100% found this document useful (1 vote)
349 views

Power Factor: What Is The Difference Between Lagging Power Factor and Leading Power Factor?

Power factor is the ratio of real power to apparent power in an AC circuit. A power factor of 1 means the current and voltage waves are in phase, while a power factor less than 1 means they are out of phase due to reactive loads like inductors and capacitors. A lagging power factor occurs when the current lags the voltage, as with an inductive load, while a leading power factor occurs when the current leads the voltage, as with a capacitive load. Power factor correction techniques like capacitors and inductors can be used to improve the power factor by offsetting the reactive effects of loads.

Uploaded by

Malik Jameel
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as DOCX, PDF, TXT or read online on Scribd
100% found this document useful (1 vote)
349 views

Power Factor: What Is The Difference Between Lagging Power Factor and Leading Power Factor?

Power factor is the ratio of real power to apparent power in an AC circuit. A power factor of 1 means the current and voltage waves are in phase, while a power factor less than 1 means they are out of phase due to reactive loads like inductors and capacitors. A lagging power factor occurs when the current lags the voltage, as with an inductive load, while a leading power factor occurs when the current leads the voltage, as with a capacitive load. Power factor correction techniques like capacitors and inductors can be used to improve the power factor by offsetting the reactive effects of loads.

Uploaded by

Malik Jameel
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as DOCX, PDF, TXT or read online on Scribd
You are on page 1/ 64

Power factor

The ratio of watts average power (the average power measured in watts) to the
apparent power of an alternating-current circuit. By definition, the equation below
holds, which is the ratio of instrument readings.

Using highly inductive loads may lead to reduction in the power factor. or
consumption of unappreciable amount of power also results in the same.

Methods to improve PF----- using synchronous condensers --- using capacitor


banks

What is the difference between lagging power factor and leading power
factor?

Power factor is the ratio of watts (true power) to VA (volt-amperes, also called apparent power).
Where the load is resistive only, the power factor is one, or unity, because the voltage waveform
and the current waveform are in phase. Thus, for resistive loads only, true power and VA are the
same.

Where the load is reactive, the load stores energy, releasing it during a different part of the cycle.
This shifts the current waveform so that it is offset, or out of phase with the voltage waveform.

Reactive loads can be inductive (electric motors), capacitive, or non-linear


(rectifier power supplies).

When the load is inductive, the inductance tends to oppose the flow of current,
storing energy then releasing it later in the cycle. The current waveform lags
behind the voltage waveform. When the load is capacitive, the opposite occurs, and
the current waveform leads the voltage waveform.

So, lagging vs. leading is another way of saying the net reactance is either
inductive or capacitive.

This is slightly simplistic, and what we are talking about above is really DPF, or
Displacement Power Factor. Non-linear loads don't really shift the current
waveform, they distort it. The current waveform starts to look like a square wave,
and square waves contain harmonics. So non-linear loads add harmonic distortion,
and this tends to mimic a capacitively reactive load, adding some leading power
factor. So when we say power factor, we really must include DPF plus harmonic
distortion in total.

One memory aid that may help to remember all this is:

ELI the ICE man

The L in ELI means inductance. The E (voltage) comes first, then the I (current)
lags behind. Inductive reactance produces a lagging power factor.

The C in ICE means capacitance. The I (current) comes first (leads) then the E
(voltage) comes later. Capacitive reactance produces a leading power factor.

Remember, it's always the current waveform that is affected by the reactive load,
so you have to think about whether the current is leading or lagging.

Most reactive loads are inductive, so at most s

A watt-meter indicates average power, and electrodynamometer or iron-vane


instruments show rms voltage and current. For the steady-state ac circuit under
sinusoidal voltage and current, pf = cos θ, where θ is the phase angle between the
voltage and current. This definition is restricted to sine waves of the same
frequency.

The power factor of an AC electric power system is defined as the ratio of the real
power flowing to the load to the apparent power,[1][2] and is a number between 0
and 1 (frequently expressed as a percentage, e.g. 0.5 pf = 50% pf). Real power is
the capacity of the circuit for performing work in a particular time. Apparent
power is the product of the current and voltage of the circuit. Due to energy stored
in the load and returned to the source, or due to a non-linear load that distorts the
wave shape of the current drawn from the source, the apparent power can be
greater than the real power.

In an electric power system, a load with low power factor draws more current than
a load with a high power factor for the same amount of useful power transferred.
The higher currents increase the energy lost in the distribution system, and require
larger wires and other equipment. Because of the costs of larger equipment and
wasted energy, electrical utilities will usually charge a higher cost to industrial or
commercial customers where there is a low power factor.

Linear loads with low power factor (such as induction motors) can be corrected
with a passive network of capacitors or inductors. Non-linear loads, such as
rectifiers, distort the current drawn from the system. In such cases, active power
factor correction is used to counteract the distortion and raise power factor. The
devices for correction of power factor may be at a central substation, or spread out
over a distribution system, or built into power-consuming equipment.

Power factor in linear circuit

Instantaneous and average power calculated from AC voltage and current with a
unity power factor (φ=0, cosφ=1)
Instantaneous and average power calculated from AC voltage and current with a
zero power factor (φ=90, cosφ=0)

Instantaneous and average power calculated from AC voltage and current with a
lagging power factor (φ=45, cosφ=0.71)

In a purely resistive AC circuit, voltage and current waveforms are in step (or in
phase), changing polarity at the same instant in each cycle. Where reactive loads
are present, such as with capacitors or inductors, energy storage in the loads result
in a time difference between the current and voltage waveforms. This stored
energy returns to the source and is not available to do work at the load. Thus, a
circuit with a low power factor will have higher currents to transfer a given
quantity of real power than a circuit with a high power factor. A linear load does
not change the shape of the waveform of the current, but may change the relative
timing (phase) between voltage and current.

Circuits containing purely resistive heating elements (filament lamps, strip heaters,
cooking stoves, etc.) have a power factor of 1.0. Circuits containing inductive or
capacitive elements (compact fluorescent lamps,[3] lamp ballasts, motors, etc.) often
have a power factor below 1.0.

Definition and calculation

AC power flow has the three components: real power (P), measured in watts (W);
apparent power (S), measured in volt-amperes (VA); and reactive power (Q),
measured in reactive volt-amperes (var).

The power factor is defined as:


. (λ = P/S)

In the case of a perfectly sinusoidal waveform, P, Q and S can be expressed as


vectors that form a vector triangle such that:

If φ is the phase angle between the current and voltage, then the power factor is
equal to , and:

Since the units are consistent, the power factor is by definition a dimensionless
number between 0 and 1. When power factor is equal to 0, the energy flow is
entirely reactive, and stored energy in the load returns to the source on each cycle.
When the power factor is 1, all the energy supplied by the source is consumed by
the load. Power factors are usually stated as "leading" or "lagging" to show the
sign of the phase angle.

If a purely resistive load is connected to a power supply, current and voltage will
change polarity in step, the power factor will be unity (1), and the electrical energy
flows in a single direction across the network in each cycle. Inductive loads such as
transformers and motors (any type of wound coil) consume reactive power with
current waveform lagging the voltage. Capacitive loads such as capacitor banks or
buried cable generate reactive power with current phase leading the voltage. Both
types of loads will absorb energy during part of the AC cycle, which is stored in
the device's magnetic or electric field, only to return this energy back to the source
during the rest of the cycle.

For example, to get 1 kW of real power, if the power factor is unity, 1 kVA of
apparent power needs to be transferred (1 kW ÷ 1 = 1 kVA). At low values of
power factor, more apparent power needs to be transferred to get the same real
power. To get 1 kW of real power at 0.2 power factor, 5 kVA of apparent power
needs to be transferred (1 kW ÷ 0.2 = 5 kVA). This apparent power must be
produced and transmitted to the load in the conventional fashion, and is subject to
the usual distributed losses in the production and transmission processes.

Linear loads

Electrical loads consuming alternating current power consume both real power and
reactive power. The vector sum of real and reactive power is the apparent power.
The presence of reactive power causes the real power to be less than the apparent
power, and so, the electric load has a power factor of less than 1.

Power factor correction of linear loads

It is often desirable to adjust the power factor of a system to near 1.0. This power
factor correction is achieved by switching in or out banks of inductors or
capacitors. For example the inductive effect of motor loads may be offset by
locally connected capacitors. When reactive elements supply or absorb reactive
power near the load, the apparent power is reduced.

Power factor correction may be applied by an electrical power transmission utility


to improve the stability and efficiency of the transmission network. Correction
equipment may be installed by individual electrical customers to reduce the costs
charged to them by their electricity supplier. A high power factor is generally
desirable in a transmission system to reduce transmission losses and improve
voltage regulation at the load.

Power factor correction brings the power factor of an AC power circuit closer to 1
by supplying reactive power of opposite sign, adding capacitors or inductors which
act to cancel the inductive or capacitive effects of the load, respectively. For
example, the inductive effect of motor loads may be offset by locally connected
capacitors. If a load had a capacitive value, inductors (also known as reactors in
this context) are connected to correct the power factor. In the electricity industry,
inductors are said to consume reactive power and capacitors are said to supply it,
even though the reactive power is actually just moving back and forth on each AC
cycle.

The reactive elements can create voltage fluctuations and harmonic noise when
switched on or off. They will supply or sink reactive power regardless of whether
there is a corresponding load operating nearby, increasing the system's no-load
losses. In a worst case, reactive elements can interact with the system and with
each other to create resonant conditions, resulting in system instability and severe
overvoltage fluctuations. As such, reactive elements cannot simply be applied at
will, and power factor correction is normally subject to engineering analysis.
1. Reactive Power Control Relay; 2. Network connection points; 3. Slow-blow
Fuses; 4. Inrush Limiting Contactors; 5. Capacitors (single-phase or three-phase
units, delta-connection); 6. Transformer Suitable voltage transformation to suit
control power (contactors, ventilation,...)

An automatic power factor correction unit is used to improve power factor. A


power factor correction unit usually consists of a number of capacitors that are
switched by means of contactors. These contactors are controlled by a regulator
that measures power factor in an electrical network. To be able to measure power
factor, the regulator uses a current transformer to measure the current in one phase.

Depending on the load and power factor of the network, the power factor controller
will switch the necessary blocks of capacitors in steps to make sure the power
factor stays above a selected value (usually demanded by the energy supplier), say
0.9.
Instead of using a set of switched capacitors, an unloaded synchronous motor can
supply reactive power. The reactive power drawn by the synchronous motor is a
function of its field excitation. This is referred to as a synchronous condenser. It is
started and connected to the electrical network. It operates at full leading power
factor and puts vars onto the network as required to support a system’s voltage or
to maintain the system power factor at a specified level.

The condenser’s installation and operation are identical to large electric motors. Its
principal advantage is the ease with which the amount of correction can be
adjusted; it behaves like an electrically variable capacitor. Unlike capacitors, the
amount of reactive power supplied is proportional to voltage, not the square of voltage;
this improves voltage stability on large networks. Synchronous condensors are often used in
connection with high voltage direct current transmission projects or in large industrial plants
such as steel mills.

Non-linear loads
A non-linear load on a power system is typically a rectifier (such as used in a
power supply), or some kind of arc discharge device such as a fluorescent lamp,
electric welding machine, or arc furnace. Because current in these systems is
interrupted by a switching action, the current contains frequency components that
are multiples of the power system frequency.

Non-sinusoidal components

Non-linear loads change the shape of the current waveform from a sine wave to
some other form. Non-linear loads create harmonic currents in addition to the
original (fundamental frequency) AC current. Addition of linear components such
as capacitors and inductors cannot cancel these harmonic currents, so other
methods such as filters or active power factor correction are required to smooth out
their current demand over each cycle of alternating current and so reduce the
generated harmonic currents.

In circuits having only sinusoidal currents and voltages, the power factor effect
arises only from the difference in phase between the current and voltage. This is
narrowly known as "displacement power factor". The concept can be generalized
to a total, distortion, or true power factor where the apparent power includes all
harmonic components. This is of importance in practical power systems which
contain non-linear loads such as rectifiers, some forms of electric lighting, electric
arc furnaces, welding equipment, switched-mode power supplies and other
devices.
A typical multimeter will give incorrect results when attempting to measure the
AC current drawn by a non-sinusoidal load. A true RMS multimeter must be used
to measure the actual RMS currents and voltages (and therefore apparent power).
To measure the real power or reactive power, a wattmeter designed to properly
work with non-sinusoidal currents must be used.

Switched-mode power supplies

A particularly important class of non-linear loads is the millions of personal


computers that typically incorporate switched-mode power supplies (SMPS) with
rated output power ranging from a few watt to more than 1 kW. Historically, these
very-low-cost power supplies incorporated a simple full-wave rectifier that
conducted only when the mains instantaneous voltage exceeded the voltage on the
input capacitors. This leads to very high ratios of peak-to-average input current,
which also lead to a low distortion power factor and potentially serious phase and
neutral loading concerns.

A typical switched-mode power supply first makes a DC bus, using a bridge


rectifier or similar circuit. The output voltage is then derived from this DC bus.
The problem with this is that the rectifier is a non-linear device, so the input
current is highly non-linear. That means that the input current has energy at
harmonics of the frequency of the voltage.

This presents a particular problem for the power companies, because they cannot
compensate for the harmonic current by adding simple capacitors or inductors, as
they could for the reactive power drawn by a linear load. Many jurisdictions are
beginning to legally require power factor correction for all power supplies above a
certain power level.

Regulatory agencies such as the EU have set harmonic limits as a method of


improving power factor. Declining component cost has hastened implementation
of two different methods. To comply with current EU standard EN61000-3-2, all
switched-mode power supplies with output power more than 75 W must include
passive PFC, at least. 80 PLUS power supply certification requires a power factor
of 0.9 or more.[4]

Passive PFC

The simplest way to control the harmonic current is to use a filter: it is possible to
design a filter that passes current only at line frequency (e.g. 50 or 60 Hz). This
filter reduces the harmonic current, which means that the non-linear device now
looks like a linear load. At this point the power factor can be brought to near unity,
using capacitors or inductors as required. This filter requires large-value high-
current inductors, however, which are bulky and expensive.

A passive PFC requires an inductor larger than the inductor in an active PFC, but
costs less.[5][6]

This is a simple way of correcting the nonlinearity of a load by using capacitor


banks. It is not as effective as active PFC[citation needed].[7][8][9][10][11]

Passive PFCs are typically more power efficient than active PFCs – a passive PFC
on a switching computer PSU has a typical power efficiency of around 96%, while
an active PFC has a typical efficiency of about 94%.[12]

Active PFC

An Active Power Factor Corrector (active PFC) is a power electronic system


that controls the amount of power drawn by a load in order to obtain a Power
factor as close as possible to unity. In most applications, the active PFC controls
the input current of the load so that the current waveform is proportional to the
mains voltage waveform (a sinewave).

Some types of active PFC are

1. Boost
2. Buck
3. Buck-boost

Active power factor correctors can be single-stage or multi-stage.

In the case of a switched-mode power supply, a boost converter is


inserted between the bridge rectifier and the main input capacitors. The
boost converter attempts to maintain a constant DC bus voltage on its
output while drawing a current that is always in phase with and at the
same frequency as the line voltage. Another switchmode converter
inside the power supply produces the desired output voltage from the
DC bus. This approach requires additional semiconductor switches and
control electronics, but permits cheaper and smaller passive components.
It is frequently used in practice. For example, SMPS with passive PFC
can achieve power factor of about 0.7–0.75, SMPS with active PFC, up
to 0.99 power factor, while a SMPS without any power factor correction
has a power factor of only about 0.55–0.65[citation needed]. Due to their very
wide input voltage range, many power supplies with active PFC can
automatically adjust to operate on AC power from about 100 V (Japan)
to 230 V (Europe). That feature is particularly welcome in power
supplies for laptops.

Importance of power factor in distribution systems

The significance of power factor lies in the fact that utility companies
supply customers with volt-amperes, but bill them for watts. Power
factors below 1.0 require a utility to generate more than the minimum
volt-amperes necessary to supply the real power (watts). This increases
generation and transmission costs. For example, if the load power factor
were as low as 0.7, the apparent power would be 1.4 times the real
power used by the load. Line current in the circuit would also be 1.4
times the current required at 1.0 power factor, so the losses in the circuit
would be doubled (since they are proportional to the square of the
current). Alternatively all components of the system such as generators,
conductors, transformers, and switchgear would be increased in size
(and cost) to carry the extra current.

Utilities typically charge additional costs to customers who have a


power factor below some limit, which is typically 0.9 to 0.95. Engineers
are often interested in the power factor of a load as one of the factors
that affect the efficiency of power transmission.

Measuring power factor

Power factor in a single-phase circuit (or balanced three-phase circuit)


can be measured with the wattmeter-ammeter-voltmeter method, where
the power in watts is divided by the product of measured voltage and
current. The power factor of a balanced polyphase circuit is the same as
that of any phase. The power factor of an unbalanced polyphase circuit
is not uniquely defined.
A direct reading power factor meter can be made with a moving coil meter of the
electrodynamic type, carrying two perpendicular coils on the moving part of the
instrument. The field of the instrument is energized by the circuit current flow. The
two moving coils, A and B, are connected in parallel with the circuit load. One
coil, A, will be connected through a resistor and the second coil, B, through an
inductor, so that the current in coil B is delayed with respect to current in A. At
unity power factor, the current in A is in phase with the circuit current, and coil A
provides maximum torque,driving the instrument pointer toward the 1.0 mark on
the scale. At zero power factor, the current in coil B is in phase with circuit current,
and coil B provides torque to drive the pointer towards 0. At intermediate values of
power factor, the torques provided by the two coils add and the pointer takes up
intermediate positions.[13]

Another electromechanical instrument is the polarized-vane type. [14] In this


instrument a stationary field coil produces a rotating magnetic field, just like a
polyphase motor. The field coils are connected either directly to polyphase voltage
sources or to a phase-shifting reactor if a single-phase application. A second
stationary field coil, perpendicular to the voltage coils, carries a current
proportional to current in one phase of the circuit. The moving system of the
instrument consists of two vanes which are magnetized by the current coil. In
operation the moving vanes take up a physical angle equivalent to the electrical
angle between the voltage source and the current source. This type of instrument
can be made to register for currents in both directions, giving a 4-quadrant display
of power factor or phase angle.

Digital instruments can be made that either directly measure the time lag between
voltage and current waveforms and so calculate the power factor, or by measuring
both true and apparent power in the circuit and calculating the quotient. The first
method is only accurate if voltage and current are sinusoidal; loads such as
rectifiers distort the waveforms from the sinusoidal shape.

Mnemonics

English-language power engineering students are advised to remember: "ELI the


ICE man" or "ELI on ICE" – the voltage E leads the current I in an inductor L, the
current leads the voltage in a capacitor C.

Or even shorter: CIVIL – in a Capacitor the I (current) leads Voltage, Voltage


leads I (current) in an inductor L.
Real, reactive, and apparent power

The apparent power is the vector sum of real and reactive power

Consider a simple alternating current (AC) circuit consisting of a source and a


load, where both the current and voltage are sinusoidal. If the load is purely
resistive, the two quantities reverse their polarity at the same time, the direction of
energy flow does not reverse, and only real power flows. If the load is purely
reactive, then the voltage and current are 90 degrees out of phase and there is no
net power flow. This energy flowing backwards and forwards is known as reactive
power. A practical load will have resistive, inductive, and capacitive parts, and so
both real and reactive power will flow to the load.

If a capacitor and an inductor are placed in parallel, then the currents flowing
through the inductor and the capacitor tend to cancel out rather than adding.
Conventionally, capacitors are considered to generate reactive power and inductors
to consume it. This is the fundamental mechanism for controlling the power factor
in electric power transmission; capacitors (or inductors) are inserted in a circuit to
partially cancel reactive power of the load.

The apparent power is the product of voltage and current. Apparent power is handy
for sizing of equipment or wiring. However, adding the apparent power for two
loads will not accurately give the total apparent power unless they have the same
displacement between current and voltage (the same power factor).

Engineers use the following terms to describe energy flow in a system (and assign
each of them a different unit to differentiate between them):

 Real power (P) - unit: watt (W)


 Reactive power (Q) - unit: volt-amperes reactive (var)
 Complex power (S) - unit: volt-ampere (VA)
 Apparent Power (|S|) , that is, the absolute value of complex power S - unit:
volt-ampere (VA)

In the diagram, P is the real power, Q is the reactive power (in this case positive),
S is the complex power and the length of S is the apparent power.

Reactive power does not transfer energy, so it is represented as the imaginary


basis. Real power moves energy, so it is the real basis.

The unit for all forms of power is the watt (symbol: W), but this unit is generally
reserved for real power. Apparent power is conventionally expressed in volt-
amperes (VA) since it is the product of rms voltage and rms current. The unit for
reactive power is expressed as "VAr", which stands for volt-amperes reactive.
Since reactive power flow transfers no net energy to the load, it is sometimes
called "wattless" power.

Understanding the relationship between these three quantities lies at the heart of
understanding power engineering. The mathematical relationship among them can
be represented by vectors or expressed using complex numbers,

(where j is the imaginary unit).

The complex value S is referred to as the complex power.

Power factor

The ratio between real power and apparent power in a circuit is called the power
factor. Where the waveforms are purely sinusoidal, the power factor is the cosine
of the phase angle (φ) between the current and voltage sinusoid waveforms.
Equipment data sheets and nameplates often will abbreviate power factor as "cosφ"
for this reason.

Power factor equals 1 when the voltage and current are in phase, and is zero when
the current leads or lags the voltage by 90 degrees. Power factors are usually stated
as "leading" or "lagging" to show the sign of the phase angle, where leading
indicates a negative sign. For two systems transmitting the same amount of real
power, the system with the lower power factor will have higher circulating currents
due to energy that returns to the source from energy storage in the load. These
higher currents in a practical system will produce higher losses and reduce overall
transmission efficiency. A lower power factor circuit will have a higher apparent
power and higher losses for the same amount of real power transfer.
Purely capacitive circuits cause reactive power with the current waveform leading
the voltage wave by 90 degrees, while purely inductive circuits cause reactive
power with the current waveform lagging the voltage waveform by 90 degrees.
The result of this is that capacitive and inductive circuit elements tend to cancel
each other out.

Reactive power flow

In power transmission and distribution, significant effort is made to control the


reactive power flow. This is typically done automatically by switching inductors or
capacitor banks in and out, by adjusting generator excitation, and by other means.
Electricity retailers may use electricity meters which measure reactive power to
financially penalize customers with low power factor loads. This is particularly
relevant to customers operating highly inductive loads such as motors at water
pumping stations.

Unbalanced polyphase systems

While real power and reactive power are well defined in any system, the definition
of apparent power for unbalanced polyphase systems is considered to be one of the
most controversial topics in power engineering. Originally, apparent power arose
merely as a figure of merit. Major delineations of the concept are attributed to
Stanley's Phenomena of Retardation in the Induction Coil (1888) and Steinmetz's
Theoretical Elements of Engineering (1915). However, with the development of
three phase power distribution, it became clear that the definition of apparent
power and the power factor could not be applied to unbalanced polyphase systems.
In 1920, a "Special Joint Committee of the AIEE and the National Electric Light
Association met to resolve the issue. They considered two definitions:

that is, the quotient of the sums of the real powers for each phase over the sum of
the apparent power for each phase.

that is, the quotient of the sums of the real powers for each phase over the
magnitude of the sum of the complex powers for each phase.
The 1920 committee found no consensus and the topic continued to dominate
discussions. In 1930 another committee formed and once again failed to resolve the
question. The transcripts of their discussions are the lengthiest and most
controversial ever published by the AIEE (Emanuel, 1993). Further resolution of
this debate did not come until the late 1990s.

Basic calculations using real numbers

A perfect resistor stores no energy, so current and voltage are in phase. Therefore
there is no reactive power and P = S. Therefore for a perfect resistor:

For a perfect capacitor or inductor on the other hand there is no net power transfer,
so all power is reactive. Therefore for a perfect capacitor or inductor:

Where X is the reactance of the capacitor or inductor.

If X is defined as being positive for an inductor and negative for a capacitor then
we can remove the modulus signs from Q and X and get.

Multiple frequency systems

Since an RMS value can be calculated for any waveform, apparent power can be
calculated from this.

For real power it would at first appear that we would have to calculate loads of
product terms and average all of them. However if we look at one of these product
terms in more detail we come to a very interesting result.
however the time average of a function of the form cos(ωt + k) is zero provided
that ω is nonzero. Therefore the only product terms that have a nonzero average are
those where the frequency of voltage and current match. In other words it is
possible to calculate real (average) power by simply treating each frequency
separately and adding up the answers.

Furthermore, if we assume the voltage of the mains supply is a single frequency


(which it usually is), this shows that harmonic currents are a bad thing. They will
increase the rms current (since there will be non-zero terms added) and therefore
apparent power, but they will have no effect on the real power transferred. Hence,
harmonic currents will reduce the power factor.

Harmonic currents can be reduced by a filter placed at the input of the device.
Typically this will consist of either just a capacitor (relying on parasitic resistance
and inductance in the supply) or a capacitor-inductor network. An active power
factor correction circuit at the input would generally reduce the harmonic currents
further and maintain the power factor closer to unity. And hence, the different
components of current is defined.

References

 "Understanding Power Factor" ST Application Note


 "AC Power Java Applet"

Real, reactive, and apparent power


The apparent power is the vector sum of real and reactive power

Consider a simple alternating current (AC) circuit consisting of a source and a


load, where both the current and voltage are sinusoidal. If the load is purely
resistive, the two quantities reverse their polarity at the same time, the direction of
energy flow does not reverse, and only real power flows. If the load is purely
reactive, then the voltage and current are 90 degrees out of phase and there is no
net power flow. This energy flowing backwards and forwards is known as reactive
power. A practical load will have resistive, inductive, and capacitive parts, and so
both real and reactive power will flow to the load.

If a capacitor and an inductor are placed in parallel, then the currents flowing
through the inductor and the capacitor tend to cancel out rather than adding.
Conventionally, capacitors are considered to generate reactive power and inductors
to consume it. This is the fundamental mechanism for controlling the power factor
in electric power transmission; capacitors (or inductors) are inserted in a circuit to
partially cancel reactive power of the load.

The apparent power is the product of voltage and current. Apparent power is handy
for sizing of equipment or wiring. However, adding the apparent power for two
loads will not accurately give the total apparent power unless they have the same
displacement between current and voltage (the same power factor).

Engineers use the following terms to describe energy flow in a system (and assign
each of them a different unit to differentiate between them):

 Real power (P) - unit: watt (W)


 Reactive power (Q) - unit: volt-amperes reactive (var)
 Complex power (S) - unit: volt-ampere (VA)
 Apparent Power (|S|) , that is, the absolute value of complex power S - unit:
volt-ampere (VA)
In the diagram, P is the real power, Q is the reactive power (in this case positive),
S is the complex power and the length of S is the apparent power.

Reactive power does not transfer energy, so it is represented as the imaginary


basis. Real power moves energy, so it is the real basis.

The unit for all forms of power is the watt (symbol: W), but this unit is generally
reserved for real power. Apparent power is conventionally expressed in volt-
amperes (VA) since it is the product of rms voltage and rms current. The unit for
reactive power is expressed as "VAr", which stands for volt-amperes reactive.
Since reactive power flow transfers no net energy to the load, it is sometimes
called "wattless" power.

Understanding the relationship between these three quantities lies at the heart of
understanding power engineering. The mathematical relationship among them can
be represented by vectors or expressed using complex numbers,

(where j is the imaginary unit).

The complex value S is referred to as the complex power.

Power factor

The ratio between real power and apparent power in a circuit is called the power
factor. Where the waveforms are purely sinusoidal, the power factor is the cosine
of the phase angle (φ) between the current and voltage sinusoid waveforms.
Equipment data sheets and nameplates often will abbreviate power factor as "cosφ"
for this reason.

Power factor equals 1 when the voltage and current are in phase, and is zero when
the current leads or lags the voltage by 90 degrees. Power factors are usually stated
as "leading" or "lagging" to show the sign of the phase angle, where leading
indicates a negative sign. For two systems transmitting the same amount of real
power, the system with the lower power factor will have higher circulating currents
due to energy that returns to the source from energy storage in the load. These
higher currents in a practical system will produce higher losses and reduce overall
transmission efficiency. A lower power factor circuit will have a higher apparent
power and higher losses for the same amount of real power transfer.

Purely capacitive circuits cause reactive power with the current waveform leading
the voltage wave by 90 degrees, while purely inductive circuits cause reactive
power with the current waveform lagging the voltage waveform by 90 degrees.
The result of this is that capacitive and inductive circuit elements tend to cancel
each other out.

Reactive power flow

In power transmission and distribution, significant effort is made to control the


reactive power flow. This is typically done automatically by switching inductors or
capacitor banks in and out, by adjusting generator excitation, and by other means.
Electricity retailers may use electricity meters which measure reactive power to
financially penalize customers with low power factor loads. This is particularly
relevant to customers operating highly inductive loads such as motors at water
pumping stations.

Unbalanced polyphase systems

While real power and reactive power are well defined in any system, the definition
of apparent power for unbalanced polyphase systems is considered to be one of the
most controversial topics in power engineering. Originally, apparent power arose
merely as a figure of merit. Major delineations of the concept are attributed to
Stanley's Phenomena of Retardation in the Induction Coil (1888) and Steinmetz's
Theoretical Elements of Engineering (1915). However, with the development of
three phase power distribution, it became clear that the definition of apparent
power and the power factor could not be applied to unbalanced polyphase systems.
In 1920, a "Special Joint Committee of the AIEE and the National Electric Light
Association met to resolve the issue. They considered two definitions:

that is, the quotient of the sums of the real powers for each phase over the sum of
the apparent power for each phase.

that is, the quotient of the sums of the real powers for each phase over the
magnitude of the sum of the complex powers for each phase.
The 1920 committee found no consensus and the topic continued to dominate
discussions. In 1930 another committee formed and once again failed to resolve the
question. The transcripts of their discussions are the lengthiest and most
controversial ever published by the AIEE (Emanuel, 1993). Further resolution of
this debate did not come until the late 1990s.

Basic calculations using real numbers

A perfect resistor stores no energy, so current and voltage are in phase. Therefore
there is no reactive power and P = S. Therefore for a perfect resistor:

For a perfect capacitor or inductor on the other hand there is no net power transfer,
so all power is reactive. Therefore for a perfect capacitor or inductor:

Where X is the reactance of the capacitor or inductor.

If X is defined as being positive for an inductor and negative for a capacitor then
we can remove the modulus signs from Q and X and get.

Multiple frequency systems

Since an RMS value can be calculated for any waveform, apparent power can be
calculated from this.

For real power it would at first appear that we would have to calculate loads of
product terms and average all of them. However if we look at one of these product
terms in more detail we come to a very interesting result.
however the time average of a function of the form cos(ωt + k) is zero provided
that ω is nonzero. Therefore the only product terms that have a nonzero average are
those where the frequency of voltage and current match. In other words it is
possible to calculate real (average) power by simply treating each frequency
separately and adding up the answers.

Furthermore, if we assume the voltage of the mains supply is a single frequency


(which it usually is), this shows that harmonic currents are a bad thing. They will
increase the rms current (since there will be non-zero terms added) and therefore
apparent power, but they will have no effect on the real power transferred. Hence,
harmonic currents will reduce the power factor.

Harmonic currents can be reduced by a filter placed at the input of the device.
Typically this will consist of either just a capacitor (relying on parasitic resistance
and inductance in the supply) or a capacitor-inductor network. An active power
factor correction circuit at the input would generally reduce the harmonic currents
further and maintain the power factor closer to unity. And hence, the different
components of current is defined.
Anyone who has heard a refrigeration unit groan through a brownout or been
frustrated by an inexplicable series of control system trips should have a healthy
respect for the effects of power quality on equipment reliability and energy
efficiency. Few doubt that sags and harmonics contribute to downtime, off-spec
production and shortened equipment life.

But, like so many potential projects that improve uptime, product quality and life
cycle cost, it’s hard to make a financial case for spending on power conditioning.
These days, it’s much easier to gain support and financial backing for the rapid
paybacks you can calculate from improving energy efficiency.

“Power-quality solutions have historically been sold as insurance policies to


protect against the next damaging sag, surge or interruption,” says Daniel
Carnovale, P.E., power quality solutions manager, Eaton (www.eaton.com).
“Those benefits are hard to justify. By contrast, energy savings offers easy
justification.”

The problem is that some of the solution providers significantly overstate the
savings and customers are deceived. Carnovale says, “While these solutions
provide excellent protection, they often provide very little, if any, energy savings.”

Common sense as well as the Electric Power Research Institute (EPRI) point out
that you can only save energy that is wasted. The losses in a typical plant power
system range from 1% to 4%, so eliminating them would, at best, reduce your bill
by 1% to 4%.

But these days, even a couple percent of energy savings can add up to a worthy
return on investment, and paying attention to power quality in the form of power
factor and peak demand can reduce some industrial power bills as much as 30%.

Sources of savings

Sags, surges, interruptions and harmonics wreak havoc on equipment. It’s easy to
believe they contribute to inefficiencies, and they do. But how much?

You might find low-hanging fruit in your existing power quality or uninterruptible
power supply (UPS) equipment. Strides have been made in the efficiency of the
rectifier stages that condition incoming power and connect capacitors or batteries
to support sags and outages. Just a few years ago, “Typical losses were 10% to
15%, and worse at lighter loads,” says Ed Spears, product marketing manager,
critical power solutions, Eaton. “Now we are in the range of 92% to 98%, even
99% efficient. This from better semiconductor technology but also topology. We
can operate in modes where we can allow more straight utility power to come
through.”

From there, savings directly from reducing inefficiencies drop off rapidly. For
example, in motors, “A 3% voltage imbalance can lead to a 25% to 30% current
imbalance, and increase motor losses by 10% to 20%. It also raises motor
operating temperatures,” says Mike Melfi, manager, advanced variable-speed
motor technologies, Baldor (www.baldor.com). But the increase in loss might not
add up to much. “Altogether, a 95% efficient motor might be reduced to 94%,”
Melfi says.

Harmonics are multiples of the fundamental frequency


(60 Hz in North America), and are produced by
nonlinear loads (Figure 1). “Unwanted harmonics
equal heat, which equals loss,” says John Perry,
product marketing manager, power factor correction
products, Eaton. “Wiring sizes, conduits and other
Figure 1. Harmonics are components can be minimized by minimizing
multiples of the harmonics.”
fundamental frequency (60
Ironically, some energy-saving technologies such as
Hz in North America), and
electronic lighting ballasts and variable-speed drives
are produced by nonlinear (VSD) can add harmonics, as can computer power
loads. supplies. These cause inefficiencies and shorten the
lives of motors, transformers and other connected
equipment. Even when the phase loads are balanced, “The third harmonics can add
up in the neutral conductor,” says Rich Vesel, product manager, electrical balance-
of-plant equipment, ABB. (www.ABB.com). “At 60 Hz, the neutral currents of a
balanced three-phase system cancel, but the third harmonics don’t cancel — they
add. An undersized neutral can run hot and might even become a fire hazard.
Balance the phases, oversize the neutral line, and use harmonic filtering to
minimize the effects.”

Furthermore, some of the ways plants deal with harmonics add to energy
inefficiencies. “People used to throw in K-rated transformers all over the place
because harmonics were eating up their standard transformers, but they would lose
2% to 4% efficiency,” says Carnovale. “Newer versions are much more efficient,
and they handle the harmonics — they can absorb them like a capacitor set. You
can use them in pairs to cancel out harmonics.”
The upshot is, while cleaning up power might pay off in higher system efficiency,
that’s not where you’ll find 25% savings. “Companies want to install a black box
and save a lot of energy, but unless they have a power factor situation and the
utility is imposing a penalty, they’re not going to save 20% to 30%,” Carnovale
says. “If they replace an inefficient UPS, they might pick up 10%. For harmonics,
it’s going to be more like 2% to 4%.”

Page 1 of 4

« Prev 1 | 2 | 3 | 4 Next » View on one page

Understand power factor

Finding the proverbial 20% to 30% reduction in electricity


bills doesn’t require an electrical engineering degree, though
you might need to consult with a few to obtain your goal. So
it’s worth having at least a rudimentary understanding of
active power (also called true power), reactive power,
apparent power and power factor.

Active or true power (P) is measured in watts (W) delivered


into a resistive load and converted into mechanical work or
useful heat.
Figure 2. The active
power component Reactive power (Q), measured in volt-amperes reactive
(P) is in phase with (VAR), is the power that energizes magnetic fields of
the applied voltage inductors (coils) and charges capacitors. Motors,
while the reactive transformers and capacitors require VARs. Creating and
component (Q) is 90º collapsing the magnetic fields occurs within each half cycle.
“Power engineers speak of motor loads as consumers of
out of phase. The VARs and capacitors as suppliers of VARs,” says Harry
power factor (PF) is Boulos, sales manager, Power Quality Group, Schneider
defined as the cosine Electric (www.reactivar.com).
of the angle between
active power and “Reactive power doesn’t perform measurable work, but it
loads the power-generating and transmission equipment the
apparent power (S).
same as if it did,” Vesel says. “So, the power company has
From the power to support both types of load.”
company’s
perspective, the Apparent power (S), measured in volt-amperes (VA) or
closer the power kilovolt-amperes (kVA), is the vector sum of the active and
factor is to unity, the
higher the efficiency
of the electrical
system.
the reactive components (Figure 2), and represents the system capacity required to
support the active plus reactive loads. Power transmission lines and transformers
must be sized in kVA to supply the total VARs and watts.

Power factor (PF) is the ratio of active to apparent power (W/VA).

The utility is paid by the kilowatt-hour (kWh) for real power, and charges a fee
based on the peak demand represented by the combination of real and reactive
power. “If a plant has a large reactive load because of poor power factor control,
along with the price of the power, you’re paying rent on an oversized ‘pipe’ to
deliver it,” Vesel says.

The demand fee might not be explicit. “Sometimes it’s obvious, sometimes not,”
Perry says. “If your bill includes terms like kVA or kVAR, they’re measuring your
power factor and they might be charging for it with a surcharge that shows up on
your bill.”

Power factor penalties take a variety of forms, calculated in different ways that are
detailed in the white paper “Power Quality Solutions and Energy Savings — What
Is Real?” at www.eaton.com/experience. Some utilities use multiple methods and
might apply more than one penalty.

A power factor penalty might be invisible. “For example, in a straight kVA


demand rate, there is nothing that explicitly mentions a power factor penalty,”
Carnavale says, “but a poor power factor will result in a higher kVA for a given
kW of load, so there is an implicit power factor penalty built into that rate. In other
cases, the utility might give a rebate for maintaining a power factor above a given
level. You might not realize you are paying a penalty if you don’t get this rebate.”

Customers of intermediaries such as municipalities or co-ops might find that


penalties are simply buried in the demand charge. “The power factor penalty might
be levied by the power supplier to the co-op and the penalty adjusted in the
demand charge,” Carnovale adds. “It might look like a reduction in kWh when the
penalty is removed.”

Be proactive about reactive

Your power factor penalty is typically determined by utility measurements of


average reactive power during 15-minute or 30-minute periods throughout your
billing cycle. You can reduce or eliminate the charge by balancing loads or using
power factor correction equipment. But, you say, this isn’t news — didn’t you take
care of that back in the 1980s with a set of tapped capacitors?

Maybe you did, but those were simpler times. “Loads have changed
tremendously,” Boulos says. “VSDs, presses, welders and automation together
cause problems on the network. With cyclic loads, traditional correction with a
fixed capacitor, or even with a controller averaging every 30 seconds or every
minute, doesn’t always work.”

VSDs offer higher system efficiency but can aggravate plant power quality and
reduce the reliability and efficiency of fixed-speed and constant-torque motors on
the same power supply. “The net energy savings can be reduced by these effects,”
says John Malinowski, senior product manager, AC and DC motors, Baldor. “The
drive design can help or hurt, and this is true of any load. Hence IEEE Standard
519, which is driven by reliability considerations but applying it also affects energy
efficiency.”

Page 3 of 4

« Prev 1 | 2 | 3 | 4 Next » View on one page

So one strategy is to be selective about the technologies you install. VSDs with
active rectifier units (ARUs) compensate for other inductive
loads on the same power supply or bus — for example, in a
group of motors on a bus where one or more run at full power
and speed, a motor with an ARU can compensate the
inductive loads of the full-power motors and correct the bus
power factor. “It’s best to consider only drives with active
front ends, where you can control reactive power and power
factor,” Malinowski says.

If today’s drives aren’t enough, separate power factor


correction devices can improve overall electrical efficiency
Figure 3. In this upstream of their point of connection in the electrical network
example, supplying (Figure 3). “The most basic example is a set of capacitors you
the reactive power switch in or out manually depending on the equipment that is
locally with a running,” says Vesel. “Static VAR compensation (SVC), or
STATCOM, might be applied at the substation.”
capacitor brings
the power factor to However, you can’t indiscriminately add power factor
unity and reduces correction capacitors without understanding how their
the power
supplier’s apparent
load from 100 kVA
to 80 kVA.
presence will affect the system. VFDs, DC power supplies, electronic ballasts and
other nonlinear loads can cause harmonics, which can interact with the inductive
reactance to cause harmonic resonance.

“Harmonic resonance is said to be a self-correcting problem,” says Carnovale.


“Most times capacitor fuses will open, capacitor cans will fail, or the source
transformer fails. Any of these events will lead to the removal of a component
from the system, eliminating the resonance condition. However, they’re all
undesirable results. In a best-case scenario, the electrical control equipment acts
erratically.”

The interactions call for combining corrections for harmonics and power factor.
For example, an active harmonic filter (AHF) can be part of a power factor
correction system (Figure 4). Some potential benefits of a holistic approach
include:

 Additional production without buying transformers and distribution


equipment by reducing the power factor.
 Ensuring compliance with the utility’s harmonic standards.
 Improving reliability of the distribution network and process equipment, for
example by eliminating logic faults of digital devices.
 Reducing overheating of transformers, motors and cables.

Keeping up with loads that vary throughout the production


cycle might require faster, automated equipment, like a
hybrid VAR compensator. It has a passive component that
works like a fixed capacitor, and a high-speed active
component. It can act as needed as a leading, lagging or
inductive load, and respond quickly to support voltage.
“They can save more energy than folks think they will,”
Figure 4. An active says Boulos. “Mars is saving an average of 5% at multiple
harmonic filter (AHF) facilities, where you’d traditionally expect a savings of 2%
senses dynamic power- to 3%.”
quality problems and
The potential savings depends on the kind of equipment.
injects the appropriate
An automaker would expect more savings with welding
harmonic and reactive and stamping equipment than in painting or assembly. In
current to limit the steel industry, it’s arc furnaces. Each part of the plant
harmonic distortion is a little different.
and improve power
factor. The result is a
nearly perfect sine
wave.
“It’s best to have someone visit and see what the problems are,” Boulos says. “Are
they sags or transients? We look at the loads and, if needed, set up a study and take
measurements. Then we can make recommendations and a proposal.” Paybacks
typically are one to two years for a normal capacitor system, two to three years for
a filtered system. “But reduced downtime and product savings by reducing scrap
can make it less than a year, sometimes less than six months,” Boulos adds.

Know where it’s going

The first step toward improving energy efficiency is to know where your energy is
going, and suppliers are supporting this measure by rapidly improving the
capabilities and reducing the costs of power monitoring systems.

Monitoring systems are being used to look at power quality, sequence of events,
trends in loads, temperatures and currents, and more. “We have a million boxes —
relays, meters, current transformers — to go into any power system,” says Vince
Tullo, general manager, sales, GE Digital Energy (www.gedigitalenergy.com).
“From inputs to motor control centers — substations, switchgear, protective relays
— you have, at every point, the ability to put on intelligence. You can have
control, metering and fault detection all in one box.

Page 3 of 4

« Prev 1 | 2 | 3 | 4 Next » View on one page

Page 4 of 4

« Prev 1 | 2 | 3 | 4 Next » View on one page

“But it’s not the boxes, it’s the ability to take that data, turn it into information and
give it to the right person in an actionable format,” Tullo continues. “The news
during the past few years is the ability to get information down to the process level
and be able to use it.”

To make educated decisions about corrective equipment for mission critical


operations, Eaton says a monitoring system should be capable of:

 Capturing triggered events like sags, swells, and interruptions (many energy
meters only captured time triggered data not threshold triggered data)
 Evaluating real-time and snapshot data for key electrical and power quality
variables
 Having sufficient frequency response and sampling rate to capture high-
speed transients such as events related to lighting or switching
 Trending appropriate variables so a change or variation from normal is
obvious and can be flagged and used for analysis.
 Trending harmonic currents and voltage distortion to evaluate issues related
to harmonics (overheating, nuisance operation or capacitor failures) 
 Time/date stamping events to within one millisecond to help determine the
correlation between cause/effect

“Electric power distribution units (EPDUs) can monitor and control, and be set to
alarm on overload or unbalance, to help maintain efficiency and prevent failures,”
says Perry. “By monitoring breakers and submetering components, you can catch
negative trends to save energy and downtime.”

A well-instrumented and characterized power system can be modeled for more


effective monitoring, engineering and maintenance. “Users can simulate,
understand, and re-engineer the way their entire electrical infrastructure will
respond in the event of unexpected faults or routine system maintenance,” says
says Mark Ascolese, CEO, EDSA (www.edsa.com). “You can detect variations
between as-designed specifications and actual electrical system performance so
you can predict potential points of failure and ensure that facilities operate as they
were designed to, down to the smallest component.”

A breaker removal or other maintenance on mission-critical equipment might


typically have to be done at night or over a weekend or holiday to allow for
surprises. “With simulation capabilities, you can see the effects, practice the
operation and do it anytime without worrying about unknowns,” Ascolese says.
“Modeling lets you see how changes would affect energy bills — whether doing
something different would save energy and cost. You can look at shifting assets
from peak to off-peak times and see what the effects would be.”

Shave a peak

With harmonics under control and a plan to keep power factors close to unity,
you’re ready to work with your utility to reduce peak loads and shift non-critical
power consumption to off-peak times. Your power company might offer to help
pay for it.

Sags can be filled and peaks can be shaved with energy storage devices. Flywheel
systems are used on material handling cranes, connected to the DC section of a
VSD. “During lift, power usage peaks and the DC bus can sag, but during let-
down, there’s energy available for regeneration,” says Frank DeLattre, chief sales
officer, Vycon (www.vycon.com). “That energy can be stored and used during lifts
to reduce consumption.” The devices also provide power for a short time during
outages.

Energy storage also can be used to cover peak loads where capacity is limited, such
as at the end of a heavily loaded power transmission line or a location where the
substation has become marginal. “You can draw more than the utility can provide
during short peaks, and use the flywheel to prevent a voltage sag,” DeLattre says.

“Clients like Valero Oil, BASF and Publix are putting in load-shedding systems to
reduce peak loads and handle outages,” Tullo says.

When hurricanes take out the power lines, Publix food stores are shut down — no
lights, no registers, and soon, the food spoils. So they put in standby power
systems with generators. “Now, when a hurricane comes through, they stay in
business, but more importantly, they also manage their peak demand,” Tullo says.
“On hot summer days, they run the generators and sell excess capacity back to
Florida Power and Light.”

Typical paybacks are 12 months to 18 months for load-shedding and critical power
initiatives, Tullo says. “They’re spending money to save money.”

Utilities often have formal demand-side reduction programs where they invest as
much to reduce demand as they would have to spend to increase capacity. “Some
have called them ‘dollar-a-watt’ programs,” Vesel says. “Talk with your utility.
Get your plant audited by people who know how to do it right and can give you a
well-analyzed and well-engineered result, and go from there.”

E-mail Editor in Chief Paul Studebaker, CMRP, at [email protected].

Page 4 of 4

« Prev 1 | 2 | 3 | 4 Next » View on one page

The Smith Chart, invented by Phillip H. Smith (1905-1987 is a graphical aid or nomogram
designed for electrical engineering specialising in radio frequency (RF) engineering to assist in
solving problems with transmission lines and impedance matching circuits.

Use of the Smith Chart utility has grown steadily over the years and it is still widely used today,
not only as a problem solving aid, but as a graphical demonstrator of how many RF parameters
behave at one or more frequencies, an alternative to using tabular information. The Smith Chart
can be used to represent many parameters including impedance, admittances, reflection
coefficients, scattering parameters, noise figure circles, constant gain contours and regions for
[[stability theory|unconditional stability

Contents
[hide]

 1 Overview
 2 Mathematical Basis
o 2.1 Actual and Normalised Impedance and Admittance
o 2.2 The Normalised Impedance Smith Chart
 2.2.1 The Variation of Complex Reflection Coefficient with Position Along the
Line
 2.2.2 The Variation of Normalised Impedance with Position Along the Line
 2.2.3 Regions of the Z Smith Chart
 2.2.4 Circles of Constant Normalised Resistance and Constant Normalised
Reactance
o 2.3 The Y Smith Chart
o 2.4 Practical Examples
o 2.5 Working with Both the Z Smith Chart and the Y Smith Charts
o 2.6 Choice of Smith Chart Type and Component Type
 3 Using the Smith Chart to Solve Conjugate Matching Problems With Distributed Components
 4 Using the Smith Chart to Analyse Lumped Element Circuits
 5 References
 6 External links

[edit] Overview
The Smith Chart is plotted on the complex number and reflection coefficient plane in two
dimension and is scaled in normalised impedance (the most common), normalised admittance or
both, using different colours to distuinguish between them. These are often known as the Z, Y
and YZ Smith Charts respectively

Normalised scaling allows the Smith Chart to be used for problems involving any characteristic
impedance or system impedance, although by far the most commonly used is 50 Ohms. With
relatively simple graphical construction it is straighforward to convert between normalised
impedance (or normalised admittance) and the corresponding complex voltage reflection
coefficient.

The Smith Chart has circumferential scaling in wavelengths and degrees. The wavelengths scale
is used in distributed component problems and represents the distance measured along the
transmission line connected between the signal generator or source and the load to the point
under consideration. The degrees scale represents the angle of the voltage reflection coefficient
at that point. The Smith Chart may also be used for lumped element model matching and
analysis problems.

Use of the Smith Chart and the interpretation of the results obtained using it requires a good
understanding of alternating current and transmission line theory, both of which are pre-
requisites for RF engineers.

As impedances and admittances change with frequency, problems using the Smith Chart can
only be solved manually using one frequency at a time, the result being represented by a
point(geometry). This is often adequate for narrow band applications (typically up to about 5%
to 10% bandwidth) but for wider bandwidths it is usually necessary to apply Smith Chart
techniques at more than one frequency across the operating frequency band. Provided the
frequencies are sufficiently close, the resulting Smith Chart points may be joined by straight
lines to create a locus.

A locus of points on a Smith Chart covering a range of frequencies can be used to visually
represent:

 how capacitance or how inductive a load is across the frequency range


 how difficult matching is likely to be at various frequencies
 how well matched a particular component is.

The accuracy of the Smith Chart is reduced for problems involving a large statistics spread of
impedances or admittances, although the scaling can be magnified for individual areas to
accommodate these.

[edit] Mathematical Basis


[edit] Actual and Normalised Impedance and Admittance

A transmission line with a characteristic impedance of may be universally considered to have


a characteristic admittance of where

Any impedance, expressed in Ohms, may be normalised by dividing it by the characteristic


impedance, so the normalised impedance using the lower case z, suffix T is given by

Similarly, for normalised admittance


The SI unit of impedance is the Ohm with the symbol of the upper case Greek letter Omega ( )
and the SI unit for admittance is the Siemens with the symbol of an upper case letter S.
Normalised impedance and normalised admittance have no units. Actual impedances and
admittances must be normalised before using them on a Smith Chart. Once the result is obtained
it may be de-normalised to obtain the actual result.

[edit] The Normalised Impedance Smith Chart

Using transmission line theory, if a transmission line is terminated in an impedance ( ) which


differs from its characteristic impedance ( ), a standing wave will be formed on the line
comprising the resultant of both the forward ( ) and the reflected ( ) waves. Using
complex exponential notation:

and

where

is the temporal part of the wave and

where

is the angular frequency in radians per second (rad/s)

is the frequency in Hertz (Hz)

is the time in seconds (s)

and are constants

is the distance measured along the transmission line from the generator in metres (m)

Also

is the propagation constant which has units 1/m

where

is the attenuation constant in Nepers per metre (Np/m)

is the phase constant in radians per metre (rad/m)


The Smith Chart is used with one frequency at a time so the temporal part of the phase (
) is fixed. All terms are actually multiplied by this to obtain the instantaneous phase,
but it is conventional and understood to omit it. Therefore

and

[edit] The Variation of Complex Reflection Coefficient with Position Along the Line

The complex voltage reflection coefficient is defined as the ratio of the reflected wave to the
incident (or forward) wave. Therefore

where C is also a constant. For a uniform transmission line (in which is constant), the complex
reflection coefficient of a standing wave varies according to the position on the line. If the line is
lossy ( is finite) this is represented on the Smith Chart by a spiral path. In most Smith Chart
problems however, losses can be assumed negligible ( ) and the task of solving them is
greatly simplified. For the loss free case therefore, the expression for complex reflection
coefficient becomes

The phase constant may also be written as

where is the wavelength within the transmission line at the test frequency. Therefore

This equation shows that, for a standing wave, the complex reflection coefficient and impedance
repeats every half wavelength along the transmission line. The complex reflection coefficient is
generally simply referred to as reflection coefficient. The outer circumferential scale of the Smith
Chart represents the distance from the generator to the load scaled in wavelengths and is
therefore scaled from zero to 0.50.

[edit] The Variation of Normalised Impedance with Position Along the Line

If and are the voltage across and the current entering the termination at the end of the
transmission line respectively, then
and

By dividing these equations and substituting for both the voltage reflection coefficient

and the normalised impedance of the termination represented by the lower case Z, subscript T

gives the result:

Alternatively, in terms of the reflection coefficient

These are the equations which are used to construct the Z Smith Chart.

Both and are expressed in complex numbers without any units. They both change with
frequency so for any particular measurement, the frequency at which it was performed must be
stated together with the characteristic impedance.

may be expressed in magnitude and angle on a polar diagram. Any actual reflection coefficient
must have a magnitude of less than or equal to unity so, at the test frequency, this may be
expressed by a point inside a circle of unity radius. The Smith Chart is actually constructed on
such a polar diagram. The Smith chart scaling is designed in such a way that reflection
coefficient can be converted to normalised impedance or vice versa. Using the Smith Chart, the
normalised impedance may be obtained with appreciable accuracy by plotting the point
representing the reflection coefficient treating the Smith Chart as a polar diagram and then
reading its value directly using the characteristic Smith Chart scaling. This technique is a
graphical alternative to substituting the values in the equations.

By substituting the expression for how reflection coefficient changes along an unmatched loss
free transmission line
for the loss free case, into the equation for normalised impedance in terms of reflection
coefficient

and using Euler's identity

yields the impedance version transmission line equation for the loss free case:<ref>Hayt,
William H Jr.; "Engineering Electromagnetics" Fourth Ed;McGraw-Hill International Book
Company; pp 428 433. IBSN 0-07-027395-2.</ref>

where is the impedance 'seen' at the input of a loss free transmission line of length l,
terminated with an impedance

Versions of the transmission line equation may be similarly derived for the admittance loss free
case and for the impedance and admittance lossy cases.

The Smith Chart graphical equivalent of using the transmission line equation is to normalise
, to plot the resulting point on a Z Smith Chart and to draw a circle through that point
centred at the Smith Chart centre. The path along the arc of the circle represents how the
impedance changes whilst moving along the transmission line. In this case the circumferential
(wavelength) scaling must be used, remembering that this is the wavelength within the
transmission line and may differ from the free space wavelength.

[edit] Regions of the Z Smith Chart

If a polar diagram is mapped on to a cartesian coordinate system it is conventional to measure


angles relative to the positive x-axis using a counter-clockwise direction for positive angles. The
magnitude of a complex number is the length of a straight line drawn from the origin to the point
representing it. The Smith Chart uses the same convention, noting that, in the normalised
impedance plane, the positive x-axis extends from the center of the Smith Chart at
to the point . The region above the x-axis represents inductive impedances and
the region below the x-axis represents capacitive impedances. Inductive impedances have
positive imaginary parts and capacitive impedances have negative imaginary parts.

If the termination is perfectly matched, the reflection coefficient will be zero, represented
effectively by a circle of zero radius or in fact a point at the centre of the Smith Chart. If the
termination was a perfect open circuit or short circuit the magnitude of the reflection coefficient
would be unity, all power would be reflected and the point would lie at some point on the unity
circumference circle. Given that Smith's chart isn't accurate enough, Rafay's chart can be used to
find the inverse impedance.

[edit] Circles of Constant Normalised Resistance and Constant Normalised Reactance

The normalised impedance Smith Chart is composed of two families of circles: circles of
constant normalised resistance and circles of constant normalised reactance. In the complex
reflection coefficient plane the Smith Chart occupies a circle of unity radius centred at the origin.
In cartesian coordinates therefore the circle would pass through the points (1,0) and (-1,0) on the
x-axis and the points (0,1) and (0,-1) on the y-axis.

Since both and are complex numbers, in general they may be expressed by the following
generic rectangular complex numbers:

Substituting these into the equation relating normalised impedance and complex reflection
coefficient:

gives the following result:

This is the equation which describes how the complex reflection coefficient changes with the
normalised impedance and may be used to construct both families of circles.<ref>Davidson, C.
W.;"Transmission Lines for Communications with CAD Programs";Macmillan; pp 80-85. ISBN
0-333-47398-1</ref>

[edit] The Y Smith Chart

The Y Smith chart is constructed in a similar way to the Z Smith Chart case but by expressing
values of voltage reflection coefficient in terms of normalised admittance instead of normalised
impedance. The normalised admittance yT is the reciprocal of the normalised impedance zT, so

Therefore:
and

The Y Smith Chart appears like the normalised impedance type but with the graphic scaling
rotated through , the numeric scaling remaining unchanged.

The region above the x-axis represents capacitive admittances and the region below the x-axis
represents inductive admittances. Capacitive admittances have positive imaginary parts and
inductive admittances have negative imaginary parts.

Again, if the termination is perfectly matched the reflection coefficient will be zero, represented
by a 'circle' of zero radius or in fact a point at the centre of the Smith Chart. If the termination
was a perfect open or short circuit the magnitude of the voltage reflection coefficient would be
unity, all power would be reflected and the point would lie at some point on the unity
circumference circle of the Smith Chart.

[edit] Practical Examples

Example points plotted on the normalised impedance Smith Chart

A point with a reflection coefficient magnitude 0.63 and angle , represented in polar form as
, is shown as point P1 on the Smith Chart. To plot this, one may use the
circumferential (reflection coefficient) angle scale to find the graduation and a ruler to
draw a line passing through this and the centre of the Smith Chart. The length of the line would
then be scaled to P1 assuming the Smith Chart radius to be unity. For example if the actual radius
measured from the paper was 100 mm, the length OP1 would be 63 mm.
The following table gives some similar examples of points which are plotted on the Z Smith
Chart. For each, the reflection coefficient is given in polar form together with the corresponding
normalised impedance in rectangular form. The conversion may be read directly from the Smith
Chart or by substitution into the equation.

Some examples of points plotted on the normalised impedance Smith Chart

Reflection Coefficient (Polar Normalised Impedance


Point Identity
Form) (Rectangular Form)

P1 (Inductive)

P2 (Inductive)

P3 (Capacitive)

[edit] Working with Both the Z Smith Chart and the Y Smith Charts

In RF circuit and matching problems sometimes it is more convenient to work with admittances
(representing conductances and susceptances) and sometimes it is more convenient to work with
impedances (representing resistances and reactances). Solving a typical matching problem will
often require several changes between both types of Smith Chart, using normalised impedance
for series elements and normalised admittances for parallel elements. For these a dual
(normalised) impedance and admittance Smith Chart may be used. Alternatively, one type may
be used and the scaling converted to the other when required. In order to change from normalised
impedance to normalised admittance or vice versa, the point representing the value of reflection
coefficient under consideration is moved through exactly 180 degrees at the same radius. For
example the point P1 in the example representing a reflection coefficient of has a
normalised impedance of . To graphically change this to the equivalent
normalised admittance point, say Q1, a line is drawn with a ruler from P1 through the Smith
Chart centre to Q1, an equal radius in the opposite direction. This is equivalent to moving the
point through a circular path of exactly 180 degrees. Reading the value from the Smith Chart for
Q1, remembering that the scaling is now in normalised admittance, gives .
Performing the calculation

manually will confirm this.

Once a transformation from impedance to admittance has been performed the scaling changes to
normalised admittance until such time that a later transformation back to normalised impedance
is performed.
The table below shows examples of normalised impedances and their equivalent normalised
admittances obtained by rotation of the point through . Again these may either be obtained
by calculation or using a Smith Chart as shown, converting between the normalised impedance
and normalised admittances planes.

Values of reflection coefficient as normalised impedances and the equivalent normalised admittances

Normalised Impedance Plane Normalised Admittance Plane

P1 ( ) Q1 ( )

P10 ( ) Q10 ( )

Values of complex reflection coefficient plotted on the normalised impedance Smith Chart and their
equivalents on the normalised admittance Smith Chart

[edit] Choice of Smith Chart Type and Component Type

The choice of whether to use the Z Smith Chart or the Y Smith Chart for any particular
calculation depends on which is more convenient. Impedances in series and admittances in
parallel add whilst impedances in parallel and admittances in series are related by a reciprocal
equation. If is the equivalent impedance of series impedances and is the equivalent
impedance of parallel impedances, then

For admittances the reverse is true, that is


Dealing with the reciprocals, especially in complex numbers, is more time consuming and error-
prone than using linear addition. In general therefore, most RF engineers work in the plane
where the circuit topography supports linear addition. The following table gives the complex
expressions for impedance (real and normalised) and admittance (real and normalised) for each
of the three basic passive circuit elements: resistance, inductance and capacitance. Knowing just
the characteristic impedance (or characteristic admittance) and test frequency can be used to find
the equivalent circuit from any impedance or admittance, or vice versa.

Expressions for Real and Normalised Impedance and Admittance with Characteristic Impedance Z 0 or
Characteristic Admittance Y0

Impedance (Z or z) or Reactance (X or x)
Element Type
Real ( ) Normalised (No Unit)

Resistance (R)

Inductance (L)

Capacitance (C)

Admittance (Y or y) or Susceptance (B or b)
Element Type
Real (S) Normalised (No Unit)

Resistance (R)

Inductance (L)

Capacitance (C)

[edit] Using the Smith Chart to Solve Conjugate Matching


Problems With Distributed Components
Usually distributed matching is only feasable at microwave frequencies since, for most
components operating at these frequencies, appreciable transmission line dimensions are
available in terms of wavelengths. Also the electrical behavior of many lumped components
becomes rather unpredictable at these frequencies.

For distributed components the effects on reflection coefficient and impedance of moving along
the transmission line must be allowed for using the outer circumferential scale of the Smith Chart
which is calibrated in wavelengths.

The following example shows how a transmission line, terminated with an arbitrary load, may be
matched at one frequency either with a series or parallel reactive component in each case
connected at precise positions.

Smith Chart construction for some distributed transmission line matching

Supposing a loss free air-spaced transmission line of characteristic impedance ,


operating at a frequency of 800 MHz, is terminated with a circuit comprising a 17.5 resistor in
series with a 6.5 nanohenry (6.5 nH) inductor. How may the line be matched?

From the table above, the reactance of the inductor forming part of the termination at 800 MHz is

so the impedance of the combination ( ) is given by

and the normalised impedance ( )is

This is plotted on the Z Smith Chart at point P20. The line OP20 is extended through to the
wavelength scale where it intersects at the point . As the transmission line is loss
free, a circle centred at the centre of the Smith Chart is drawn through the point P20 to represent
the path of the constant magnitude reflection coefficient due to the termination. At point P21 the
circle intersects with the unity circle of constant normalised resistance at

The extension of the line OP21 intersects the wavelength scale at , therefore the
distance from the termination to this point on the line is given by

Since the transmission line is air-spaced, the wavelength at 800 MHz in the line is the same as
that in free space and is given by

where is the velocity of electromagnetic radiation in free space and is the frequency in hertz.
The result gives , making the position of the matching component 29.6 mm from
the load.

The conjugate match for the impedance at P21 ( ) is

As the Smith Chart is still in the normalised impedance plane, from the table above a series
capacitor is required where

Therefore

To match the termination at 800 MHz, a series capacitor of 2.6 pF must be placed in series with
the transmission line at a distance of 29.6 mm from the termination.

An alternative shunt match could be calculated after performing a Smith Chart transformation
from normalised impedance to normalised admittance. Point Q20 is the equivalent of P20 but
expressed as a normalised admittance. Reading from the Smith Chart scaling, remembering that
this is now a normalised admittance gives
(In fact this value is not actually used). However, the extension of the line OQ20 through to the
wavelength scale gives . The earliest point at which a shunt conjugate match
could be introduced,moving towards the generator, would be at Q21, the same position as the
previous P21, but this time representing a normalised admittance given by

The distance along the transmission line is in this case

which converts to 123 mm.

The conjugate matching component is required to have a normalised admittance ( ) of

From the table it can be seen that a negative admittance would require to be an inductor,
connected in parallel with the transmission line. If its value is , then

This gives the result

A suitable inductive shunt matching would therefore be a 6.5 nH inductor in parallel with the
line positioned at 123 mm from the load.

[edit] Using the Smith Chart to Analyse Lumped Element


Circuits
The analysis of lumped element components assumes that the wavelength at the frequency of
operation is much greater than the dimensions of the components themselves. The Smith Chart
may be used to analyse such circuits in which case the movements around the chart are generated
by the (normalised) impedances and admittances of the components at the frequency of
operation. In this case the wavelength scaling on the Smith Chart circumference is not used. The
following circuit will be analysed using a Smith Chart at an operating frequency of 100 MHz. At
this frequency the free space wavelength is 3 m. The component dimensions themselves will be
in the order of millimetres so the assumption of lumped components will be valid. Despite there
being no transmission line as such, a system impedance must still be defined to enable
normalisation and de-normalisation calculations and is a good choice here as
. If there were very different values of resistance present a value closer to these
might be a better choice.

A lumped element circuit which may be analysed using a Smith Chart

Smith Chart with graphical construction for analysis of a lumped circuit

The analysis starts with a Z Smith Chart looking into R1 only with no other components present.
As is the same as the system impedance, this is represented by a point at the centre
of the Smith Chart. The first transformation is OP1 along the line of constant normalised
resistance in this case the addition of a normalised reactance of -j0.80, corresponding to a series
capacitor of 40 pF. Points with suffix P are in the Z plane and points with suffix Q are in the Y
plane. Therefore transformations P1 to Q1 and P3 to Q3 are from the Z Smith Chart to the Y Smith
Chart and transformation Q2 to P2 is from the Y Smith Chart to the Z Smith Chart. The following
table shows the steps taken to work through the remaining components and transformations,
returning eventually back to the centre of the Smith Chart and a perfect 50 Ohm match.

Smith Chart steps for analysing a lumped element circuit

x or y
Transformation Plane Normalised Capacitance/Inductance Formula to Solve Result
Value

Capacitance (Series)
Inductance (Shunt)

Z Capacitance (Series)

Y Capacitance (Shunt)

Active Harmonic Filter - Digital Power Manager - DPM

Incomparable Features!

ELIMINATION OF HARMONICS actively reduces


harmonic distortion to less than 5%
INSTANTANEOUS REACTION instantly reacts to and
corrects varying load conditions 
EASY INSTALLATION simple, zero downtime
installation (Split CT option required)
UTILITY SAVINGS reduced utility costs
MTBF all Mesta equipment is designed to operate for at
least 200,000 hours
COMPENSATION FOR ALL REACTIVE
CURRENTS  
BALANCING OF THREE PHASE LOADS
EXTENSION OF PLANT EQUIPMENT LIFE

Digital Power Factor and Harmonic Manager

See how powerful the Mesta DPM is, click into the DPM Power Lab.

 
 

True Active Filter


The Mesta DPM is a true active filter, continuously monitoring the load
currents and responding to the amount and type of harmonics encountered. 
The system injects a corrective current, dramatically
Effectively reduces Harmonics 
The Mesta DPM reduces harmonic distortion to less than 5%,while balancing
the power and improving the power factor to the plant utility grid.   
Helps meet ANSI/IEEE-519-1992 Requirements 
When placed on a plant load, it helps meet the ANSI/IEEE-519-1992
Requirements, under all constant or variable load conditions. Mesta DPM
reduces Harmonic distortion to less than 5%, while balancing the power and
improving the power factor to the plant utility grid.   
Highly Sophisticated Microcontroller Intelligence  
The patented design of the MESTA DPM incorporates a powerful
microcontroller, DSP microcomputer, and the use of highly sophisticated
software.   
Sizing a DPM  
When sizing a DPM to your facility, you need only consider the harmonics and
fundamental phase shift.  The DPM Current rating indicates the amount of
correction current that the unit is capable of producing.  As an example, a 125
HP motor drive with 32% THD would have approximately 40 amps of
harmonic current.  If there is also 25 amps of reactive current, the filter will
generate a total of 48 amps of correction current.  Therefore, installing a 50
amp DPM corrects the harmonics and reactive current, even though the drive
draws 130 amps of total current.     

Figure 1. Mesta Power Lab


DPM Harmonics Analysis
B) DPM Correction
A) Utility Current Current C) Load Current
 
The Mesta DPM, placed in parallel with your non-linear load (as seen above),
continuously interacts with line in order to ensure that purely sinusoidal, balanced,
and current at unity power factor current is drawn from the utility.

Observing the Analysis in Figure 1, the power of the Mesta DPM is immediately
obvious.  The graphs (A, B, C) display data at three different points of the system:
(A) the targeted, purely sinusoidal utility current with unity power factor (B) the
necessary current harmonics in order for the Mesta DPM to achieve those results
and (C) the non-linear load with harmonics.     

By investigation of the line current, the Mesta DPM instantaneously finds and
corrects any non-fundamental current that is requested by a non-linear load.

To further demonstrate the achieved end results, Figure 2 summarizes the above
data into a table.  

Figure 2.

(A) Targeted (C) Non-Linear Load


 
Utility Current Current
Current (Amps) 41 50
Total Harmonic Distortion
1.6 62.1
THD(%)
Power Factor 0.99 0.64

ELECTRICAL DEMAND AND POWER FACTOR

     Many industrial electrical distribution systems are being used in ways not
foreseen by the designers.  These changes in use can cause some problems in
energy consumption and in safety. If a motor is operating at a lower voltage than it
was designed for, it is probably using more amperage than was intended and is
causing unnecessary losses in transmission lines.  If the wires are too small for the
load, line losses can be large, and fire hazards increase significantly.  Other
problems that can create unnecessary energy loss are voltage imbalance in three-
phase motors and leaks from voltage sources to ground.  Another problem that may
be costing money is a low power factor.  The optimal use of a plants' electrical use
can reduce operating and production costs.  The following module containing the
recommendations below illustrates this energy savings potential.

General Rules of Thumb:

 The average cost of electricity is $0.05/kWh ($15/MMBtu)


 There are 2000 hours per year per shift (based on the assumption that one
shift is 8 hours per day, 5 days per week, 50 weeks per year)
 Switching from electric heat to natural gas or #2 fuel oil can reduce
heating costs by 78%
 Average cost savings for demand reduction:  By shifting an operation to
off-peak hours, the following savings are achieved:  $75/Hp/year
 The average benefit of shifting other electrical equipment to off-peak
hours is: $120/kW/year

 
Notes:
                    Before choosing the following targeted recommendations READ
THE FOLLOWING:
    Pay back estimates for the following recommendations will use the
equation below.  They will vary depending on the, application, type of
installation, and purchase quantity of material and labor associated
with each recommendation.  It will be up to the person doing the
analyses to use the URL references below each equation to help
estimate an implementation cost.

    The data correlating to the variables below each equation will
be prompted for in order to execute a calculation.  Frequently the
fuel cost (FC) associated with the specific recommendation will be
prompted for in order to calculate the annual cost savings (ACS).
Unless otherwise specific to a particular recommendation the ACS
will calculated as follows:
 

1. Optimize plant power factors


2. Install demand controller/load shedder
3. Install power factor controllers on motors
4. Reduce transformer capacity
5. Check accuracy of power meter
6. Reduce rates
7. Take advantage of utility controlled power management for price
reduction
8. Reduce late fees
9. Install efficient rectifiers

1.  Optimize plant power factors


      The power factor is the ratio between the resistive component of AC
power and the total (resistive and inductive) power supplied.  The
installation of capacitors to improve power factor will reduce utility demand
charges. There will also be some additional savings from the reduced in-
plant power line losses (improving the power factor results in a
proportionally decreased current, and thus a reduction in the line power
losses, which are proportional to the square of the current). There will be no
energy savings from the reduction in utility demand charge. Since it costs
more to provide electricity with a low power factor it is common for
electrical utilities to make an additional charge if the power factor is less
than some value (typically 0.90).  Equipment that contributes to a low power
factor includes welding machines, induction motors, power transformers,
electric arc furnaces, and fluorescent light ballasts. The following equation
shows the utility savings that can occur.
 

Power Engineering Books


Thomas Register
NEMA

AUCS = annual utility cost savings, $


KVA1 = current average monthly billing demand, KW/month = KW/PFc
(average kilowatt demand/current power factor)
KVA2 = anticipated average monthly billing demand, KW/month =
KW/PFa (average kilowatt demand/anticipated power factor)
DC = demand charge, $/KW
MY = months per year demand is charged

Data Conversions?
RPN Calculator?

Press the calculate button to execute an estimation

ALCS = annual line cost savings, $


PFc = current power factor
PFa = anticipated power factor
LLP = line loss percentage (decision made by plant personnel)*
KWH = plant energy consumption per year, KWhr/yr (from utility bills)
AEC = yearly average energy charge, $/KWhr (from utility bills)

Data Conversions?
RPN Calculator?

Press the calculate button to execute an estimation

BACK TO FRONT PAGE

* A conservative estimate is an average of a 2.5% power loss in the plant


wiring, since accepted practice calls for allowing a maximum of a 5%
voltage drop to any machine in the plant.

2.  Install demand controller/load shedder

     Install demand controller or load shedder to limit peak demand. There
will be no energy savings since demand is shifted, not reduced.  The
following equation shows the savings that can be achieved.
 

Power Engineering Books


Thomas Register
NEMA

DS = demand savings, KW
DC = demand charge, $/KW
MY = months per year savings will occur

Data Conversions?
RPN Calculator?

Press the calculate button to execute an estimation

BACK TO FRONT PAGE


 
3.  Install power factor controllers on motors

    Power factor controllers when properly installed on AC motors will


reduce reactive currents produced by the quality of the AC power (ie how
much the current leads or lags the voltage supplied).  The power loss from
the reactive currents does no useful work and contributes to energy losses. 
The size of power factor controllers for T-frame motors may be higher than
those of U-frame motors of the same horsepower, speed and design.  
Typically the power factor for T-frame motors is less than that of U-frame
motors.  The following equation illustrates the potential savings that can be
derived when the installation of a power factor controller is considered.
 
 

Power Engineering Books


NEMA
Thomas Register

PFLKW = Present Full Load KW = HP x CON x 1/PME


HP = motor horsepower
CON = conversion factor, 0.746 KW/hp
PME = present full load motor efficiency (choose Motor efficiency table for
a value)

PIKW = Present Idling KW = PFLKW x LF


LF = load factor at idling

NFLKW = New Full Load KW = HP x CON x 1/NME


NME = new full load motor efficiency

NIKW = New Idling KW = PFLFW x NLF


NLF = new load factor at idling
FOH = operating hours at full load
IOH = operating hours of idling

Data Conversions?
RPN Calculator?
Press the calculate button to execute an estimation

BACK TO FRONT PAGE

4.  Reduce transformer capacity

     De-energize excess transformer capacity to avoid utility charges;


redistribute existing loads to permit removal of under loaded transformers. 
Note:  consider power loss as well as initial loads and growth in sizing
transformers.
    Transformers have substantial continuing no-load losses, related to the
primary side voltage, so they incur power losses on the basis of their full
load rating (ref.- South Carolina Energy Conservation Manual, Governor's
Division of Energy Agriculture and Natural Resources).  The following
equation illustrates the potential energy savings that can be achieved.

Power Engineering Books


Thomas Register
NEMA

RC = rated capacity, KW
NL = fractional loss with no load*
OT = off time for transformers, hrs/yr

* According to "Kent's Mechanical Engineers Handbook, 12th Edition", the


no load loss of transformers is between 0.3% of the rated capacity for small
power transformers (500 kVA and more) and 1% for small distribution
transformers (500 kVA and less).

Data Conversions?
RPN Calculator?

Press the calculate button to execute an estimation

BACK TO FRONT PAGE

5.  Check accuracy of power meter


     The accuracy of a power meter can be an underlying reason why a
particular trend of unknown energy loss has occurred in a manufacturing
plant  The following equations are used to determine the annual cost savings
with the installment of an accurate metering device.
 

 
Power Engineering Books
Thomas Register
NEMA

ED = excess demand due to meter error (KW) *


RU = rate per unit of electrical power, ($/KW)
*
User demand can be calculated from a watt-hour meter using the following
formula:

Kh = gearing ratio between disk rotation and movement of hands on the


meter
M = product of the ratio of current or potential transformers, if in use
R = number of disk revolutions measured in time T
T = time in seconds

Data Conversions?
RPN Calculator?

Press the calculate button to execute an estimation

BACK TO FRONT PAGE

6.  Reduce rates

    Reduce rates and/or restructure rate schedules or make other changes in
electric service in order to obtain lowest possible rates.  The following
equation illustrates the potential savings that can occur when rate schedule
changes are made to obtain the least possible rates.
 
Power Engineering Books
Thomas Register
 

DS = demand shifted (amount of KW saved)


DC = demand cost, $/KW (from electric bills or utility)
MY = months per year
LF = average load factor over operating hours
OH = operating hours shifted from on-peak to off-peak, per year (due to
demand shift)
EC = differential electric cost between on-peak and off-peak hours, $/KWhr

Data Conversions?
RPN Calculator?

Press the calculate button to execute an estimation

BACK TO FRONT PAGE


 

7.  Take advantage of utility controlled power management for price


reduction,

     Capitalize on interruptable Rate Schedules (non-critical loads placed


under control of utility in exchange for lower rates).  The following equation
illustrates the monetary savings that can be achieved.
 

Power Engineering Books


Thomas Register
 
TD = total demand affected, KW
DR = difference in rates, $/KW (contact utility)

Data Conversions?
RPN Calculator?

Press the calculate button to execute an estimation

BACK TO FRONT PAGE

8.  Reduce late fees


    Pay utility bills on schedule to avoid late fees. Annual cost savings will
equal the amount which would have been paid as late fees.  The following
equation provides an insight to the potential savings that can be achieved.
 

Power Engineering Books


Thomas Register
 

NB = original billed amount, $ (from utility bill)


PLC = percent charged as late fees (from utility bill)

Data Conversions?
RPN Calculator?

Press the calculate button to execute an estimation

BACK TO FRONT PAGE

9.  Install efficient rectifiers

    Install efficient rectifiers in place of existing rectification or DC


generating equipment  The following equation shows the savings that can be
achieved.
 
Power Engineering Books
Thomas Register
NEMA
 

HP = total power rating of connected motors, HP


CON = conversion factor, 0.746 KW/HP
HY = operating hours per year
hc = current average efficiency of all connected motors, (click Motor
efficiency table for a value)
ha = anticipated average efficiency

KVAR Power Factor Optimizer

All Residential Units have a 12 Year Warranty!


ALL Commercial Units have a 5 Year Warranty!
ALL PRODUCTS HAVE FREE SHIPPING!

 Product:  Desc. Residential  Wiring  Price


PFCD 1500 50amp/240v, single phase 3wire Rt/ L mount $350.00
 PFCD 1100  100amp/240v, single phase 3wire  Rt/ L mount  $399.00
 PFCD1200  200amp/240v, single phase 3 wire  Rt/L mount  $399.00
 PFCD 1400  400amp/240v, single phase 3 wire  Rt/L mount  $999.00
 Product:  Desc. Commercial,    
 PFCD3100  100amp/208,240,480v, 3phase 4 wire  Rt/L mount $499.00
 PFCD3200  200amp/208,240,480v, 3phase  4 wire  Rt/L mount $1099.00
 PFCD3400  400amp/208,240,480v, 3phase 4 wire  Rt/L mount $1999.00

** Note: When ordering please specify whether you need a right or left hand unit; The picture
    above denotes a left hand unit wiring on left side.
    All units come prewired with 3' of wire and flex for your convenience.
***Important: to assure you order the correct size Power Factor Unit, check your main breaker
or
     fuse size this will determine the size of your unit!
****LINK: TO VIEW WARRANTY'S AND SPECIFICATIONS

WE ACCEPT

 
 The PFCD 3100 is a 100amp 208/ 480v, three phase/ 4wire unit.
 This unit is fused and needs to be installed on a 3pole 30amp breaker closest to the main breaker.
 Lowers electric bills up to 20%.
 Reduces power consumption.
 Eliminates power surges and spikes.
 Enhances capacity of existing electrical system.
 Protects appliances and sensitive electronic equipment.
 Reduces harmful (EMFs) electro magnetic fields.
 Improves power factor.
 Reduces electricity required by inductive loads (motors).
 increases the life of appliances and motors.
 Certified Green Product, Reduces Carbon Footprint
 
 The PFCD 3200 is a 200amp 208/ 480v, three phase/ 4wire unit.
 This unit is fused and needs to be installed on a 3pole 30amp breaker closest to the main breaker.
 Lowers electric bills up to 20%.
 Reduces power consumption.
 Eliminates power surges and spikes.
 Enhances capacity of existing electrical system.
 Protects appliances and sensitive electronic equipment.
 Reduces harmful (EMFs) electro magnetic fields.
 Improves power factor.
 Reduces electricity required by inductive loads (motors).
 increases the life of appliances and motors.
 Certified Green Product, Reduces Carbon Footprint

March 
1996                                                                                                                                    
  Volume 2   Issue 5

IMPLEMENTING POWER FACTOR CORRECTION

    Correcting power factor in a production facility often reduces the electricity
bill by thousands of dollars a year.  That was the message in the last issue of
$mart Energy User.  This issue deals with actually doing it, with getting the
technical advice you need, selecting the best correction option for your plant, and
pricing out the job.

    Let’s assume that the Energy and Minerals Section has done an electrical energy
audit of your facility.  We came, took measurements, and found that your plant’s
peak electrical demand was accompanied by a low power factor.  We then
calculated what the annual electricity bill savings would be if the power factor was
corrected.  We may also have come up with a rough cost estimate, based on the
experience of others, and calculated an approximate payback figure.  You liked the
preliminary numbers, and would now like to proceed.  What’s the next step?
 

1.      SELECT THE RIGHT POWER FACTOR CORRECTION


APPROACH FOR YOUR SITUATION

Because production facilities differ greatly from each other, there is no single
approach to power factor correction that is best in all situations.  Someone must
analyze your situation and decide which approach is best for your plant. 
Depending on plant complexity and other issues, that analysis might be done by

·        a qualified in-house person,

·        an electrical contractor who is experienced in power factor correction,

·        a manufacturer of P.F. correction systems, or

·        a professional electrical engineer. 

The following questions indicate what needs looking into, and might also help you
decide who should do the looking.

·        Correct P.F. at the service entrance, at individual motors, or both?

The electric utility is concerned with the plant’s power factor at the point of
metering, not whether the compensating capacitors are installed at the electrical
service entrance or at individual motors.  Because installing one capacitor bank at
the service entrance is less expensive than installing individual capacitors at each
motor, this is the approach that is usually taken.  There are exceptions, however. 
Perhaps the facility has only one or two large motors.  Or a large motor causes an
excessive drop on the circuit that feeds it.  In these and some other situations,
compensation at the motor can be more cost effective.

·        Install capacitors only, or capacitors plus an automatic P.F. controller?

The electric utility is also concerned that any power factor of less than 100% be
inductive, not capacitive.  A capacitive (leading) power factor can occur if an
inductive (lagging) power factor has been overcompensated by putting too much
capacitance across the line.  This can happen if, for instance, a fixed capacitor bank
is installed at the service entrance and then some large motor loads are
subsequently turned off.  Compensating individual motor loads avoids this
problem.  So does installing a power factor controller. A power factor controller
continuously monitors the service entrance power factor, and connects, at any
given time, only the amount of capacitance needed to accomplish the desired level
of correction.  These controllers often come packaged together with a capacitor
bank in a steel cabinet.  Here the engineering has been done by the manufacturer,
eliminating (in many cases) the need for on-site design by an engineer, and
eliminating that cost.

P.F.-correction capacitor banks can be supplied together with a switching


controller in a metal cabinet.

·        Does plant expansion need to be considered?

If plant expansion is likely at some point in the near or medium-term future, it


makes sense to think about the relationship between expansion and power factor
correction.  As mentioned in the last issue of $mart Energy User, correcting power
factor can sometimes postpone the day when a higher capacity electrical service
entrance will be required.  On the other hand, if the electrical room must be
expanded now to house capacitor banks and a P.F. controller, perhaps it should be
made large enough to allow for future additions.

·        How will P.F. correction affect the supply voltage?

Power factor correction tends to increase voltage levels within the plant.  Where
plant voltages are on the low side, this increase is welcome.  But if the voltage is
already on the high side, it may not be.  Someone should estimate the magnitude of
this increase, consider its possible effects, and contact the electric utility if it would
be unacceptably high.

·        Are power-line harmonics a factor? 


With our present monitoring equipment, the Energy and Minerals Section can
monitor the overall harmonic level on your power circuits.  If the level we measure
is sufficiently low, the designer of your power factor correction system can safely
ignore harmonics.  Above a certain level, however, power line harmonics can
cause capacitor heating and even failure.  Our meter readings are just a first-order
indicator.  If they indicate that a problem may exist, it is important to have more
detailed measurements made — measurements that indicate the amplitude of
specific harmonics such as the fifth, seventh, ninth, etc.  Some manufacturers of 
P.F.-correction equipment will make these measurements for you.

·        Where will the capacitor banks and the controller go?

Any changes or additions to a plant’s electrical system must meet electrical code
requirements.  There are usually costs associated with this, and they should be
considered when deciding what approach to take.  For instance, if installing a
capacitor bank and controller also means enlarging the electrical room, it might be
less expensive to eliminate the controller and mount capacitors near individual
motors.

2.      SETTLE ON SPECIFIC HARDWARE, AND GET A FIRM PRICE ON


THE  JOB

    Once the best approach has been pinned down, the next step is to decide what
specific hardware to purchase, and to get firm quotes.  Unless your organization
happens to be blessed with a staff person experienced in specifying and installing
power factor correction systems, or you have engaged an electrical engineer to
analyze your situation and design your solution, you will probably want a
competent electrical contractor to price out the job for you — preferably a
contractor who works regularly with suppliers of capacitors and P.F. control
systems.

You might also like