
P

EE 307_Topic #3 (Sem. I, 2014/2015) 17 of 20


3.5.1. Load-Flow Control Using Synchronous Generators
This control method is realised by use of synchronous generators. By controlling the amount of real and
reactive power output of synchronous generators forming part of a large power system, we can control
the power flowing into the network.
To explain this, let us consider Figure 4.4.1. In this figure the generator at the i-th bus is modelled by a
synchronous reactance XG,i and voltage behind the synchronous reactance EG,i. Applying Kirchhoff’s
voltage law and assuming phase angle of the terminal (bus) voltage to be equal to zero, we get

Vi EG ,i ∠(90 − δ ) − Vi 2∠90
SG ,i = PG ,i + jQG ,i = Vi I i* = (1)
X G ,i

It follows from the above formula that,


Vi EG ,i Vi EG , i
PG ,i = cos(90 − δ ) = sin δ (2)
X G ,i X G ,i

= i (EG ,i cos δ − Vi )
Vi EG ,i Vi 2
sin (90 − δ ) −
V (3)
QG , i =
X G ,i X G , i X G ,i

Equation (2) shows that if Vi and EG,i are held constant, the real-power output of the generator is directly
dependent on the load (power) angle, δ. Of course, in steady-state operation real electric power output
can only be increased by an increase of real mechanical power input. Therefore, the power angle must
also be directly dependent on the mechanical input. The mechanical power input is normally controlled
by use of appropriate governor settings. In Equation (3), if Vi and EG,i are held constant, Q would appear
to be affected by changes in the power angle or changes in mechanical power input. However, the
normal operating range for the power angle is less than 150. For small changes of the power angle in this
operating range, the cosine term is relatively unchanged while the sine term will vary significantly. This
implies that reactive-power output of a generator is not directly dependent on its mechanical power
input. It is primarily dependent on the magnitude of the generator voltage, EG,i if Vi and δ are held
constant. For a generator operating at a constant speed, the generator voltage is a function of the DC
excitation on the rotor.
From this discussion the reactive power output of a generator appears to be controlled by adjusting the
rotor DC excitation, and the real-power output appears to be controlled by varying the mechanical
power input. By controlling the amount of real and reactive power generated at each generating station
in a power system, the general flow of power in the system can be controlled.
3.5.2. Load-Flow Control Using kVAr Sources
Most loads in a power system are inductive. Reactive power must, therefore, be generated to supply this
inductive load. On the real power that is delivered to the load actually does real mechanical work.
However, delivering the same amount of real power to a load with a reactive part requires more current
than supplying a load with no reactive part. Larger currents mean greater voltage drops along
transmission and distribution lines and more ohmic line losses. Both of these conditions are undesirable
since they may result in violation of line thermal, voltage and stability constraints.
One way to overcome these disadvantages is to generate reactive power at the location that it is needed.
For this purpose, a variety of reactive power (VAR) generators can be used. But one source of reactive
power, which is widely used in power systems, is the capacitor banks. Capacitors consume reactive
power, which is the same as supplying VARs. Therefore, the installation of a capacitor bank at the load
bus will have the effect of reducing the amount of current required from the generating station to supply
a given amount of load. This phenomenon is illustrated in Figure 4.1.3. The result of adding the properly

EE 307_Topic #3 (Sem. I, 2014/2015) 18 of 20


sized capacitor at the load is that no reactive power (Q) must flow through the lines of the system to
supply the load reactance. That is, use of VAR generators controls the flow in reactive power in the
system.
3.5.3. Load-Flow Control Using Transformers
Transformers provide a convenient means of controlling the flow of both real and reactive power along
a transmission line by redistributing the flow of power between transmission or distribution lines. They
effect this control by changing the bus voltage magnitudes and phase angles, which may be performed
using tap-changing transformers and regulating transformers. The former controls only the voltage
magnitude, while the latter can be used to control both magnitude and phase angle (i.e., both as a
booster and phase shifter).
(a) Load-flow Control Using Tap-Changing Transformers
Tap-changing transformers (both under-no-load and under-load) effects load-flow control by changing
the magnitude of the voltage at the bus connected to its secondary (i.e., boosting the secondary voltage
without changing the phase angle).
(b) Load-flow Control Using Regulating Transformers
Regulating transformers are transformers specially designed to change, by a relatively small amount, the
voltage in a point of the system as shown in the figure below. They are not used to transform a large
amount of energy. The voltage change introduced ∆V, by regulating transformers, can in general be
changed in both magnitude and phase angle. A regulating transformer that adds a voltage increment of
variable magnitude is often referred to as booster transformer, while that which adds a voltage
increment of variable phase is called a phase-shifting transformer. Thus, a booster transformer effects
load-flow control through voltage magnitude control (VMC), while the phase-shifting transformer
effects the same through phase-angle control (PAC).
Below, we describe the operation of these two types of regulating transformers and how they achieve
load-flow control in the links where they are connected. That is, we illustrate that the added voltage
magnitude or phase angle, although of small magnitude, can have a drastic effect on the power flow in
the link in question (i.e., in the line where the RT is connected). Regulating transformers in lines effects
control of power flow on those lines by modifying the bus impedance matrix, thereby modifying the
load flow solution.
Va Va + ∆Va

Vb Regulating Vb + ∆Vb
Transformer

Vc Vc + ∆Vc

3.5.4. Load-Flow Control Using FACTS Devices


FACTS is an acronym for flexible AC transmission systems and refers to a family of power electronics-
based devices or other static equipment which provide control of one or more AC transmission system
parameters/quantities (e.g., current, voltage, reactance, phase angle) to enhance controllability and
increase power transfer capability. FACTS achieve load-flow control not by introducing external power
generation, but by influencing the parameters contained in the power transfer equation. FACTS, thus,
works like dependent (or controlled) sources of voltage or current, where the controlling quantity can be
any of the quantities in the power transfer equation. FACTS can be connected in series or in parallel
relative to the line whose flow is to be controlled. FACTS working as dependent voltage sources are
often connected in series with the line, while FACTS designed to work as dependent current sources are

EE 307_Topic #3 (Sem. I, 2014/2015) 19 of 20


often connect as a shunt to the line. FACTS may also be connected both in series and in shunt with the
system or line.
Some of the most common of these devices are:
(a) Static Shunt Compensators
(b) Static Series Compensators
(c) Unified Power Flow Controllers (UPFC)
(d) Inter-Line Power Flow Controller (IPFC)
(e) High-Voltage DC Transmission (HVDC) Devices
Table 1: Summary of Load-Flow Controllers
S.No. Control Equipment or Device Controllable Parameter(s) Control Mechanism
(i.e., Underlying
Principle of Operation)
1 Synchronous Generators - Voltage Magnitude
- Voltage Phase Angle
2 Shunt Capacitor Bank - Voltage Magnitude
3 Transformer Load-Tap Changer (LTC) - Voltage Magnitude
4 Regulating Transformer – Booster - Voltage Magnitude
Transformer
5 Regulating Transformer – Phase-Shifting - Voltage Phase Angle
Transformer
6 FACTS (Flexible AC Transmission - “Apparent” Line
System) Devices – Series FACTS Impedance
Controller; Shunt FACTS Controller; - Voltage Magnitude
Combined Series-Series FACTS - Voltage Phase Angle
Controller; and Combined Series-Shunt
FACTS Controller
NB: The series and/or shunt FACTS
controllers may take the form of variable
impedance, variable source (dependent
voltage source or dependent current
source), or a combination of variable
impedance & variable source
7 Line Upgrade or Reconductoring, that is, -
Replacing Line Conductor with that of a
Higher Current Rating (NB: This is a
means of enhancing the loading or power-
transfer capability of a line, rather than a
means of regulating the current/power
flow through it)
8 Conversion of a Single-Circuit Line to a -
Double-Circuit Line or Line Phase
Splitting (NB: This is a means of
enhancing the loading or power-transfer
capability of a line, rather than a means of
regulating the current/power flowing
through it)

– END –

EE 307_Topic #3 (Sem. I, 2014/2015) 20 of 20

You might also like