Sei sulla pagina 1di 81

DC Circuit Theory

The fundamental relationship between voltage, current and resistance in an electrical or electronic circuit is
called Ohm’s Law.

All materials are made up from atoms, and all atoms consist of protons, neutrons and electrons. Protons,
have a positive electrical charge. Neutrons have no electrical charge (that is they are Neutral), while
Electrons have a negative electrical charge. Atoms are bound together by powerful forces of attraction
existing between the atoms nucleus and the electrons in its outer shell.

When these protons, neutrons and electrons are together within the atom they are happy and stable. But if we
separate them from each other they want to reform and start to exert a potential of attraction called a
potential difference.

Now if we create a closed circuit these loose electrons will start to move and drift back to the protons due to
their attraction creating a flow of electrons. This flow of electrons is called an electrical current. The
electrons do not flow freely through the circuit as the material they move through creates a restriction to the
electron flow. This restriction is called resistance.

Then all basic electrical or electronic circuits consist of three separate but very much related electrical
quantities called: Voltage, ( v ), Current, ( i ) and Resistance, ( Ω ).

Electrical Voltage
Voltage, ( V ) is the potential energy of an electrical supply stored in the form of an electrical charge.
Voltage can be thought of as the force that pushes electrons through a conductor and the greater the voltage
the greater is its ability to “push” the electrons through a given circuit. As energy has the ability to do work
this potential energy can be described as the work required in joules to move electrons in the form of an
electrical current around a circuit from one point or node to another.

Then the difference in voltage between any two points, connections or junctions (called nodes) in a circuit is
known as the Potential Difference, ( p.d. ) commonly called the Voltage Drop.

The Potential difference between two points is measured in Volts with the circuit symbol V, or lowercase
“v“, although Energy, E lowercase “e” is sometimes used to indicate a generated emf (electromotive force).
Then the greater the voltage, the greater is the pressure (or pushing force) and the greater is the capacity to
do work.
A constant voltage source is called a DC Voltage with a voltage that varies periodically with time is called
an AC voltage. Voltage is measured in volts, with one volt being defined as the electrical pressure required
to force an electrical current of one ampere through a resistance of one Ohm. Voltages are generally
expressed in Volts with prefixes used to denote sub-multiples of the voltage such as microvolts ( μV = 10-6
V ), millivolts ( mV = 10-3 V ) or kilovolts ( kV = 103 V ). Voltage can be either positive or negative.

Batteries or power supplies are mostly used to produce a steady D.C. (direct current) voltage source such as
5v, 12v, 24v etc in electronic circuits and systems. While A.C. (alternating current) voltage sources are
available for domestic house and industrial power and lighting as well as power transmission. The mains
voltage supply in the United Kingdom is currently 230 volts a.c. and 110 volts a.c. in the USA.

General electronic circuits operate on low voltage DC battery supplies of between 1.5V and 24V dc The
circuit symbol for a constant voltage source usually given as a battery symbol with a positive, + and
negative, – sign indicating the direction of the polarity. The circuit symbol for an alternating voltage source
is a circle with a sine wave inside.
Voltage Symbols

A simple relationship can be made between a tank of water and a voltage supply. The higher the water tank
above the outlet the greater the pressure of the water as more energy is released, the higher the voltage the
greater the potential energy as more electrons are released.
Voltage is always measured as the difference between any two points in a circuit and the voltage between
these two points is generally referred to as the “Voltage drop“. Note that voltage can exist across a circuit
without current, but current cannot exist without voltage and as such any voltage source whether DC or AC
likes an open or semi-open circuit condition but hates any short circuit condition as this can destroy it.
Electrical Current
Electrical Current, ( I ) is the movement or flow of electrical charge and is measured in Amperes, symbol
i, for intensity). It is the continuous and uniform flow (called a drift) of electrons (the negative particles of an
atom) around a circuit that are being “pushed” by the voltage source. In reality, electrons flow from the
negative (–ve) terminal to the positive (+ve) terminal of the supply and for ease of circuit understanding
conventional current flow assumes that the current flows from the positive to the negative terminal.
Generally in circuit diagrams the flow of current through the circuit usually has an arrow associated with the
symbol, I, or lowercase i to indicate the actual direction of the current flow. However, this arrow usually
indicates the direction of conventional current flow and not necessarily the direction of the actual flow.
Conventional Current Flow

Conventionally this is the flow of positive charge around a circuit, being positive to negative. The diagram
at the left shows the movement of the positive charge (holes) around a closed circuit flowing from the
positive terminal of the battery, through the circuit and returns to the negative terminal of the battery. This
flow of current from positive to negative is generally known as conventional current flow.
This was the convention chosen during the discovery of electricity in which the direction of electric current
was thought to flow in a circuit. To continue with this line of thought, in all circuit diagrams and schematics,
the arrows shown on symbols for components such as diodes and transistors point in the direction of
conventional current flow.
Then Conventional Current Flow gives the flow of electrical current from positive to negative and which
is the opposite in direction to the actual flow of electrons.
Electron Flow

The flow of electrons around the circuit is opposite to the direction of the conventional current flow being
negative to positive.The actual current flowing in an electrical circuit is composed of electrons that flow
from the negative pole of the battery (the cathode) and return back to the positive pole (the anode) of the
battery.
This is because the charge on an electron is negative by definition and so is attracted to the positive terminal.
This flow of electrons is called Electron Current Flow. Therefore, electrons actually flow around a circuit
from the negative terminal to the positive.
Both conventional current flow and electron flow are used by many textbooks. In fact, it makes no
difference which way the current is flowing around the circuit as long as the direction is used consistently.
The direction of current flow does not affect what the current does within the circuit. Generally it is much
easier to understand the conventional current flow – positive to negative.
In electronic circuits, a current source is a circuit element that provides a specified amount of current for
example, 1A, 5A 10 Amps etc, with the circuit symbol for a constant current source given as a circle with an
arrow inside indicating its direction.
Current is measured in Amps and an amp or ampere is defined as the number of electrons or charge (Q in
Coulombs) passing a certain point in the circuit in one second, (t in Seconds).
Electrical current is generally expressed in Amps with prefixes used to denote micro amps ( μA = 10-6A ) or
milliamps ( mA = 10-3A ). Note that electrical current can be either positive in value or negative in value
depending upon its direction of flow around the circuit.
Current that flows in a single direction is called Direct Current, or D.C. and current that alternates back
and forth through the circuit is known as Alternating Current, or A.C.. Whether AC or DC current only
flows through a circuit when a voltage source is connected to it with its “flow” being limited to both the
resistance of the circuit and the voltage source pushing it.
Also, as alternating currents (and voltages) are periodic and vary with time the “effective” or “RMS”, (Root
Mean Squared) value given as Irms produces the same average power loss equivalent to a DC current I average .
Current sources are the opposite to voltage sources in that they like short or closed circuit conditions but
hate open circuit conditions as no current will flow.
Using the tank of water relationship, current is the equivalent of the flow of water through the pipe with the
flow being the same throughout the pipe. The faster the flow of water the greater the current. Note that
current cannot exist without voltage so any current source whether DC or AC likes a short or semi-short
circuit condition but hates any open circuit condition as this prevents it from flowing.
Resistance
Resistance, ( R ) is the capacity of a material to resist or prevent the flow of current or, more specifically,
the flow of electric charge within a circuit. The circuit element which does this perfectly is called the
“Resistor”.
Resistance is a circuit element measured in Ohms, Greek symbol ( Ω, Omega ) with prefixes used to denote
Kilo-ohms ( kΩ = 103Ω ) and Mega-ohms ( MΩ = 106Ω ). Note that resistance cannot be negative in value
only positive.
Resistor Symbols

The amount of resistance a resistor has is determined by the relationship of the current through it to the
voltage across it which determines whether the circuit element is a “good conductor” – low resistance, or a
“bad conductor” – high resistance. Low resistance, for example 1Ω or less implies that the circuit is a good
conductor made from materials such as copper, aluminium or carbon while a high resistance, 1MΩ or more
implies the circuit is a bad conductor made from insulating materials such as glass, porcelain or plastic.
A “semiconductor” on the other hand such as silicon or germanium, is a material whose resistance is half
way between that of a good conductor and a good insulator. Hence the name “semi-conductor”.
Semiconductors are used to make Diodes and Transistors etc.
Resistance can be linear or non-linear in nature, but never negative. Linear resistance obeys Ohm’s Law as
the voltage across the resistor is linearly proportional to the current through it. Non-linear resistance, does
not obey Ohm’s Law but has a voltage drop across it that is proportional to some power of the current.
Resistance is pure and is not affected by frequency with the AC impedance of a resistance being equal to its
DC resistance and as a result can not be negative. Remember that resistance is always positive, and never
negative.
A resistor is classed as a passive circuit element and as such cannot deliver power or store energy. Instead
resistors absorbed power that appears as heat and light. Power in a resistance is always positive regardless of
voltage polarity and current direction.
For very low values of resistance, for example milli-ohms, ( mΩ ) it is sometimes much easier to use the
reciprocal of resistance ( 1/R ) rather than resistance ( R ) itself. The reciprocal of resistance is called
Conductance, symbol ( G ) and represents the ability of a conductor or device to conduct electricity.
In other words the ease by which current flows. High values of conductance implies a good conductor such
as copper while low values of conductance implies a bad conductor such as wood. The standard unit of
measurement given for conductance is the Siemen, symbol (S).
The unit used for conductance is mho (ohm spelt backward), which is symbolized by an inverted Ohm sign
℧. Power can also be expressed using conductance as: p = i2/G = v2G.
The relationship between Voltage, ( v ) and Current, ( i ) in a circuit of constant Resistance, ( R ) would
produce a straight line i-v relationship with slope equal to the value of the resistance as shown.
Voltage, Current and Resistance Summary
Hopefully by now you should have some idea of how electrical Voltage, Current and Resistance are closely
related together. The relationship between Voltage, Current and Resistance forms the basis of Ohm’s law.
In a linear circuit of fixed resistance, if we increase the voltage, the current goes up, and similarly, if we
decrease the voltage, the current goes down. This means that if the voltage is high the current is high, and if
the voltage is low the current is low.
Likewise, if we increase the resistance, the current goes down for a given voltage and if we decrease the
resistance the current goes up. Which means that if resistance is high current is low and if resistance is low
current is high.
Then we can see that current flow around a circuit is directly proportional ( ∝ ) to voltage, ( V↑ causes I↑ )
but inversely proportional ( 1/∝ ) to resistance as, ( R↑ causes I↓ ).
A basic summary of the three units is given below.
 Voltage or potential difference is the measure of potential energy between two points in a circuit and
is commonly referred to as its ” volt drop “.
 When a voltage source is connected to a closed loop circuit the voltage will produce a current
flowing around the circuit.
 In DC voltage sources the symbols +ve (positive) and −ve (negative) are used to denote the polarity
of the voltage supply.
 Voltage is measured in Volts and has the symbol V for voltage or E for electrical energy.
 Current flow is a combination of electron flow and hole flow through a circuit.
 Current is the continuous and uniform flow of charge around the circuit and is measured in Amperes
or Amps and has the symbol I.
 Current is Directly Proportional to Voltage ( I ∝ V )
 The effective (rms) value of an alternating current has the same average power loss equivalent to a
direct current flowing through a resistive element.
 Resistance is the opposition to current flowing around a circuit.
 Low values of resistance implies a conductor and high values of resistance implies an insulator.
 Current is Inversely Proportional to Resistance ( I 1/∝ R )
 Resistance is measured in Ohms and has the Greek symbol Ω or the letter R.
Unit of
Quantity Symbol Abbreviation
Measure
Voltage V or E Volt V
Current I Ampere A
Resistance R Ohms Ω
In the next tutorial about DC Circuits we will look at Ohms Law which is a mathematical equation
explaining the relationship between Voltage, Current, and Resistance within electrical circuits and is the
foundation of electronics and electrical engineering. Ohm’s Law is defined as: V = I*R.

Ohms Law and Power


The relationship between Voltage, Current and Resistance in any DC electrical circuit was firstly discovered
by the German physicist Georg Ohm.
Georg Ohm found that, at a constant temperature, the electrical current flowing through a fixed linear
resistance is directly proportional to the voltage applied across it, and also inversely proportional to the
resistance. This relationship between the Voltage, Current and Resistance forms the basis of Ohms Law and
is shown below.
Ohms Law Relationship

By knowing any two values of the Voltage, Current or Resistance quantities we can use Ohms Law to find
the third missing value. Ohms Law is used extensively in electronics formulas and calculations so it is “very
important to understand and accurately remember these formulas”.
To find the Voltage, ( V )
[ V = I x R ]      V (volts) = I (amps) x R (Ω)
To find the Current, ( I )
[ I = V ÷ R ]      I (amps) = V (volts) ÷ R (Ω)
To find the Resistance, ( R )
[ R = V ÷ I ]      R (Ω) = V (volts) ÷ I (amps)
It is sometimes easier to remember this Ohms law relationship by using pictures. Here the three quantities of
V, I and R have been superimposed into a triangle (affectionately called the Ohms Law Triangle) giving
voltage at the top with current and resistance below. This arrangement represents the actual position of each
quantity within the Ohms law formulas.
Ohms Law Triangle
 
Transposing the standard Ohms Law equation above will give us the following combinations of the same
equation:

 
Then by using Ohms Law we can see that a voltage of 1V applied to a resistor of 1Ω will cause a current of
1A to flow and the greater the resistance value, the less current that will flow for a given applied voltage.
Any Electrical device or component that obeys “Ohms Law” that is, the current flowing through it is
proportional to the voltage across it ( I α V ), such as resistors or cables, are said to be “Ohmic” in nature,
and devices that do not, such as transistors or diodes, are said to be “Non-ohmic” devices.
Electrical Power in Circuits
Electrical Power, ( P ) in a circuit is the rate at which energy is absorbed or produced within a circuit. A
source of energy such as a voltage will produce or deliver power while the connected load absorbs it. Light
bulbs and heaters for example, absorb electrical power and convert it into either heat, or light, or both. The
higher their value or rating in watts the more electrical power they are likely to consume.
The quantity symbol for power is P and is the product of voltage multiplied by the current with the unit of
measurement being the Watt ( W ). Prefixes are used to denote the various multiples or sub-multiples of a
watt, such as: milliwatts (mW = 10-3W) or kilowatts (kW = 103W).
Then by using Ohm’s law and substituting for the values of V, I and R the formula for electrical power can
be found as:
To find the Power (P)
[ P = V x I ]      P (watts) = V (volts) x I (amps)
Also:
[ P = V2 ÷ R ]      P (watts) = V2 (volts) ÷ R (Ω)
Also:
[ P = I2 x R ]      P (watts) = I2 (amps) x R (Ω)
Again, the three quantities have been superimposed into a triangle this time called a Power Triangle with
power at the top and current and voltage at the bottom. Again, this arrangement represents the actual
position of each quantity within the Ohms law power formulas.
The Power Triangle

 
and again, transposing the basic Ohms Law equation above for power gives us the following combinations
of the same equation to find the various individual quantities:

 
So we can see that there are three possible formulas for calculating electrical power in a circuit. If the
calculated power is positive, (+P) in value for any formula the component absorbs the power, that is it is
consuming or using power. But if the calculated power is negative, (–P) in value the component produces or
generates power, in other words it is a source of electrical power such as batteries and generators.
Electrical Power Rating
Electrical components are given a “power rating” in watts that indicates the maximum rate at which the
component converts the electrical power into other forms of energy such as heat, light or motion. For
example, a 1/4W resistor, a 100W light bulb etc.
Electrical devices convert one form of power into another. So for example, an electrical motor will covert
electrical energy into a mechanical force, while an electrical generator converts mechanical force into
electrical energy. A light bulb converts electrical energy into both light and heat.
Also, we now know that the unit of power is the WATT, but some electrical devices such as electric motors
have a power rating in the old measurement of “Horsepower” or hp. The relationship between horsepower
and watts is given as: 1hp = 746W. So for example, a two-horsepower motor has a rating of 1492W, (2 x
746) or 1.5kW.
Ohms Law Pie Chart
To help us understand the the relationship between the various values a little further, we can take all of the
Ohm’s Law equations from above for finding Voltage, Current, Resistance and of course Power and
condense them into a simple Ohms Law pie chart for use in AC and DC circuits and calculations as shown.
Ohms Law Pie Chart
 
As well as using the Ohm’s Law Pie Chart shown above, we can also put the individual Ohm’s Law
equations into a simple matrix table as shown for easy reference when calculating an unknown value.
Ohms Law Matrix Table

Ohms Law Example No1


For the circuit shown below find the Voltage (V), the Current (I), the Resistance (R) and the Power (P).

 
Voltage   [ V = I x R ] = 2 x 12Ω = 24V
Current   [ I = V ÷ R ] = 24 ÷ 12Ω = 2A
Resistance   [ R = V ÷ I ] = 24 ÷ 2 = 12 Ω
Power   [ P = V x I ] = 24 x 2 = 48W
 
Power within an electrical circuit is only present when BOTH voltage and current are present. For example,
in an open-circuit condition, voltage is present but there is no current flow I = 0 (zero), therefore V*0 is 0 so
the power dissipated within the circuit must also be 0. Likewise, if we have a short-circuit condition, current
flow is present but there is no voltage V = 0, therefore 0*I = 0 so again the power dissipated within the
circuit is 0.
As electrical power is the product of V*I, the power dissipated in a circuit is the same whether the circuit
contains high voltage and low current or low voltage and high current flow. Generally, electrical power is
dissipated in the form of Heat (heaters), Mechanical Work such as motors, Energy in the form of radiated
(Lamps) or as stored energy (Batteries).
Electrical Energy in Circuits
Electrical Energy is the capacity to do work, and the unit of work or energy is the joule ( J ). Electrical
energy is the product of power multiplied by the length of time it was consumed. So if we know how much
power, in Watts is being consumed and the time, in seconds for which it is used, we can find the total energy
used in watt-seconds. In other words, Energy = power x time and Power = voltage x current. Therefore
electrical power is related to energy and the unit given for electrical energy is the watt-seconds or joules.

 
Electrical power can also be defined as the rate of by which energy is transferred. If one joule of work is
either absorbed or delivered at a constant rate of one second, then the corresponding power will be
equivalent to one watt so power can be defined as “1Joule/sec = 1Watt”. Then we can say that one watt is
equal to one joule per second and electrical power can be defined as the rate of doing work or the
transferring of energy.
Electrical Power and Energy Triangle

 
or to find the various individual quantities:
 
We said previously that electrical energy is define as being watts per second or joules. Although electrical
energy is measured in Joules it can become a very large value when used to calculate the energy consumed
by a component.
For example, if a 100 watt light bulb is left-“ON” for 24 hours, the energy consumed will be 8,640,000
Joules (100W x 86,400 seconds), so prefixes such as kilojoules (kJ = 103J) or megajoules (MJ = 106J) are
used instead and in this simple example, the energy consumed will be 8.64MJ (mega-joules).
But dealing with joules, kilojoules or megajoules to express electrical energy, the maths involved can end up
with some big numbers and lots of zero’s, so it is much more easier to express electrical energy consumed in
Kilowatt-hours.
If the electrical power consumed (or generated) is measured in watts or kilowatts (thousands of watts) and
the time is measure in hours not seconds, then the unit of electrical energy will be the kilowatt-hours,
(kWhr). Then our 100 watt light bulb above will consume 2,400 watt hours or 2.4kWhr, which is much
easier to understand the 8,640,000 joules.
1 kWhr is the amount of electricity used by a device rated at 1000 watts in one hour and is commonly called
a “Unit of Electricity”. This is what is measured by the utility meter and is what we as consumers purchase
from our electricity suppliers when we receive our bills.
Kilowatt-hours are the standard units of energy used by the electricity meter in our homes to calculate the
amount of electrical energy we use and therefore how much we pay. So if you switch ON an electric fire
with a heating element rated at 1000 watts and left it on for 1 hour you will have consumed 1 kWhr of
electricity. If you switched on two electric fires each with 1000 watt elements for half an hour the total
consumption would be exactly the same amount of electricity – 1kWhr.
So, consuming 1000 watts for one hour uses the same amount of power as 2000 watts (twice as much) for
half an hour (half the time). Then for a 100 watt light bulb to use 1 kWhr or one unit of electrical energy it
would need to be switched on for a total of 10 hours (10 x 100 = 1000 = 1kWhr).
Now that we know what is the relationship between voltage, current and resistance in a circuit, in the next
tutorial relating to DC Circuits, we will look at the Standard Electrical Units used in electrical and electronic
engineering to enable us to calculate these values and see that each value can be represented by either
multiples or sub-multiples of the standard unit.

Electrical Units of Measure


Electrical Units of Measurement are used to express standard electrical units along with their prefixes when
the units are too small or too large to express as a base unit
The standard units of electrical measurement used for the expression of voltage, current and resistance are
the Volt [ V ], Ampere [ A ] and Ohm [ Ω ] respectively.
These electrical units of measurement are based on the International (metric) System, also known as the SI
System with other commonly used electrical units being derived from SI base units.
Sometimes in electrical or electronic circuits and systems it is necessary to use multiples or sub-multiples
(fractions) of these standard electrical measuring units when the quantities being measured are very large or
very small.
The following table gives a list of some of the standard electrical units of measure used in electrical
formulas and component values.
Standard Electrical Units of Measure
Electrical Measuring
Symbol Description
Parameter Unit
Unit of Electrical Potential
Voltage Volt V or E
V = I × R
Unit of Electrical Current
Current Ampere I or i
I = V ÷ R
Unit of DC Resistance
Resistance Ohm R or Ω
R = V ÷ I
Reciprocal of Resistance
Conductance Siemen G or ℧
G = 1 ÷ R
Unit of Capacitance
Capacitance Farad C
C = Q ÷ V
Unit of Electrical Charge
Charge Coulomb Q
Q = C × V
Unit of Inductance
Inductance Henry L or H
VL = -L(di/dt)
Unit of Power
Power Watts W
P = V × I  or  I2 × R
Unit of AC Resistance
Impedance Ohm Z 2 2 2
Z  = R  + X
Unit of Frequency
Frequency Hertz Hz
ƒ = 1 ÷ T
Multiples and Sub-multiples
There is a huge range of values encountered in electrical and electronic engineering between a maximum
value and a minimum value of a standard electrical unit. For example, resistance can be lower than 0.01Ω or
higher than 1,000,000Ω. By using multiples and submultiple’s of the standard unit we can avoid having to
write too many zero’s to define the position of the decimal point. The table below gives their names and
abbreviations.
Prefix Symbol Multiplier Power of Ten
Terra T 1,000,000,000,000 1012
Giga G 1,000,000,000 109
Mega M 1,000,000 106
kilo k 1,000 103
none none 1 100
centi c 1/100 10-2
milli m 1/1,000 10-3
micro µ 1/1,000,000 10-6
nano n 1/1,000,000,000 10-9
pico p 1/1,000,000,000,000 10-12
So to display the units or multiples of units for either Resistance, Current or Voltage we would use as an
example:
 1kV = 1 kilo-volt  –  which is equal to 1,000 Volts.
 1mA = 1 milli-amp  –  which is equal to one thousandths (1/1000) of an Ampere.
 47kΩ = 47 kilo-ohms  –  which is equal to 47 thousand Ohms.
 100uF = 100 micro-farads  –  which is equal to 100 millionths (100/1,000,000) of a Farad.
 1kW = 1 kilo-watt  –  which is equal to 1,000 Watts.
 1MHz = 1 mega-hertz  –  which is equal to one million Hertz.
To convert from one prefix to another it is necessary to either multiply or divide by the difference between
the two values. For example, convert 1MHz into kHz.
Well we know from above that 1MHz is equal to one million (1,000,000) hertz and that 1kHz is equal to one
thousand (1,000) hertz, so one 1MHz is one thousand times bigger than 1kHz. Then to convert Mega-hertz
into Kilo-hertz we need to multiply mega-hertz by one thousand, as 1MHz is equal to 1000 kHz.
Likewise, if we needed to convert kilo-hertz into mega-hertz we would need to divide by one thousand. A
much simpler and quicker method would be to move the decimal point either left or right depending upon
whether you need to multiply or divide.
As well as the “Standard” electrical units of measure shown above, other units are also used in electrical
engineering to denote other values and quantities such as:
 •  Wh – The Watt-Hour, The amount of electrical energy consumed by a circuit over a period of
time. Eg, a light bulb consumes one hundred watts of electrical power for one hour. It is commonly
used in the form of: Wh (watt-hours), kWh (Kilowatt-hour) which is 1,000 watt-hours or MWh
(Megawatt-hour) which is 1,000,000 watt-hours.
 •  dB – The Decibel, The decibel is a one tenth unit of the Bel (symbol B) and is used to represent
gain either in voltage, current or power. It is a logarithmic unit expressed in dB and is commonly
used to represent the ratio of input to output in amplifier, audio circuits or loudspeaker systems.
For example, the dB ratio of an input voltage (V IN) to an output voltage (VOUT) is expressed as
20log10 (Vout/Vin). The value in dB can be either positive (20dB) representing gain or negative (-
20dB) representing loss with unity, ie input = output expressed as 0dB.
 •  θ – Phase Angle, The Phase Angle is the difference in degrees between the voltage waveform and
the current waveform having the same periodic time. It is a time difference or time shift and
depending upon the circuit element can have a “leading” or “lagging” value. The phase angle of a
waveform is measured in degrees or radians.
 •  ω – Angular Frequency, Another unit which is mainly used in a.c. circuits to represent the Phasor
Relationship between two or more waveforms is called Angular Frequency, symbol ω. This is a
rotational unit of angular frequency 2πƒ with units in radians per second, rads/s. The complete
revolution of one cycle is 360 degrees or 2π, therefore, half a revolution is given as 180 degrees or π
rad.
 •  τ – Time Constant, The Time Constant of an impedance circuit or linear first-order system is the
time it takes for the output to reach 63.7% of its maximum or minimum output value when subjected
to a Step Response input. It is a measure of reaction time.
In the next tutorial about DC circuit theory we will look at Kirchhoff’s Circuit Law which along with Ohms
Law allows us to calculate the different voltages and currents circulating around a complex circuit.

Kirchhoffs Circuit Law


Kirchhoffs Circuit Laws allow us to solve complex circuit problems by defining a set of basic network laws
and theorems for the voltages and currents around a circuit
We saw in the Resistors tutorial that a single equivalent resistance, ( RT ) can be found when two or more
resistors are connected together in either series, parallel or combinations of both, and that these circuits obey
Ohm’s Law.
However, sometimes in complex circuits such as bridge or T networks, we can not simply use Ohm’s Law
alone to find the voltages or currents circulating within the circuit. For these types of calculations we need
certain rules which allow us to obtain the circuit equations and for this we can use Kirchhoffs Circuit Law.
In 1845, a German physicist, Gustav Kirchhoff developed a pair or set of rules or laws which deal with the
conservation of current and energy within electrical circuits. These two rules are commonly known as:
Kirchhoffs Circuit Laws with one of Kirchhoffs laws dealing with the current flowing around a closed
circuit, Kirchhoffs Current Law, (KCL) while the other law deals with the voltage sources present in a
closed circuit, Kirchhoffs Voltage Law, (KVL).
Kirchhoffs First Law – The Current Law, (KCL)
Kirchhoffs Current Law or KCL, states that the “total current or charge entering a junction or node is
exactly equal to the charge leaving the node as it has no other place to go except to leave, as no charge is
lost within the node“. In other words the algebraic sum of ALL the currents entering and leaving a node
must be equal to zero, I(exiting) + I(entering) = 0. This idea by Kirchhoff is commonly known as the Conservation
of Charge.
Kirchhoffs Current Law

 
Here, the three currents entering the node, I 1, I2, I3 are all positive in value and the two currents leaving the
node, I4 and I5 are negative in value. Then this means we can also rewrite the equation as;
I1 + I2 + I3 – I4 – I5 = 0
The term Node in an electrical circuit generally refers to a connection or junction of two or more current
carrying paths or elements such as cables and components. Also for current to flow either in or out of a node
a closed circuit path must exist. We can use Kirchhoff’s current law when analysing parallel circuits.
Kirchhoffs Second Law – The Voltage Law, (KVL)
Kirchhoffs Voltage Law or KVL, states that “in any closed loop network, the total voltage around the loop
is equal to the sum of all the voltage drops within the same loop” which is also equal to zero. In other words
the algebraic sum of all voltages within the loop must be equal to zero. This idea by Kirchhoff is known as
the Conservation of Energy.
Kirchhoffs Voltage Law

 
Starting at any point in the loop continue in the same direction noting the direction of all the voltage drops,
either positive or negative, and returning back to the same starting point. It is important to maintain the same
direction either clockwise or anti-clockwise or the final voltage sum will not be equal to zero. We can use
Kirchhoff’s voltage law when analysing series circuits.
When analysing either DC circuits or AC circuits using Kirchhoffs Circuit Laws a number of definitions
and terminologies are used to describe the parts of the circuit being analysed such as: node, paths, branches,
loops and meshes. These terms are used frequently in circuit analysis so it is important to understand them.
Common DC Circuit Theory Terms:
 • Circuit – a circuit is a closed loop conducting path in which an electrical current flows.
 • Path – a single line of connecting elements or sources.
 • Node – a node is a junction, connection or terminal within a circuit were two or more circuit
elements are connected or joined together giving a connection point between two or more branches.
A node is indicated by a dot.
 • Branch – a branch is a single or group of components such as resistors or a source which are
connected between two nodes.
 • Loop – a loop is a simple closed path in a circuit in which no circuit element or node is encountered
more than once.
 • Mesh – a mesh is a single open loop that does not have a closed path. There are no components
inside a mesh.
Note that:
    Components are said to be connected together in Series if the same current value flows through all the
components.
    Components are said to be connected together in Parallel if they have the same voltage applied across
them.
A Typical DC Circuit
Kirchhoffs Circuit Law Example No1
Find the current flowing in the 40Ω Resistor, R3
 

 
The circuit has 3 branches, 2 nodes (A and B) and 2 independent loops.
Using Kirchhoffs Current Law, KCL the equations are given as:
At node A :    I1 + I2 = I3
At node B :    I3 = I1 + I2
Using Kirchhoffs Voltage Law, KVL the equations are given as:
Loop 1 is given as :    10 = R1 I1 + R3 I3 = 10I1 + 40I3
Loop 2 is given as :    20 = R2 I2 + R3 I3 = 20I2 + 40I3
Loop 3 is given as :    10 – 20 = 10I1 – 20I2
As I3 is the sum of I1 + I2 we can rewrite the equations as;
Eq. No 1 :    10 = 10I1 + 40(I1 + I2)  =  50I1 + 40I2
Eq. No 2 :    20 = 20I2 + 40(I1 + I2)  =  40I1 + 60I2
We now have two “Simultaneous Equations” that can be reduced to give us the values of I1 and I2 
Substitution of I1 in terms of I2 gives us the value of I1 as -0.143 Amps
Substitution of I2 in terms of I1 gives us the value of I2 as +0.429 Amps
As :    I3 = I1 + I2
The current flowing in resistor R3 is given as :    -0.143 + 0.429 = 0.286 Amps
and the voltage across the resistor R3 is given as :    0.286 x 40 = 11.44 volts
The negative sign for I1 means that the direction of current flow initially chosen was wrong, but never the
less still valid. In fact, the 20v battery is charging the 10v battery.
Application of Kirchhoffs Circuit Laws
These two laws enable the Currents and Voltages in a circuit to be found, ie, the circuit is said to be
“Analysed”, and the basic procedure for using Kirchhoff’s Circuit Laws is as follows:
 1. Assume all voltages and resistances are given. ( If not label them V1, V2,… R1, R2, etc. )
 2. Label each branch with a branch current. ( I1, I2, I3 etc. )
 3. Find Kirchhoff’s first law equations for each node.
 4. Find Kirchhoff’s second law equations for each of the independent loops of the circuit.
 5. Use Linear simultaneous equations as required to find the unknown currents.
As well as using Kirchhoffs Circuit Law to calculate the various voltages and currents circulating around a
linear circuit, we can also use loop analysis to calculate the currents in each independent loop which helps to
reduce the amount of mathematics required by using just Kirchhoff’s laws. In the next tutorial about DC
circuits, we will look at Mesh Current Analysis to do just that.

Mesh Current Analysis


Mesh Current Analysis is a technique used to find the currents circulating around a loop or mesh with in any
closed path of a circuit.
While Kirchhoff´s Laws give us the basic method for analysing any complex electrical circuit, there are
different ways of improving upon this method by using Mesh Current Analysis or Nodal Voltage Analysis
that results in a lessening of the math’s involved and when large networks are involved this reduction in
maths can be a big advantage.
For example, consider the electrical circuit example from the previous section.
Mesh Current Analysis Circuit

 
One simple method of reducing the amount of math’s involved is to analyse the circuit using Kirchhoff’s
Current Law equations to determine the currents, I1 and I2 flowing in the two resistors. Then there is no need
to calculate the current I3 as its just the sum of I1 and I2. So Kirchhoff’s second voltage law simply becomes:
 Equation No 1 :    10 =  50I1 + 40I2
 Equation No 2 :    20 =  40I1 + 60I2
therefore, one line of math’s calculation have been saved.
Mesh Current Analysis
An easier method of solving the above circuit is by using Mesh Current Analysis or Loop Analysis which
is also sometimes called Maxwell´s Circulating Currents method. Instead of labelling the branch currents
we need to label each “closed loop” with a circulating current.
As a general rule of thumb, only label inside loops in a clockwise direction with circulating currents as the
aim is to cover all the elements of the circuit at least once. Any required branch current may be found from
the appropriate loop or mesh currents as before using Kirchhoff´s method.
For example: :    i1 = I1 , i2 = -I2  and  I3 = I1 – I2
We now write Kirchhoff’s voltage law equation in the same way as before to solve them but the advantage
of this method is that it ensures that the information obtained from the circuit equations is the minimum
required to solve the circuit as the information is more general and can easily be put into a matrix form.
For example, consider the circuit from the previous section.

 
These equations can be solved quite quickly by using a single mesh impedance matrix Z. Each element ON
the principal diagonal will be “positive” and is the total impedance of each mesh. Where as, each element
OFF the principal diagonal will either be “zero” or “negative” and represents the circuit element connecting
all the appropriate meshes.
First we need to understand that when dealing with matrices, for the division of two matrices it is the same
as multiplying one matrix by the inverse of the other as shown.
 
 
having found the inverse of R, as V/R is the same as V x R -1, we can now use it to find the two circulating
currents.

Where:
 [ V ]   gives the total battery voltage for loop 1 and then loop 2
 [ I ]     states the names of the loop currents which we are trying to find
 [ R ]   is the resistance matrix
 [ R-1 ]   is the inverse of the [ R ] matrix
and this gives I1 as -0.143 Amps and I2 as -0.429 Amps
As :    I3 = I1 – I2
The combined current of I3 is therefore given as :   -0.143 – (-0.429) = 0.286 Amps
This is the same value of  0.286 amps current, we found previously in the Kirchhoffs circuit law tutorial.
Mesh Current Analysis Summary
This “look-see” method of circuit analysis is probably the best of all the circuit analysis methods with the
basic procedure for solving Mesh Current Analysis equations is as follows:
 1. Label all the internal loops with circulating currents. (I1, I2, …IL etc)
 2. Write the [ L x 1 ] column matrix [ V ] giving the sum of all voltage sources in each loop.
 3. Write the [ L x L ] matrix, [ R ] for all the resistances in the circuit as follows:

o   R11 = the total resistance in the first loop.
o   Rnn = the total resistance in the Nth loop.
o   RJK = the resistance which directly joins loop J to Loop K.
 4. Write the matrix or vector equation [V]  =  [R] x [I] where [I] is the list of currents to be found.
As well as using Mesh Current Analysis, we can also use node analysis to calculate the voltages around the
loops, again reducing the amount of mathematics required using just Kirchoff’s laws. In the next tutorial
relating to DC circuit theory, we will look at Nodal Voltage Analysis to do just that.

Nodal Voltage Analysis


Nodal Voltage Analysis finds the unknown voltage drops around a circuit between different nodes that
provide a common connection for two or more circuit components
Nodal Voltage Analysis complements the previous mesh analysis in that it is equally powerful and based on
the same concepts of matrix analysis. As its name implies, Nodal Voltage Analysis uses the “Nodal”
equations of Kirchhoff’s first law to find the voltage potentials around the circuit.
So by adding together all these nodal voltages the net result will be equal to zero. Then, if there are “n”
nodes in the circuit there will be “n-1” independent nodal equations and these alone are sufficient to describe
and hence solve the circuit.
At each node point write down Kirchhoff’s first law equation, that is: “the currents entering a node are
exactly equal in value to the currents leaving the node” then express each current in terms of the voltage
across the branch. For “n” nodes, one node will be used as the reference node and all the other voltages will
be referenced or measured with respect to this common node.
For example, consider the circuit from the previous section.
Nodal Voltage Analysis Circuit
 
In the above circuit, node D is chosen as the reference node and the other three nodes are assumed to have
voltages, Va, Vb and Vc with respect to node D. For example;

 
As Va = 10v and Vc = 20v , Vb can be easily found by:

 
again is the same value of 0.286 amps, we found using Kirchhoff’s Circuit Law in the previous tutorial.
From both Mesh and Nodal Analysis methods we have looked at so far, this is the simplest method of
solving this particular circuit. Generally, nodal voltage analysis is more appropriate when there are a larger
number of current sources around. The network is then defined as: [ I ] = [ Y ] [ V ] where [ I ] are the
driving current sources, [ V ] are the nodal voltages to be found and [ Y ] is the admittance matrix of the
network which operates on [ V ] to give [ I ].
Nodal Voltage Analysis Summary.
The basic procedure for solving Nodal Analysis equations is as follows:
 1. Write down the current vectors, assuming currents into a node are positive. ie, a (N x 1)  matrices
for “N” independent nodes.
 2. Write the admittance matrix [Y] of the network where:
o   Y11 = the total admittance of the first node.
o   Y22 = the total admittance of the second node.
o   RJK = the total admittance joining node J to node K.
 3. For a network with “N” independent nodes, [Y] will be an (N x N) matrix and that Ynn will be
positive and Yjk will be negative or zero value.
  
 4. The voltage vector will be (N x L) and will list the “N” voltages to be found.
We have now seen that a number of theorems exist that simplify the analysis of linear circuits. In the next
tutorial we will look at Thevenins Theorem which allows a network consisting of linear resistors and sources
to be represented by an equivalent circuit with a single voltage source and a series resistance.

Thevenin’s Theorem
Thevenin theorem is an analytical method used to change a complex circuit into a simple equivalent circuit
consisting of a single resistance in series with a source voltage
In the previous three tutorials we have looked at solving complex electrical circuits using Kirchhoff’s
Circuit Laws, Mesh Analysis and finally Nodal Analysis. But there are many more “Circuit Analysis
Theorems” available to choose from which can calculate the currents and voltages at any point in a circuit.
In this tutorial we will look at one of the more common circuit analysis theorems (next to Kirchhoff´s) that
has been developed, Thevenin’s Theorem.
Thevenin’s Theorem states that “Any linear circuit containing several voltages and resistances can be
replaced by just one single voltage in series with a single resistance connected across the load“. In other
words, it is possible to simplify any electrical circuit, no matter how complex, to an equivalent two-terminal
circuit with just a single constant voltage source in series with a resistance (or impedance) connected to a
load as shown below.
Thevenin’s Theorem is especially useful in the circuit analysis of power or battery systems and other
interconnected resistive circuits where it will have an effect on the adjoining part of the circuit.
Thevenin’s equivalent circuit

 
As far as the load resistor RL is concerned, any complex “one-port” network consisting of multiple resistive
circuit elements and energy sources can be replaced by one single equivalent resistance Rs and one single
equivalent voltage Vs. Rs is the source resistance value looking back into the circuit and Vs is the open
circuit voltage at the terminals.
For example, consider the circuit from the previous tutorials.

 
Firstly, to analyse the circuit we have to remove the centre 40Ω load resistor connected across the terminals
A-B, and remove any internal resistance associated with the voltage source(s). This is done by shorting out
all the voltage sources connected to the circuit, that is v = 0, or open circuit any connected current sources
making i = 0. The reason for this is that we want to have an ideal voltage source or an ideal current source
for the circuit analysis.
The value of the equivalent resistance, Rs is found by calculating the total resistance looking back from the
terminals A and B with all the voltage sources shorted. We then get the following circuit.

Find the Equivalent Resistance (Rs)

 
The voltage Vs is defined as the total voltage across the terminals A and B when there is an open circuit
between them. That is without the load resistor RL connected.
Find the Equivalent Voltage (Vs)

 
We now need to reconnect the two voltages back into the circuit, and as V S  =  VAB the current flowing
around the loop is calculated as:

This current of 0.33 amperes (330mA) is common to both resistors so the voltage drop across the 20Ω
resistor or the 10Ω resistor can be calculated as:
 
VAB  =  20  –  (20Ω x 0.33amps)  =   13.33 volts.
or
VAB  =  10  +  (10Ω x 0.33amps)  =   13.33 volts, the same.
 
Then the Thevenin’s Equivalent circuit would consist or a series resistance of 6.67Ω and a voltage source of
13.33v. With the 40Ω resistor connected back into the circuit we get:

 
and from this the current flowing around the circuit is given as:

which again, is the same value of 0.286 amps, we found using Kirchhoff’s circuit law in the previous circuit
analysis tutorial.
Thevenin’s theorem can be used as another type of circuit analysis method and is particularly useful in the
analysis of complicated circuits consisting of one or more voltage or current source and resistors that are
arranged in the usual parallel and series connections.
While Thevenin’s circuit theorem can be described mathematically in terms of current and voltage, it is not
as powerful as Mesh Current Analysis or Nodal Voltage Analysis in larger networks because the use of
Mesh or Nodal analysis is usually necessary in any Thevenin exercise, so it might as well be used from the
start. However, Thevenin’s equivalent circuits of Transistors, Voltage Sources such as batteries etc, are very
useful in circuit design.
Thevenin’s Theorem Summary
We have seen here that Thevenin’s theorem is another type of circuit analysis tool that can be used to reduce
any complicated electrical network into a simple circuit consisting of a single voltage source, Vs in series
with a single resistor, Rs.
When looking back from terminals A and B, this single circuit behaves in exactly the same way electrically
as the complex circuit it replaces. That is the i-v relationships at terminals A-B are identical.
The basic procedure for solving a circuit using Thevenin’s Theorem is as follows:
 1. Remove the load resistor RL or component concerned.
 2. Find RS by shorting all voltage sources or by open circuiting all the current sources.
 3. Find VS by the usual circuit analysis methods.
 4. Find the current flowing through the load resistor RL.
In the next tutorial we will look at Nortons Theorem which allows a network consisting of linear resistors
and sources to be represented by an equivalent circuit with a single current source in parallel with a single
source resistance.

Nortons Theorem
Nortons theorem is an analytical method used to change a complex circuit into a simple equivalent circuit
consisting of a single resistance in parallel with a current source
Norton on the other hand reduces his circuit down to a single resistance in parallel with a constant current
source.
Nortons Theorem states that “Any linear circuit containing several energy sources and resistances can be
replaced by a single Constant Current generator in parallel with a Single Resistor“.
As far as the load resistance, RL is concerned this single resistance, RS is the value of the resistance looking
back into the network with all the current sources open circuited and I S is the short circuit current at the
output terminals as shown below.
Nortons equivalent circuit

 
The value of this “constant current” is one which would flow if the two output terminals where shorted
together while the source resistance would be measured looking back into the terminals, (the same as
Thevenin).
For example, consider our now familiar circuit from the previous section.

 
To find the Nortons equivalent of the above circuit we firstly have to remove the centre 40Ω load resistor
and short out the terminals A and B to give us the following circuit.
When the terminals A and B are shorted together the two resistors are connected in parallel across their two
respective voltage sources and the currents flowing through each resistor as well as the total short circuit
current can now be calculated as:
with A-B Shorted Out

 
If we short-out the two voltage sources and open circuit terminals A and B, the two resistors are now
effectively connected together in parallel. The value of the internal resistor Rs is found by calculating the
total resistance at the terminals A and B giving us the following circuit.

Find the Equivalent Resistance (Rs)

 
Having found both the short circuit current, Is and equivalent internal resistance, Rs this then gives us the
following Nortons equivalent circuit.
Nortons equivalent circuit

Ok, so far so good, but we now have to solve with the original 40Ω load resistor connected across terminals
A and B as shown below.
Again, the two resistors are connected in parallel across the terminals A and B which gives us a total
resistance of:

The voltage across the terminals A and B with the load resistor connected is given as:

Then the current flowing in the 40Ω load resistor can be found as:

Once again and using Nortons theorem, the value of current for I 3 is still calculated as 0.286 amps, which we
found using Kirchhoff´s circuit law in the previous tutorials.
Nortons Theorem Summary
The basic procedure for solving a circuit using Nortons Theorem is as follows:
 1. Remove the load resistor RL or component concerned.
 2. Find RS by shorting all voltage sources or by open circuiting all the current sources.
 3. Find IS by placing a shorting link on the output terminals A and B.
 4. Find the current flowing through the load resistor RL.
In a circuit, power supplied to the load is at its maximum when the load resistance is equal to the source
resistance. In the next tutorial we will look at Maximum Power Transfer. The application of the maximum
power transfer theorem can be applied to either simple and complicated linear circuits having a variable load
and is used to find the load resistance that leads to transfer of maximum power to the load.

Maximum Power Transfer


Maximum Power Transfer occurs when the resistive value of the load is equal in value to that of the voltage
sources internal resistance allowing maximum power to be supplied
Generally, this source resistance or even impedance if inductors or capacitors are involved is of a fixed value
in Ohm´s.
However, when we connect a load resistance, RL across the output terminals of the power source, the
impedance of the load will vary from an open-circuit state to a short-circuit state resulting in the power
being absorbed by the load becoming dependent on the impedance of the actual power source. Then for the
load resistance to absorb the maximum power possible it has to be “Matched” to the impedance of the power
source and this forms the basis of Maximum Power Transfer.
The Maximum Power Transfer Theorem is another useful circuit analysis method to ensure that the
maximum amount of power will be dissipated in the load resistance when the value of the load resistance is
exactly equal to the resistance of the power source. The relationship between the load impedance and the
internal impedance of the energy source will give the power in the load. Consider the circuit below.
Thevenins Equivalent Circuit

 
In our Thevenin equivalent circuit above, the maximum power transfer theorem states that “the maximum
amount of power will be dissipated in the load resistance if it is equal in value to the Thevenin or Norton
source resistance of the network supplying the power“.
In other words, the load resistance resulting in greatest power dissipation must be equal in value to the
equivalent Thevenin source resistance, then RL = RS but if the load resistance is lower or higher in value than
the Thevenin source resistance of the network, its dissipated power will be less than maximum.
For example, find the value of the load resistance, RL that will give the maximum power transfer in the
following circuit.
Maximum Power Transfer Example No1

Where:
  RS = 25Ω
  RL is variable between 0 – 100Ω
  VS = 100v

 
Then by using the following Ohm’s Law equations:

 
We can now complete the following table to determine the current and power in the circuit for different
values of load resistance.
Table of Current against Power
RL (Ω) I (amps) P (watts) RL (Ω) I (amps) P (watts)
0 4.0 0 25 2.0 100
5 3.3 55 30 1.8 97
10 2.8 78 40 1.5 94
15 2.5 93 60 1.2 83
20 2.2 97 100 0.8 64
Using the data from the table above, we can plot a graph of load resistance, RL against power, P for different
values of load resistance. Also notice that power is zero for an open-circuit (zero current condition) and also
for a short-circuit (zero voltage condition).
Graph of Power against Load Resistance

 
From the above table and graph we can see that the Maximum Power Transfer occurs in the load when the
load resistance, RL is equal in value to the source resistance, RS that is: RS = RL = 25Ω. This is called a
“matched condition” and as a general rule, maximum power is transferred from an active device such as a
power supply or battery to an external device when the impedance of the external device exactly matches the
impedance of the source.
One good example of impedance matching is between an audio amplifier and a loudspeaker. The output
impedance, ZOUT of the amplifier may be given as between 4Ω and 8Ω, while the nominal input impedance,
ZIN of the loudspeaker may be given as 8Ω only.
Then if the 8Ω speaker is attached to the amplifiers output, the amplifier will see the speaker as an 8Ω load.
Connecting two 8Ω speakers in parallel is equivalent to the amplifier driving one 4Ω speaker and both
configurations are within the output specifications of the amplifier.
Improper impedance matching can lead to excessive power loss and heat dissipation. But how could you
impedance match an amplifier and loudspeaker which have very different impedances. Well, there are
loudspeaker impedance matching transformers available that can change impedances from 4Ω to 8Ω, or to
16Ω’s to allow impedance matching of many loudspeakers connected together in various combinations such
as in PA (public address) systems.
Transformer Impedance Matching
One very useful application of impedance matching in order to provide maximum power transfer between
the source and the load is in the output stages of amplifier circuits. Signal transformers are used to match the
loudspeakers higher or lower impedance value to the amplifiers output impedance to obtain maximum sound
power output. These audio signal transformers are called “matching transformers” and couple the load to the
amplifiers output as shown below.
Transformer Impedance Matching

 
The maximum power transfer can be obtained even if the output impedance is not the same as the load
impedance. This can be done using a suitable “turns ratio” on the transformer with the corresponding ratio of
load impedance, ZLOAD to output impedance, ZOUT matches that of the ratio of the transformers primary turns
to secondary turns as a resistance on one side of the transformer becomes a different value on the other.
If the load impedance, ZLOAD is purely resistive and the source impedance is purely resistive, ZOUT then the
equation for finding the maximum power transfer is given as:

 
Where: NP is the number of primary turns and NS the number of secondary turns on the transformer. Then by
varying the value of the transformers turns ratio the output impedance can be “matched” to the source
impedance to achieve maximum power transfer. For example,
Maximum Power Transfer Example No2
If an 8Ω loudspeaker is to be connected to an amplifier with an output impedance of 1000Ω, calculate the
turns ratio of the matching transformer required to provide maximum power transfer of the audio signal.
Assume the amplifier source impedance is Z1, the load impedance is Z2 with the turns ratio given as N.
 
Generally, small high frequency audio transformers used in low power amplifier circuits are nearly always
regarded as ideal for simplicity, so any losses can be ignored.
In the next tutorial about DC circuit theory, we will look at Star Delta Transformation which allows us to
convert balanced star connected circuits into equivalent delta and vice versa.

Star Delta Transformation


Star-Delta Transformations and Delta-Star Transformations allow us to convert impedances connected
together in a 3-phase configuration from one type of connection to another
We can now solve simple series, parallel or bridge type resistive networks using Kirchhoff´s Circuit Laws,
mesh current analysis or nodal voltage analysis techniques but in a balanced 3-phase circuit we can use
different mathematical techniques to simplify the analysis of the circuit and thereby reduce the amount of
math’s involved which in itself is a good thing.
Standard 3-phase circuits or networks take on two major forms with names that represent the way in which
the resistances are connected, a Star connected network which has the symbol of the letter, Υ (wye) and a
Delta connected network which has the symbol of a triangle, Δ (delta).
If a 3-phase, 3-wire supply or even a 3-phase load is connected in one type of configuration, it can be easily
transformed or changed it into an equivalent configuration of the other type by using either the Star Delta
Transformation or Delta Star Transformation process.
A resistive network consisting of three impedances can be connected together to form a T or “Tee”
configuration but the network can also be redrawn to form a Star or Υ type network as shown below.
T-connected and Equivalent Star Network

 
As we have already seen, we can redraw the T resistor network above to produce an electrically equivalent
Star or Υ type network. But we can also convert a Pi or π type resistor network into an electrically
equivalent Delta or Δ type network as shown below.
Pi-connected and Equivalent Delta Network

 
Having now defined exactly what is a Star and Delta connected network it is possible to transform the Υ
into an equivalent Δ circuit and also to convert a Δ into an equivalent Υ circuit using a the transformation
process.
This process allows us to produce a mathematical relationship between the various resistors giving us a Star
Delta Transformation as well as a Delta Star Transformation.
These circuit transformations allow us to change the three connected resistances (or impedances) by their
equivalents measured between the terminals 1-2, 1-3 or 2-3 for either a star or delta connected circuit.
However, the resulting networks are only equivalent for voltages and currents external to the star or delta
networks, as internally the voltages and currents are different but each network will consume the same
amount of power and have the same power factor to each other.
Delta Star Transformation
To convert a delta network to an equivalent star network we need to derive a transformation formula for
equating the various resistors to each other between the various terminals. Consider the circuit below.
Delta to Star Network

 
Compare the resistances between terminals 1 and 2.
Resistance between the terminals 2 and 3.

Resistance between the terminals 1 and 3.

This now gives us three equations and taking equation 3 from equation 2 gives:

Then, re-writing Equation 1 will give us:

Adding together equation 1 and the result above of equation 3 minus equation 2 gives:
From which gives us the final equation for resistor P as:

 
Then to summarize a little about the above maths, we can now say that resistor P in a Star network can be
found as Equation 1 plus (Equation 3 minus Equation 2) or  Eq1 + (Eq3 – Eq2).
Similarly, to find resistor Q in a star network, is equation 2 plus the result of equation 1 minus equation 3
or  Eq2 + (Eq1 – Eq3) and this gives us the transformation of Q as:

and again, to find resistor R in a Star network, is equation 3 plus the result of equation 2 minus equation 1
or  Eq3 + (Eq2 – Eq1) and this gives us the transformation of R as:

When converting a delta network into a star network the denominators of all of the transformation formulas
are the same: A + B + C, and which is the sum of ALL the delta resistances. Then to convert any delta
connected network to an equivalent star network we can summarized the above transformation equations as:
Delta to Star Transformations Equations

 
If the three resistors in the delta network are all equal in value then the resultant resistors in the equivalent
star network will be equal to one third the value of the delta resistors. This gives each resistive branch in the
star network a value of: RSTAR = 1/3*RDELTA which is the same as saying: (RDELTA)/3
Delta – Star Example No1
Convert the following Delta Resistive Network into an equivalent Star Network.
Star Delta Transformation
Star Delta transformation is simply the reverse of above. We have seen that when converting from a delta
network to an equivalent star network that the resistor connected to one terminal is the product of the two
delta resistances connected to the same terminal, for example resistor P is the product of resistors A and B
connected to terminal 1.
By rewriting the previous formulas a little we can also find the transformation formulas for converting a
resistive star network to an equivalent delta network giving us a way of producing a star delta transformation
as shown below.
Star to Delta Transformation

The value of the resistor on any one side of the delta, Δ network is the sum of all the two-product
combinations of resistors in the star network divide by the star resistor located “directly opposite” the delta
resistor being found. For example, resistor A is given as:

with respect to terminal 3 and resistor B is given as:

with respect to terminal 2 with resistor C given as:

with respect to terminal 1.


By dividing out each equation by the value of the denominator we end up with three separate transformation
formulas that can be used to convert any Delta resistive network into an equivalent star network as given
below.
Star Delta Transformation Equations

 
One final point about converting a star resistive network to an equivalent delta network. If all the resistors in
the star network are all equal in value then the resultant resistors in the equivalent delta network will be
three times the value of the star resistors and equal, giving:  RDELTA = 3*RSTAR
Star – Delta Example No2
Convert the following Star Resistive Network into an equivalent Delta Network.

 
Both Star Delta Transformation and Delta Star Transformation allows us to convert one type of circuit
connection into another type in order for us to easily analyse the circuit. These transformation techniques
can be used to good effect for either star or delta circuits containing resistances or impedances.

Voltage Sources
A Voltage Source is a device that generates an exact output voltage which, in theory, does not change
regardless of the load current
We have seen throughout this Basic Electronics Tutorials website that there are two types of elements
within an electrical or electronics circuit: passive elements and active elements. An active element is one that
is capable of continuously supplying energy to a circuit, such as a battery, a generator, an operational
amplifier, etc. A passive element on the other hand are physical elements such as resistors, capacitors,
inductors, etc, which cannot generate electrical energy by themselves but only consume it.
The types of active circuit elements that are most important to us are those that supply electrical energy to
the circuits or network connected to them. These are called “electrical sources” with the two types of
electrical sources being the voltage source and the current source. The current source is usually less
common in circuits than the voltage source, but both are used and can be regarded as complements of each
other.
An electrical supply or simply, “a source”, is a device that supplies electrical power to a circuit in the form
of a voltage source or a current source. Both types of electrical sources can be classed as a direct (DC) or
alternating (AC) source in which a constant voltage is called a DC voltage and one that varies sinusoidally
with time is called an AC voltage. So for example, batteries are DC sources and the 230V wall socket or
mains outlet in your home is an AC source.
We said earlier that electrical sources supply energy, but one of the interesting characteristic of an electrical
source, is that they are also capable of converting non-electrical energy into electrical energy and vice versa.
For example, a battery converts chemical energy into electrical energy, while an electrical machine such as a
DC generator or an AC alternator converts mechanical energy into electrical energy.
Renewable technologies can convert energy from the sun, the wind, and waves into electrical or thermal
energy. But as well as converting energy from one source to another, electrical sources can both deliver or
absorb energy allowing it to flow in both directions.
Another important characteristic of an electrical source and one which defines its operation, are its I-V
characteristics. The I-V characteristic of an electrical source can give us a very nice pictorial description of
the source, either as a voltage source and a current source as shown.
Electrical Sources

 
Electrical sources, both as a voltage source or a current source can be classed as being either independent
(ideal) or dependent, (controlled) that is whose value depends upon a voltage or current elsewhere within the
circuit, which itself can be either constant or time-varying.
When dealing with circuit laws and analysis, electrical sources are often viewed as being “ideal”, that is the
source is ideal because it could theoretically deliver an infinite amount of energy without loss thereby
having characteristics represented by a straight line. However, in real or practical sources there is always a
resistance either connected in parallel for a current source, or series for a voltage source associated with the
source affecting its output.
The Voltage Source
A voltage source, such as a battery or generator, provides a potential difference (voltage) between two points
within an electrical circuit allowing current to flowing around it. Remember that voltage can exist without
current. A battery is the most common voltage source for a circuit with the voltage that appears across the
positive and negative terminals of the source being called the terminal voltage.
Ideal Voltage Source

 
An ideal voltage source is defined as a two terminal active element that is capable of supplying and
maintaining the same voltage, (v) across its terminals regardless of the current, (i) flowing through it. In
other words, an ideal voltage source will supply a constant voltage at all times regardless of the value of the
current being supplied producing an I-V characteristic represented by a straight line.
Then an ideal voltage source is known as an Independent Voltage Source as its voltage does not depend on
either the value of the current flowing through the source or its direction but is determined solely by the
value of the source alone. So for example, an automobile battery has a 12V terminal voltage that remains
constant as long as the current through it does not become to high, delivering power to the car in one
direction and absorbing power in the other direction as it charges.
On the other hand, a Dependent Voltage Source or controlled voltage source, provides a voltage supply
whose magnitude depends on either the voltage across or current flowing through some other circuit
element. A dependent voltage source is indicated with a diamond shape and are used as equivalent electrical
sources for many electronic devices, such as transistors and operational amplifiers.
Connecting Voltage Sources Together
Ideal voltage sources can be connected together in both parallel or series the same as for any circuit element.
Series voltages add together while parallel voltages have the same value. Note that unequal ideal voltage
sources cannot be connected directly together in parallel.
Voltage Source in Parallel
While not best practice for circuit analysis, ideal voltage sources can be connected in parallel provided they
are of the same voltage value. Here in this example, two 10 volt voltage source are combined to produce 10
volts between terminals A and B. Ideally, there would be just one single voltage source of 10 volts given
between terminals A and B.
What is not allowed or is not best practice, is connecting together ideal voltage sources that have different
voltage values as shown, or are short-circuited by an external closed loop or branch.
Badly Connected Voltage Sources

 
However, when dealing with circuit analysis, voltage sources of different values can be used providing there
are other circuit elements in between them to comply with Kirchoff’s Voltage Law, KVL.
Unlike parallel connected voltage sources, ideal voltage sources of different values can be connected
together in series to form a single voltage source whose output will be the algebraic addition or subtraction
of the voltages used. Their connection can be as: series-aiding or series-opposing voltages as shown.
Voltage Source in Series

 
Series aiding voltage sources are series connected sources with their polarities connected so that the plus
terminal of one is connected to the negative terminal of the next allowing current to flow in the same
direction. In the example above, the two voltages of 10V and 5V of the first circuit can be added, for a V S of
10 + 5 = 15V. So the voltage across terminals A and B is 15 volts.
Series opposing voltage sources are series connected sources which have their polarities connected so that
the plus terminal or the negative terminals are connected together as shown in the second circuit above. The
net result is that the voltages are subtracted from each other. Then the two voltages of 10V and 5V of the
second circuit are subtracted with the smaller voltage subtracted from the larger voltage. Resulting in a V S of
10 – 5 = 5V.
The polarity across terminals A and B is determined by the larger polarity of the voltage sources, in this
example terminal A is positive and terminal B is negative resulting in +5 volts. If the series-opposing
voltages are equal, the net voltage across A and B will be zero as one voltage balances out the other. Also
any currents (I) will also be zero, as without any voltage source, current can not flow.
Voltage Source Example No1
Two series aiding ideal voltage sources of 6 volts and 9 volts respectively are connected together to supply a
load resistance of 100 Ohms. Calculate: the source voltage, V S, the load current through the resistor, I R and
the total power, P dissipated by the resistor. Draw the circuit.

Thus, VS = 15V, IR = 150mA or 0.15A, and PR = 2.25W.


Practical Voltage Source
We have seen that an ideal voltage source can provide a voltage supply that is independent of the current
flowing through it, that is, it maintains the same voltage value always. This idea may work well for circuit
analysis techniques, but in the real world voltage sources behave a little differently as for a practical voltage
source, its terminal voltage will actually decrease with an increase in load current.
As the terminal voltage of an ideal voltage source does not vary with increases in the load current, this
implies that an ideal voltage source has zero internal resistance, R S = 0. In other words, it is a resistorless
voltage source. In reality all voltage sources have a very small internal resistance which reduces their
terminal voltage as they supply higher load currents.
For non-ideal or practical voltage sources such as batteries, their internal resistance (R S) produces the same
effect as a resistance connected in series with an ideal voltage source as these two series connected elements
carry the same current as shown.
Ideal and Practical Voltage Source

 
You may have noticed that a practical voltage source closely resembles that of a Thevenin’s equivalent
circuit as Thevenin’s theorem states that “any linear network containing resistances and sources of emf and
current may be replaced by a single voltage source, V S in series with a single resistance, R S“. Note that if the
series source resistance is low, the voltage source is ideal. When the source resistance is infinite, the voltage
source is open-circuited.
In the case of all real or practical voltage sources, this internal resistance, R S no matter how small has an
effect on the I-V characteristic of the source as the terminal voltage falls off with an increase in load current.
This is because the same load current flows through RS.
Ohms law tells us that when a current, (i) flows through a resistance, a voltage drop is produce across the
same resistance. The value of this voltage drop is given as i*RS. Then VOUT will equal the ideal voltage
source, VS minus the i*RS voltage drop across the resistor. Remember that in the case of an ideal source
voltage, RS is equal to zero as there is no internal resistance, therefore the terminal voltage is same as VS.
Then the voltage sum around the loop given by Kirchoff’s voltage law, KVL is: V OUT = VS – i*RS. This
equation can be plotted to give the I-V characteristics of the actual output voltage. It will give a straight line
with a slope –RS which intersects the vertical voltage axis at the same point as V S when the current i = 0 as
shown.
Practical Voltage Source Characteristics

 
Therefore, all ideal voltage sources will have a straight line I-V characteristic but non-ideal or real practical
voltage sources will not but instead will have an I-V characteristic that is slightly angled down by an amount
equal to i*RS where RS is the internal source resistance (or impedance). The I-V characteristics of a real
battery provides a very close approximation of an ideal voltage source since the source resistance R S is
usually quite small.
The decrease in the angle of the slope of the I-V characteristics as the current increases is known as
regulation. Voltage regulation is an important measure of the quality of a practical voltage source as it
measures the variation in terminal voltage between no load, that is when I L = 0, (an open-circuit) and full
load, that is when IL is at maximum, (a short-circuit).
Voltage Source Example No2
A battery supply consists of an ideal voltage source in series with an internal resistor. The voltage and
current measured at the terminals of the battery were found to be V OUT1 = 130V at 10A, and VOUT2 = 100V at
25A. Calculate the voltage rating of the ideal voltage source and the value of its internal resistance. Draw the
I-V characteristics.
Firstly lets define in simple “simultaneous equation form“, the two voltage and current outputs of the
battery supply given as: VOUT1 and VOUT2.

 
As with have the voltages and currents in a simultaneous equation form, to find V S we will first multiply
VOUT1 by five, (5) and VOUT2 by two, (2) as shown to make the value of the two currents, (i) the same for both
equations.
 
Having made the co-efficients for RS the same by multiplying through with the previous constants, we now
multiply the second equation VOUT2 by minus one, (-1) to allow for the subtraction of the two equations so
that we can solve for VS as shown.

 
Knowing that the ideal voltage source, VS is equal to 150 volts, we can use this value for equation V OUT1 (or
VOUT2 if so wished) and solve to find the series resistance, RS.
 
Then for our simple example, the batteries internal voltage source is calculated as: V S = 150 volts, and its
internal resistance as: RS = 2Ω. The I-V characteristics of the battery are given as:
Battery I-V Characteristics

Dependent Voltage Source


Unlike an ideal voltage source which produces a constant voltage across its terminals regardless of what is
connected to it, a controlled or dependent voltage source changes its terminal voltage depending upon the
voltage across, or the current through, some other element connected to the circuit, and as such it is
sometimes difficult to specify the value of a dependent voltage source, unless you know the actual value of
the voltage or current on which it depends.
Dependent voltage sources behave similar to the electrical sources we have looked at so far, both practical
and ideal (independent) the difference this time is that a dependent voltage source can be controlled by an
input current or voltage. A voltage source that depends on a voltage input is generally referred to as a
Voltage Controlled Voltage Source or VCVS. A voltage source that depends on a current input is referred
too as a Current Controlled Voltage Source or CCVS.
Ideal dependent sources are commonly used in the analysing the input/output characteristics or the gain of
circuit elements such as operational amplifiers, transistors and integrated circuits. Generally, an ideal voltage
dependent source, either voltage or current controlled is designated by a diamond-shaped symbol as shown.
Dependent Voltage Source Symbols

 
An ideal dependent voltage-controlled voltage source, VCVS, maintains an output voltage equal to some
multiplying constant (basically an amplification factor) times the controlling voltage present elsewhere in
the circuit. As the multiplying constant is, well, a constant, the controlling voltage, V IN will determine the
magnitude of the output voltage, VOUT. In other words, the output voltage “depends” on the value of input
voltage making it a dependent voltage source and in many ways, an ideal transformer can be thought of as a
VCVS device with the amplification factor being its turns ratio.
Then the VCVS output voltage is determined by the following equation: VOUT = μVIN. Note that the
multiplying constant μ is dimensionless as it is purely a scaling factor because μ = VOUT/VIN, so its units will
be volts/volts.
An ideal dependent current-controlled voltage source, CCVS, maintains an output voltage equal to some
multiplying constant (rho) times a controlling current input generated elsewhere within the connected
circuit. Then the output voltage “depends” on the value of the input current, again making it a dependent
voltage source.
As a controlling current, IIN determines the magnitude of the output voltage, VOUT times the magnification
constant ρ (rho), this allows us to model a current-controlled voltage source as a trans-resistance amplifier as
the multiplying constant, ρ gives us the following equation: V OUT = ρIIN. This multiplying constant ρ (rho)
has the units of Ohm’s because ρ = VOUT/IIN, and its units will therefore be volts/amperes.
Voltage Source Summary
We have seen here that a Voltage Source can be either an ideal independent voltage source, or a controlled
dependent voltage source. Independent voltage sources supply a constant voltage that does not depend on
any other quantity within the circuit. Ideal independent sources can be batteries, DC generators or time-
varying AC voltage supplies from alternators.
Independent voltage sources can be modelled as either an ideal voltage source, (R S = 0) where the output is
constant for all load currents, or a non-ideal or practical, such as a battery with a resistance connected in
series with the circuit to represent the internal resistance of the source. Ideal voltage sources can be
connected together in parallel only if they are of the same voltage value. Series-aiding or series-opposing
connections will affect the output value.
Also for solving circuit analysis and complex theorems, voltage sources become short-circuited sources
making their voltage equal to zero to help solve the network. Note also that voltage sources are capable of
both delivering or absorbing power.
Ideal dependent voltage sources represented by a diamond-shaped symbol, are dependent on, and are
proportional too an external controlling voltage or current. The multiplying constant, μ for a VCVS has no
units, while the multiplying constant ρ for a CCVS has units of Ohm’s. A dependent voltage source is of
great interest to model electronic devices or active devices such as operational amplifiers and transistors that
have gain.
In the next tutorial about electrical sources, we will look at the compliment of the voltage source, that is the
current source and see that current sources can also be classed as dependent or independent electrical
sources.

Current Sources
A Current Source is an active circuit element that is capable of supplying a constant current flow to a
circuit regardless of the voltage developed across its terminals
As its name implies, a current source is a circuit element that maintains a constant current flow regardless of
the voltage developed across its terminals as this voltage is determined by other circuit elements. That is, an
ideal constant current source continually provides a specified amount of current regardless of the impedance
that it is driving and as such, an ideal current source could, in theory, supply an infinite amount of energy.
So just as a voltage source may be rated, for example, as 5 volts or 10 volts, etc, a current source will also
have a current rating, for example, 3 amperes or 15 amperes, etc.
Ideal constant current sources are represented in a similar manner to voltage sources, but this time the
current source symbol is that of a circle with an arrow inside to indicates the direction of the flow of the
current. The direction of the current will correspond to the polarity of the corresponding voltage, flowing out
from the positive terminal. The letter “i” is used to indicate that it is a current source as shown.
Ideal Current Source

 
Then an ideal current source is called a “constant current source” as it provides a constant steady state
current independent of the load connected to it producing an I-V characteristic represented by a straight line.
As with voltage sources, the current source can be either independent (ideal) or dependent (controlled) by a
voltage or current elsewhere in the circuit, which itself can be constant or time-varying.
Ideal independent current sources are typically used to solve circuit theorems and for circuit analysis
techniques for circuits that containing real active elements. The simplest form of a current source is a
resistor in series with a voltage source creating currents ranging from a few milli-amperes to many hundreds
of amperes. Remember that a zero-value current source is an open circuit as R = 0.
The concept of a current source is that of a two-terminal element that allows the flow of current indicated by
the direction of the arrow. Then a current source has a value, i, in units of amperes, (A) which are typically
abbreviated to amps. The physical relationship between a current source and voltage variables around a
network is given by Ohm’s law as these voltage and current variables will have specified values.
It may be difficult to specify the magnitude and polarity of voltage of an ideal current source as a function of
the current especially if there are other voltage or current sources in the connected circuit. Then we may
know the current supplied by the current source but not the voltage across it unless the power supplied by
the current source is given, as P = V*I.
However, if the current source is the only source within the circuit, then the polarity of voltage across the
source will be easier to establish. If however there is more than one source, then the terminal voltage will be
dependent upon the network in which the source is connected.
Connecting Current Sources Together
Just like voltage sources, ideal current sources can also be connected together to increase (or decrease) the
available current. But there are rules on how two or more independent current sources with different values
can be connected, either in series or parallel.
Current Source in Parallel
Connecting two or more current sources in parallel is equivalent to one current source whose total current
output is given as the algebraic addition of the individual source currents. Here in this example, two 5 amp
current sources are combined to produce 10 amps as IT = I1 + I2.
Current sources of different values may be connected together in parallel. For example, one of 5 amps and
one of 3 amps would combined to give a single current source of 8 amperes as the arrows representing the
current source both point in the same direction. Then as the two currents add together, their connection is
said to be: parallel-aiding.
While not best practice for circuit analysis, parallel-opposing connections use current sources that are
connected in opposite directions to form a single current source whose value is the algebraic subtraction of
the individual sources.
Parallel Opposing Current Sources

 
Here, as the two current sources are connected in opposite directions (indicated by their arrows), the two
currents subtract from each other as the two provide a closed-loop path for a circulating current complying
with Kirchoff’s Current Law, KCL. So for example, two current sources of 5 amps each would result in zero
output as 5A -5A = 0A. Likewise, if the two currents are of different values, 5A and 3A, then the output will
be the subtracted value with the smaller current subtracted from the larger current. Resulting in a I T of
5 - 3 = 2A.
We have seen that ideal current sources can be connected together in parallel to form parallel-aiding or
parallel-opposing current sources. What is not allowed or is not best practice for circuit analysis, is
connecting together ideal current sources in series combinations.
Current Sources in Series

 
Current sources are not allowed to be connected together in series, either of the same value or ones with
different values. Here in this example, two current sources of 5 amps each are connected together in series,
but what is the resulting current value. Is it equal to one source of 5 amps, or is it equal to the addition of the
two sources, that is 10 amps. Then series connected current sources add an unknown factor into circuit
analysis, which is not good.
Also, another reason why series connected sources are not allowed for circuit analysis techniques is that they
may not supply the same current in the same direction. Series-aiding or series-opposing currents do not exist
for ideal current sources.
Current Source Example No1
Two current sources of 250 milli-amps and 150 milli-amps respectively are connected together in a parallel-
aiding configuration to supply a connected load of 20 ohms. Calculate the voltage drop across the load and
the power dissipated. Draw the circuit.

Then, IT = 0.4A or 400mA, VR = 8V, and PR = 3.2W


Practical Current Source
We have seen that an ideal constant current source can supply the same amount of current indefinitely
regardless of the voltage across its terminals, thus making it an independent source. This therefore implies
that the current source has an infinite internal resistance, (R = ∞). This idea works well for circuit analysis
techniques, but in the real world current sources behave a little differently as practical current sources
always have an internal resistance, no matter how large (usually in the mega-ohms range), causing the
generated source to vary somewhat with the load.
A practical or non-ideal current source can be represented as an ideal source with an internal resistance
connected across it. The internal resistance (RP) produces the same effect as a resistance connected in
parallel (shunt) with the current source as shown. Remember that circuit elements in parallel have exactly
the same voltage drop across them.
Ideal and Practical Current Source

 
You may have noticed that a practical current source closely resembles that of a Norton’s equivalent circuit
as Norton’s theorem states that “any linear dc network can be replaced by an equivalent circuit consisting of
a constant-current source, IS in parallel with a resistor, RP“. Note that if this parallel resistance is very low,
RP = 0, the current source is short-circuited. When the parallel resistance is very high or infinite, R P ≈ ∞, the
current source can be modelled as ideal.
An ideal current source plots a horizontal line on the I-V characteristic as shown previously above. However
as practical current sources have an internal source resistance, this takes some of the current so the
characteristic of this practical source is not flat and horizontal but will reduce as the current is now splitting
into two parts, with one part of the current flowing into the parallel resistance, R P and the other part of the
current flowing straight to the output terminals.
Ohms law tells us that when a current, (i) flows through a resistance, (R) a voltage drop is produce across
the same resistance. The value of this voltage drop will be given as i*R P. Then VOUT will be equal to the
voltage drop across the resistor with no load attached. We remember that for an ideal source current, R P is
infinite as there is no internal resistance, therefore the terminal voltage will be zero as there is no voltage
drop.
The sum of the current around the loop given by Kirchoff’s current law, KCL is: I OUT = IS - VS/RP. This
equation can be plotted to give the I-V characteristics of the output current. It is given as a straight line with
a slope –RP which intersects the vertical voltage axis at the same point as IS when the source is ideal as
shown.
Practical Current Source Characteristics

 
Therefore, all ideal current sources will have a straight line I-V characteristic but non-ideal or real practical
current sources will have an I-V characteristic that is slightly angled down by an amount equal to V OUT/RP
where RP is the internal source resistance.
Current Source Example No2
A practical current source consists of a 3A ideal current source which has an internal resistance of 500
Ohms. With no-load attached, calculate the current sources open-circuit terminal voltage and the no-load
power absorbed by the internal resistor.
1. No-load values:

 
Then the open circuit voltage across the internal source resistance and terminals A and B (V AB) is calculated
at 1500 volts.
Part 2: If a 250 Ohm load resistor is connected to the terminals of the same practical current source,
calculate the current through each resistance, the power absorbed by each resistance and the voltage drop
across the load resistor. Draw the circuit.
2. Data given with load connected: IS = 3A, RP = 500Ω and RL = 250Ω

 
2a. To find the currents in each resistive branch, we can use the current-division rule.

 
2b. The power absorbed by each resistor is given as:

 
2c. Then the voltage drop across the load resistor, RL is given as:

 
We can see that the terminal voltage of an open-circuited practical current source can be very high it will
produce whatever voltage is needed, 1500 volts in this example, to supply the specified current. In theory,
this terminal voltage can be infinite as the source attempts to deliver the rated current.
Connecting a load across its terminals will reduce the voltage, 500 volts in this example, as now the current
has somewhere to go and for a constant current source, the terminal voltage is directly proportional to the
load resistance.
In the case of non-ideal current sources that each have an internal resistance, the total internal resistance (or
impedance) will be the result of combining them together in parallel, exactly the same as for resistors in
parallel.
Dependent Current Source
We now know that an ideal current source provides a specified amount of current completely independent of
the voltage across it and as such will produce whatever voltage is necessary to maintain the required current.
This then makes it completely independent of the circuit to which it is connected to resulting in it being
called an ideal independent current source.
A controlled or dependent current source on the other hand changes its available current depending upon the
voltage across, or the current through, some other element connected to the circuit. In other words, the
output of a dependent current source is controlled by another voltage or current.
Dependent current sources behave similar to the current sources we have looked at so far, both ideal
(independent) and practical. The difference this time is that a dependent current source can be controlled by
an input voltage or current. A current source that depends on a voltage input is generally referred to as a
Voltage Controlled Current Source or VCCS. A current source that depends on a current input is
generally referred too as a Current Controlled Current Source or CCCS.
Generally, an ideal current dependent source, either voltage or current controlled is designated by a
diamond-shaped symbol where an arrow indicates the direction of the current, i as shown.
Dependent Current Source Symbols

 
An ideal dependent voltage-controlled current source, VCCS, maintains an output current, IOUT that is
proportional to the controlling input voltage, VIN. In other words, the output current “depends” on the value
of input voltage making it a dependent current source.
Then the VCCS output current is defined by the following equation: I OUT = αVIN. This multiplying constant
α (alpha) has the SI units of mhos, ℧ (an inverted Ohms sign) because α = IOUT/VIN, and its units will
therefore be amperes/volt.
An ideal dependent current-controlled current source, CCCS, maintains an output current that is proportional
to a controlling input current. Then the output current “depends” on the value of the input current, again
making it a dependent current source.
As a controlling current, IIN determines the magnitude of the output current, IOUT times the magnification
constant β (beta), the output current for a CCCS element is determined by the following equation:
IOUT = βIIN. Note that the multiplying constant β is a dimensionless scaling factor as β = IOUT/IIN, so therefore
its units would be amperes/amperes.
Current Source Summary
We have seen in this tutorial about Current Sources, that an ideal current source, (R = ∞) is an active
element that provides a constant current which is totally independent of the voltage across it as a result of
the load connected to it producing an I-V characteristic represented by a straight line.
Ideal independent current sources can be connected together in parallel for circuit analysis techniques as
either parallel-aiding or parallel-opposing configurations, but they can not be connected together in series.
Also for solving circuit analysis and theorems, current sources become open-circuited sources to make their
current equal to zero. Note also that current sources are capable of either delivering or absorbing power.
In the case of non-ideal or practical current sources, they can be modelled as an equivalent ideal current
source and an internal parallel (shunt) connected resistance which is not infinite but of a value that is very
high as R ≈ ∞ producing an I-V characteristic which is not straight but slopes down as the load decreases.
We have also seen here that current sources can be dependent or independent. A dependent source is one
whose value depends on some other circuit variable. Voltage-controlled current source, VCCS, and current-
controlled current source, CCCS, are types of dependent current sources.
Constant current sources with very high internal resistances find numerous applications in electronic circuits
and analysis and can be built using bipolar transistors, diodes, zeners and FETs as well as a combination of
these solid-state devices.

Kirchhoff’s Current Law


Kirchhoff’s Current Law (KCL) is Kirchhoff’s first law that deals with the conservation of charge entering
and leaving a junction.
To determine the amount or magnitude of the electrical current flowing around an electrical or electronic
circuit, we need to use certain laws or rules that allows us to write down these currents in the form of an
equation. The network equations used are those according to Kirchhoff’s laws, and as we are dealing with
circuit currents, we will be looking at Kirchhoff’s current law, (KCL).
Gustav Kirchhoff’s Current Law is one of the fundamental laws used for circuit analysis. His current law
states that for a parallel path the total current entering a circuits junction is exactly equal to the total
current leaving the same junction. This is because it has no other place to go as no charge is lost.
In other words the algebraic sum of ALL the currents entering and leaving a junction must be equal to zero
as: Σ IIN = Σ IOUT.
This idea by Kirchhoff is commonly known as the Conservation of Charge, as the current is conserved
around the junction with no loss of current. Lets look at a simple example of Kirchhoff’s current law (KCL)
when applied to a single junction.
A Single Junction
Here in this simple single junction example, the current IT leaving the junction is the algebraic sum of the
two currents, I1 and I2 entering the same junction. That is IT = I1 + I2.
Note that we could also write this correctly as the algebraic sum of: IT - (I1 + I2) = 0.
So if I1 equals 3 amperes and I2 is equal to 2 amperes, then the total current, IT leaving the junction will be 3
+ 2 = 5 amperes, and we can use this basic law for any number of junctions or nodes as the sum of the
currents both entering and leaving will be the same.
Also, if we reversed the directions of the currents, the resulting equations would still hold true for I 1 or I2. As
I1 = IT - I2 = 5 - 2 = 3 amps, and I2 = IT - I1 = 5 - 3 = 2 amps. Thus we can think of the currents entering the
junction as being positive (+), while the ones leaving the junction as being negative (-).
Then we can see that the mathematical sum of the currents either entering or leaving the junction and in
whatever direction will always be equal to zero, and this forms the basis of Kirchhoff’s Junction Rule, more
commonly known as Kirchhoff’s Current Law, or (KCL).
Resistors in Parallel
Let’s look how we could apply Kirchhoff’s current law to resistors in parallel, whether the resistances in
those branches are equal or unequal. Consider the following circuit diagram:

 
In this simple parallel resistor example there are two distinct junctions for current. Junction one occurs at
node B, and junction two occurs at node E. Thus we can use Kirchhoff’s Junction Rule for the electrical
currents at both of these two distinct junctions, for those currents entering the junction and for those currents
flowing leaving the junction.
To start, all the current, IT leaves the 24 volt supply and arrives at point A and from there it enters node B.
Node B is a junction as the current can now split into two distinct directions, with some of the current
flowing downwards and through resistor R1 with the remainder continuing on through resistor R2 via node C.
Note that the currents flowing into and out of a node point are commonly called branch currents.
We can use Ohm’s Law to determine the individual branch currents through each resistor as: I = V/R, thus:
For current branch B to E through resistor R1

For current branch C to D through resistor R2

 
From above we know that Kirchhoff’s current law states that the sum of the currents entering a junction
must equal the sum of the currents leaving the junction, and in our simple example above, there is one
current, IT going into the junction at node B and two currents leaving the junction, I1 and I2.
Since we now know from calculation that the currents leaving the junction at node B is I 1 equals 3 amps and
I2 equals 2 amps, the sum of the currents entering the junction at node B must equal 3 + 2 = 5 amps. Thus ΣIN
= IT = 5 amperes.
In our example, we have two distinct junctions at node B and node E, thus we can confirm this value for I T
as the two currents recombine again at node E. So, for Kirchhoff’s junction rule to hold true, the sum of the
currents into point F must equal the sum of the currents flowing out of the junction at node E.
As the two currents entering junction E are 3 amps and 2 amps respectively, the sum of the currents entering
point F is therefore: 3 + 2 = 5 amperes. Thus Σ IN = IT = 5 amperes and therefore Kirchhoff’s current law
holds true as this is the same value as the current leaving point A.
Applying KCL to more complex circuits.
We can use Kirchhoff’s current law to find the currents flowing around more complex circuits. We
hopefully know by now that the algebraic sum of all the currents at a node (junction point) is equal to zero
and with this idea in mind, it is a simple case of determining the currents entering a node and those leaving
the node. Consider the circuit below.
Kirchhoff’s Current Law Example No1

 
In this example there are four distinct junctions for current to either separate or merge together at nodes A,
C, E and node F. The supply current IT separates at node A flowing through resistors R1 and R2, recombining
at node C before separating again through resistors R3, R4 and R5 and finally recombining once again at node
F.
But before we can calculate the individual currents flowing through each resistor branch, we must first
calculate the circuits total current, IT. Ohms law tells us that I = V/R and as we know the value of V, 132
volts, we need to calculate the circuit resistances as follows.
Circuit Resistance RAC
 
Thus the equivalent circuit resistance between nodes A and C is calculated as 1 Ohm.
Circuit Resistance RCF

 
Thus the equivalent circuit resistance between nodes C and F is calculated as 10 Ohms. Then the total circuit
current, IT is given as:

Giving us an equivalent circuit of:


Kirchhoff’s Current Law Equivalent Circuit

 
Therefore, V = 132V, RAC = 1Ω, RCF = 10Ω’s and IT = 12A.
Having established the equivalent parallel resistances and supply current, we can now calculate the
individual branch currents and confirm using Kirchhoff’s junction rule as follows.
 
Thus, I1 = 5A, I2 = 7A, I3 = 2A, I4 = 6A, and I5 = 4A.
We can confirm that Kirchoff’s current law holds true around the circuit by using node C as our reference
point to calculate the currents entering and leaving the junction as:

 
We can also double check to see if Kirchhoffs Current Law holds true as the currents entering the junction
are positive, while the ones leaving the junction are negative, thus the algebraic sum is: I 1 + I2 - I3 - I4 - I5 = 0
which equals 5 + 7 – 2 – 6 – 4 = 0.
So we can confirm by analysis that Kirchhoff’s current law (KCL) which states that the algebraic sum of the
currents at a junction point in a circuit network is always zero is true and correct in this example.
Kirchhoff’s Current Law Example No2
Find the currents flowing around the following circuit using Kirchhoff’s Current Law only.

 
IT is the total current flowing around the circuit driven by the 12V supply voltage. At point A, I 1 is equal to
IT, thus there will be an I1*R voltage drop across resistor R1.
The circuit has 2 branches, 3 nodes (B, C and D) and 2 independent loops, thus the I*R voltage drops around
the two loops will be:
 Loop ABC  ⇒  12 = 4I1 + 6I2
 Loop ABD  ⇒  12 = 4I1 + 12I3
Since Kirchhoff’s current law states that at node B, I1 = I2 + I3, we can therefore substitute current I1 for
(I2 + I3) in both of the following loop equations and then simplify.
Kirchhoff’s Loop Equations

 
We now have two simultaneous equations that relate to the currents flowing around the circuit.
Eq. No 1 :    12 = 10I2 + 4I3
Eq. No 2 :    12 = 4I2 + 16I3
By multiplying the first equation (Loop ABC) by 4 and subtracting Loop ABD from Loop ABC, we can be
reduced both equations to give us the values of I2 and I3
Eq. No 1 :    12 = 10I2 + 4I3 ( x4 )    ⇒   48 = 40I2 + 16I3
Eq. No 2 :    12 = 4I2 + 16I3 ( x1 )    ⇒   12 = 4I2 + 16I3
Eq. No 1 – Eq. No 2   ⇒   36 = 36I2 + 0
Substitution of I2 in terms of I3 gives us the value of I2 as 1.0 Amps
Now we can do the same procedure to find the value of I 3 by multiplying the first equation (Loop ABC) by 4
and the second equation (Loop ABD) by 10. Again by subtracting Loop ABC from Loop ABD, we can be
reduced both equations to give us the values of I2 and I3
Eq. No 1 :    12 = 10I2 + 4I3 ( x4 )    ⇒   48 = 40I2 + 16I3
Eq. No 2 :    12 = 4I2 + 16I3 ( x10 )    ⇒   120 = 40I2 + 160I3
Eq. No 2 – Eq. No 1   ⇒   72 = 0 + 144I3
Thus substitution of I3 in terms of I2 gives us the value of I3 as 0.5 Amps
As Kirchhoff’s junction rule states that :  I1 = I2 + I3
The supply current flowing through resistor R1 is given as :  1.0 + 0.5 = 1.5 Amps
Thus I1 = IT = 1.5 Amps, I2 = 1.0 Amps and I3 = 0.5 Amps and from that information we could calculate the
I*R voltage drops across the devices and at the various points (nodes) around the circuit.
We could have solved the circuit of example two simply and easily just using Ohm’s Law, but we have used
Kirchhoff’s Current Law here to show how it is possible to solve more complex circuits when we can not
just simply apply Ohm’s Law.

Kirchhoff’s Voltage Law


Kirchhoff’s Voltage Law (KVL) is Kirchhoff’s second law that deals with the conservation of energy
around a closed circuit path.
Gustav Kirchhoff’s Voltage Law is the second of his fundamental laws we can use for circuit analysis. His
voltage law states that for a closed loop series path the algebraic sum of all the voltages around any
closed loop in a circuit is equal to zero. This is because a circuit loop is a closed conducting path so no
energy is lost.
In other words the algebraic sum of ALL the potential differences around the loop must be equal to zero as:
ΣV = 0. Note here that the term “algebraic sum” means to take into account the polarities and signs of the
sources and voltage drops around the loop.
This idea by Kirchhoff is commonly known as the Conservation of Energy, as moving around a closed
loop, or circuit, you will end up back to where you started in the circuit and therefore back to the same initial
potential with no loss of voltage around the loop. Hence any voltage drops around the loop must be equal to
any voltage sources met along the way.
So when applying Kirchhoff’s voltage law to a specific circuit element, it is important that we pay special
attention to the algebraic signs, (+ and -) of the voltage drops across elements and the emf’s of sources
otherwise our calculations may be wrong.
But before we look more closely at Kirchhoff’s voltage law (KVL) lets first understand the voltage drop
across a single element such as a resistor.
A Single Circuit Element

For this simple example we will assume that the current, I is in the same direction as the flow of positive
charge, that is conventional current flow.
Here the flow of current through the resistor is from point A to point B, that is from positive terminal to a
negative terminal. Thus as we are travelling in the same direction as current flow, there will be a fall in
potential across the resistive element giving rise to a -IR voltage drop across it.
If the flow of current was in the opposite direction from point B to point A, then there would be a rise in
potential across the resistive element as we are moving from a - potential to a + potential giving us a +I*R
voltage drop.
Thus to apply Kirchhoff’s voltage law correctly to a circuit, we must first understand the direction of the
polarity and as we can see, the sign of the voltage drop across the resistive element will depend on the
direction of the current flowing through it. As a general rule, you will loose potential in the same direction
of current across an element and gain potential as you move in the direction of an emf source.
The direction of current flow around a closed circuit can be assumed to be either clockwise or anticlockwise
and either one can be chosen. If the direction chosen is different from the actual direction of current flow,
the result will still be correct and valid but will result in the algebraic answer having a minus sign.
To understand this idea a little more, lets look at a single circuit loop to see if Kirchhoff’s Voltage Law
holds true.
A Single Circuit Loop

 
Kirchhoff’s voltage law states that the algebraic sum of the potential differences in any loop must be equal
to zero as: ΣV = 0. Since the two resistors, R1 and R2 are wired together in a series connection, they are both
part of the same loop so the same current must flow through each resistor.
Thus the voltage drop across resistor, R 1 = I*R1 and the voltage drop across resistor, R 2 = I*R2 giving by
KVL:
 
We can see that applying Kirchhoff’s Voltage Law to this single closed loop produces the formula for the
equivalent or total resistance in the series circuit and we can expand on this to find the values of the voltage
drops around the loop.

Kirchhoff’s Voltage Law Example No1


Three resistor of values: 10 ohms, 20 ohms and 30 ohms, respectively are connected in series across a 12
volt battery supply. Calculate: a) the total resistance, b) the circuit current, c) the current through each
resistor, d) the voltage drop across each resistor, e) verify that Kirchhoff’s voltage law, KVL holds true.
a) Total Resistance (RT)
RT = R1 + R2 + R3  =  10Ω + 20Ω + 30Ω = 60Ω
Then the total circuit resistance RT is equal to 60Ω
b) Circuit Current (I)

Thus the total circuit current I is equal to 0.2 amperes or 200mA


c) Current Through Each Resistor
The resistors are wired together in series, they are all part of the same loop and therefore each experience the
same amount of current. Thus:
IR1 = IR2 = IR3 = ISERIES  =  0.2 amperes
d) Voltage Drop Across Each Resistor
VR1 = I x R1 = 0.2 x 10  =  2 volts
VR2 = I x R2 = 0.2 x 20  =  4 volts
VR3 = I x R3 = 0.2 x 30  =  6 volts
e) Verify Kirchhoff’s Voltage Law
Thus Kirchhoff’s voltage law holds true as the individual voltage drops around the closed loop add up to the
total.
Kirchhoff’s Circuit Loop

 
We have seen here that Kirchhoff’s voltage law, KVL is Kirchhoff’s second law and states that the algebraic
sum of all the voltage drops, as you go around a closed circuit from some fixed point and return back to the
same point, and taking polarity into account, is always zero. That is ΣV = 0
The theory behind Kirchhoff’s second law is also known as the law of conservation of voltage, and this is
particularly useful for us when dealing with series circuits, as series circuits also act as voltage dividers and
the voltage divider circuit is an important application of many series circuits.

Voltage Dividers
Voltage Divider circuits are used to produce different voltage levels from a common voltage source but the
current is the same for all components in a series cicruit
Voltage Divider Circuits are useful in providing different voltage levels from a common supply voltage.
This common supply can be a single supply either positive or negative, for example, +5V, +12V, -5V or
-12V, etc. with respect to a common point or ground, usually 0V, or it could be across a dual supply, for
example ±5V, or ±12V, etc.
Voltage dividers are also known as potential dividers, because the unit of voltage, the “Volt” represents the
amount of potential difference between two points. A voltage or potential divider is a simple passive circuit
that takes advantage of the effect of voltages being dropped across components which are connected in
series.
The potentiometer, which is a variable resistor with a sliding contact, is the most basic example of a voltage
divider as we can apply a voltage across its terminals and produce an output voltage in proportion to the
mechanical position of its sliding contact. But we can also make voltage dividers using individual resistors,
capacitors and inductors as they are two-terminal components which can be connected together in series.
Resistive Voltage Divider
The simplest, easiest to understand, and most basic form of a passive voltage divider network is that of two
resistors connected together in series. This basic combination allows us to use the Voltage Divider Rule to
calculate the voltage drops across each series resistor.
Resistive Voltage Divider Circuit

Here the circuit consists of two resistors connected together in series: R 1, and R2. Since the two resistors are
connected in series, it must therefore follow that the same value of electric current must flow through each
resistive element of the circuit as it has nowhere else to go. Thus providing an I*R voltage drop across each
resistive element.
With a supply or source voltage, VS applied across this series combination, we can apply Kirchhoff’s
Voltage Law, (KVL) and also using Ohm’s Law to find the voltage dropped across each resistor derived in
terms of the common current, I flowing through them. So solving for the current (I) flowing through the
series network gives us:

 
The current flowing through the series network is simply I = V/R following Ohm’s Law. Since the current is
common to both resistors, (IR1 = IR2) we can calculate the voltage dropped across resistor, R 2 in the above
series circuit as being:
 
Likewise for resistor R1 as being:

Voltage Divider Example No1


How much current will flow through a 20Ω resistor connected in series with a 40Ω resistor when the supply
voltage across the series combination is 12 volts dc. Also calculate the voltage drop produced across each
resistor.

 
Each resistance provides an I*R voltage drop which is proportionaly equal to its resistive value across the
supply voltage. Using the voltage divider ratio rule, we can see that the largest resistor produces the largest
I*R voltage drop. Thus, R1 = 4V and R2 = 8V. Applying Kirchhoff’s Voltage Law shows that the sum of the
voltage drops around the resistive circuit is exactly equal to the supply voltage, as 4V + 8V = 12V.
Note that if we use two resistors of equal value, that is R 1 = R2, then the voltage dropped across each resistor
would be exactly half the supply voltage for two resistances in series as the voltage divider ratio would equal
50%.
Another use of a voltage divider network is that to produce a variable voltage output. If we replace resistor
R2 with a variable resistor (potentiometer), then the voltage dropped across R 2 and therefore VOUT can be
controlled by an amount dependant on the postion of the potentiometers wiper and therefore the ratio of the
two resistive values as we have one fixed and one variable resistor. Potentiometers, trimmers, rheostats and
variacs are all examples of variable voltage division devices.
We could also take this idea of variable voltage division one step further by replacing the fixed resistor R 2
with a sensor such as a light dependent resistor, or LDR. Thus as the resistive value of the sensor changes
with changes in light levels, the output voltage VOUT also changes by a proportional amount. Thermistors and
strain guages are other examples of resistive sensors.
Since the two voltage division expressions above relate to the same common current, mathematically they
must therefore be related to each other. So for any number of individual resistors forming a series network,
the voltage dropped across any given resistor is given as:
Voltage Divider Equation

 
Where: VR(x) is the voltage drop across the resistor, RX is the value of the resistor, and RT is the total
resistance of the series network. This voltage divider equation can be used for any number of series
resistances connected together because of the proportional relationship between each resistance, R and its
corresponding voltage drop, V. Note however, that this equation is given for an unloaded voltage divider
network without any additional resistive load connected or parallel branch currents.
Voltage Divider Example No2
Three resistive elements of 6kΩ, 12kΩ and 18kΩ are connected together in series across a 36 volt supply.
Calculate, the total resistance, the value of the current flowing around the circuit, and the voltage drops
across each resistor.
Data given: VS = 36 volts, R1 = 6kΩ, R2 = 12kΩ and R3 = 18kΩ

Voltage Divider Circuit


 
The voltage drops across all three resistors should add up to the supply voltage as defined by Kirchhoff’s
Voltage Law (KVL). So the sum of the voltage drops is: V T = 6 V + 12 V + 18 V = 36.0 V the same value of
the supply voltage, VS and so is correct. Again notice that the largest resistor produces the largest voltage
drop.
Voltage Tapping Points in a Divider Network

Consider a long series of resistors connected to a voltage source, V S. Along the series network there are
different voltage tapping points, A, B, C, D, and E.
The total series resistance can be found by simply adding together the individual series resistance values
giving a total resistance, RT value of 15kΩ. This resistive value will limit the flow of current through the
circuit produced by the supply voltage, VS.
The individual voltage drops across the resistors are found using the equations above, so V R1 = VAB, VR2 =
VBC, VR3 = VCD, and VR4 = VDE.
The voltage levels at each tapping point is measured with respect to ground (0V). Thus the voltage level at
point D will be equal to VDE, and the voltage level at point C will be equal to VCD + VDE. In other words, the
voltage at point C is the sum of the two voltage drops across R3 and R4.
So hopefully we can see that by choosing a suitable set of resistive values, we can produce a sequence of
voltage drops which will have a proportional voltage value obtained from a single supply volatge. Note also
that in this example each output voltage point will be positive in value because the negative terminal of the
voltage supply, VS is grounded.
Voltage Divider Example No3
1. Calculate the noload voltage output for each tapping point of the voltage divider circuit above if the
series-connected resistive network is connected to a 15 volt DC supply.

 
2. Calculate the noload voltage output from between points B and E.

 
A Negative and Positive Voltage Divider
In the simple voltage divider circuit above all the output voltages are referenced from a common zero-
voltage ground point, but sometimes it is necessary to produce both positive and negative voltages from a
single source voltage supply. For example the different voltage levels from a computer PSU, -12V, +3.3V,
+5V and +12V, with respect to a common reference ground terminal.
Voltage Divider Example No4
Using Ohm’s Law, find the values of resistors R 1, R2, R3 and R4 required to produce the voltage levels of
-12V, +3.3V, +5V and +12V if the total power supplied to the unloaded voltage divider circuit is 24 volts
DC, 60 watts.

 
In this example, the zero-voltage ground reference point has been moved to produce the required positive
and negative voltages, while maintaining the voltage divider network across the supply. Thus the four
voltages are all measured with respect to this common reference point reulting in point D being at the
required negative potential of -12V with respect to ground.
We have seen so far that series resistive circuits can be used to create a voltage divider, or potential divider
network which can be widely used in electronic circuits. By selecting appropriate values for the series
resistances, any value of output voltage can be obtained which is lower than the input or supply voltage. But
as well as using resistances and a DC supply voltage to create a resistive voltage divider network, we can
also use capacitors (C) and inductors (L), but with a sinusoidal AC supply as capacitors and inductors are
reactive components, meaning that their resistance “reacts” against the flow of electric current.
Capacitive Voltage Dividers
As the name suggests, Capacitive Voltage Divider circuits produce voltage drops across capacitors
connected in series to a common AC supply. Generally capacitive voltage dividers are used to “step-down”
very high voltages to provide a low voltage output signal which can then be used for protection or metering.
Nowadays, high frequency capacitive voltage dividers are used more in display devices and touch screen
technologies found in mobile phones and tablets.
Unlike resistive voltage divider circuits which operate on both AC and DC supplies, voltage division using
capacitors is only possible with a sinusoidal AC supply. This is because the voltage division between series
connected capacitors is calculated using the reactance of the capacitors, XC which is dependent on the
frequency of the AC supply.
We remember from our tutorials about capacitors in AC circuits, that capacitive reactance, XC (measured in
Ohms) is inversely proportional to both frequency and capacitance, and is therefore given by the following
equation of:
Capacitive Reactance Formula
 Where:
    Xc = Capacitive Reactance in Ohms, (Ω)
    π (pi) = a numeric constant of 3.142
    ƒ = Frequency in Hertz, (Hz)
    C = Capacitance in Farads, (F)
Therefore by knowing the voltage and frequency of the AC supply, we can calculate the reactances of the
individual capacitors, substitute them in the above equation for the resistive voltage divider rule, and obtain
the corresponding voltage drops across each capacitor as shown.
Capacitive Voltage Divider

 
Using the two capacitors of 10uF and 22uF in the series circuit above, we can calculate the rms voltage
drops across each capacitor in terms of their reactance when connected to a 100 volts, 50Hz rms supply.

 
When using pure capacitors the sum of all the series voltage drops equals the source voltage, the same as for
series resistances. While the amount of voltage drop across each capacitors is proportional to its reactance, it
is inversely proportional to its capacitance.
As a result, the smaller 10uF capacitor has more reactance (318.3Ω) so therefore a greater voltage drop of 69
volts compared to the larger 22uF capacitor which has a reactance of 144.7Ω and a voltage drop of 31 volts
respectively. The current in the series circuit, IC will be 216mA, and is the same value for C1 and C2 as they
are in series.
One final point about capacitive voltage divider circuits is that as long as there is no series resistance,
purely capacitive, the two capacitor voltage drops of 69 and 32 volts will arithmetically be equal to the
supply voltage of 100 volts as the two voltages produced by the capacitors are in-phase with each other. If
for whatever reason the two voltages are out-of-phase with each other then we can not just simple add them
together as we would using Kirchhoffs voltage law, but instead phasor addition of the two waveforms is
required.
Inductive Voltage Dividers
As its name suggests, Inductive Voltage Dividers create voltage drops across inductors or coils connected
together in series to a common AC supply. An inductive voltage divider can consist of a single winding or
coil which is divided into two sections where the output voltage is taken from across one of the section, or
from two individual coils connected together. The most common example of an inductive voltage divider is
the auto-transformer with multiple tapping points along its secondary winding.
When used with steady state DC supplies or with sinusoids having a very low frequency, approaching 0 Hz,
inductors act as a short circuit. This is because their reactance is almost zero allowing any DC current to
easily pass through them, so like the previous capacitive voltage divider network, we must perform any
inductive voltage division using a sinusoidal AC supply. Inductive voltage division between series
connected inductors can be calculated using the reactance of the inductors, X L which like capacitive
inductance, is dependent on the frequency of the AC supply.
In the tutorials about inductors in AC circuits, we saw that inductive reactance, XL (also measured in Ohms)
is proportional to both frequency and inductance so any increases in the supply frequency increases an
inductors reactance. Thus inductive reactance is defined as:
Inductive Reactance Formula

 Where:
    XL = Inductive Reactance in Ohms, (Ω)
    π (pi) = a numeric constant of 3.142
    ƒ = Frequency in Hertz, (Hz)
    L = Inductance in Henries, (H)
If we know the voltage and frequency of the AC supply, we can calculate the reactances of the two inductors
and use them along with the voltage divider rule to obtain the voltage drops across each inductor as shown.
Inductive Voltage Divider
 
Using the two inductors of 10mH and 20mH in the series circuit above, we can calculate the rms voltage
drops across each capacitor in terms of their reactance when connected to a 60 volts, 200Hz rms supply.

 
Like the previous resistive and capacitive voltage division circuits, the sum of all the series voltage drops
across the inductors will equal the source voltage, as long as there are no series resistances. Meaning a pure
inductor. The amount of voltage drop across each inductor is proportional to its reactance.
The result is that the smaller 10mH inductor has less reactance (12.56Ω), so therefore less of a voltage drop
at 30 volts compared to the larger 20mH inductor which has a reactance of 25.14Ω and a voltage drop of 40
volts respectively. The current, IL in the series circuit is 1.6mA, and will be the same value for L 1 and L2 as
these two inductors are connected in series.
Voltage Divider Summary
We have seen here that the voltage divider, or network is a very common and useful circuit configuration
allowing us to produce different voltage levels from a single voltage supply, thus eliminating the need to
have separate power supplies for different parts of a circuit operating at different voltage levels.
As its name suggests, a voltage or potential divider, “divides” a fixed voltage into precise proportions using
resistors, capacitors or inductors. The most basic and commonly used voltage divider circuit is that of two
fixed-value series resistors, but a potentiometer or rheostat can also be used for voltage division by simply
adjusting its wiper position.
A very common application of a voltage divider circuit is to replace one of the fixed-value resistors with a
sensor. Resistive sensores such as light sensores, temperature sensores, pressure sensores and strain guages,
which change their resistive value as they respond to environmental changes can all be used in a voltage
divider network to provide an analogue voltage output. The biasing of bipolar transistors and MOSFETs is
also another common application of a Voltage Divider.

Current Dividers
Current Divider circuits have two or more parallel branches for currents to flow through but the voltage is
the same for all components in the parallel cicruit
Current Divider Circuits are parallel circuits in which the source or supply current divides into a number
of parallel paths. In a parallel connected circuit, all the components have their terminals connected together
sharing the same two end nodes. This results in different paths and branches for the current to flow or pass
along. However, the currents can have different values through each component.
The main characteristic of parallel circuits is that while they may produce different currents flowing through
different branches, the voltage is common to all the connected paths. That is V R1 = VR2 = VR3 … etc.
Therefore the need to find the individual resistor voltages is eliminated allowing branch currents to be easily
found with Kirchhoff’s Current Law, (KCL) and of course Ohm’s Law.
Resistive Voltage Divider
The easiest to understand, and most basic form of a passive current divider network is that of two resistors
connected together in parallel. The Current Divider Rule allows us to calculate the current flowing through
each parallel resistive branch as a percentage of the total current. Consider the circuit below.
Resistive Current Divider Circuit

Here this basic current divider circuit consists of two resistors: R1, and R2 in parallel which splits the supply
or source current IS between them into two separate currents I R1 and IR2 before joining together again and
returning back to the source.
As the source or total current equals the sum of the individual branch currents, then the total current, I T
flowing in the circuit is given by Kirchoffs current law KCL as being:
IT = IR1 + IR2
As the two resistors are connected in parallel, for Kirchhoff’s Current Law, (KCL) to hold true it must
therefore follow that the current flowing through resistor R1 will be equal to:
IR1 = IT – IR2
and the current flowing through resistor R2 will be equal to:
IR2 = IT – IR1
As the same voltage, (V) is present across each resistive element, we can find the current flowing through
each resistor in terms of this common voltage as it is simply V = I*R following Ohm’s Law. So solving for
the voltage (V) across the parallel combination gives us:

 
Solving for IR1 gives:

 
Likewise, solving for IR2 gives:
 
Notice that the above equations for each branch current has the opposite resistor in its numerator. That is to
solve for I1 we use R2, and to solve for I2 we use R1. This is because each branch current is inversely
proportional to its resistance resulting in the smaller resistance having the larger current.
Current Divider Example No1
A 20Ω resistor is connected in parallel with a 60Ω resistor. If the combination is connected across a 30 volts
battery supply, find the current flowing through each resistor and the total current supplied by the source.

 
Note that the smaller 20Ω resistor has the larger current because by its very nature, current will always flow
through the path or branch of least resistance. This implies then that a short-circuit will produce maximum
current flow, while an open-circuit will result in zero current flow. Remember also that the equivalent
resistance, REQ of parallel connected resistors will always be less than the ohmic value of the smallest
resistor with the equivalent resistance decreasing as more parallel resistances are added.
Sometimes it is not necessary to calculate all the branch currents, if the supply or total current, I T is already
known, then the final branch current can be found by simply subtracting the calculated currents from the
total current as defined by Kirchhoffs current law.
Current Divider Example No2
Three resistors are connected together to form a current divider circuit as shown below. If the circuit is fed
from a 100 volts 1.5kW power supply, calculate the individual branch currents using the current division
rule and the equivalent circuit resistance.

 
1)  Total circuit current IT
 
2)  Equivalent resistance REQ

 
3)  Branch currents IR1, IR2, IR3

 
We can check our calculations as according to Kirchhoff’s Current Rule, all the branch currents will be
equal to the total current, so: I T = IR1 + IR2 + IR3 = 10 + 4 + 1 = 15 amperes, as expected. Thus we can see that
the total current, IT is divided according to a simple ratio determined by the branch resistances. Also, as the
number of resistors connected in parallel increases, the supply ot total current, I T will also increase for a
given supply voltage, VS as there are more parallel branches taking current.
Current Division using Conductances
Another simple method of finding the branch currents in a parallel circuit is to use the conductance method.
In DC circuits, Conductance is the reciprocal of resistance, and is denoted by the letter “G“. As
conductance (G) is the reciprocal of resistance (R) which is measured in Ohm’s (Ω), the reciprocal of Ohm’s
is called “mho” (℧), (an inverted ohm sign). Thus G = 1/R. The electrical units given to conductance is the
Siemen (symbol S).
So for parallel connected resistors, the equavalent or total conductance, CT will be equal to the sum of the
individual conductances as shown.
Parallel Conductance

 
Therefore, if a resistance has a fixed value of 10Ω, it will have an equivalent conductance of 0.1S and so on.
Because of the reciprocal, a high value of conductance represents a low value of resistance, and vice versa.
We can also use prefixes in the form of milli-Siemens, mS, micro-Siemens, uS and even nano-Siemens, nS
for very small conductances. So a resistor of 10kΩ will have a conductance of 100uS.
Using the Ohm’s Law equation for current in which I = V/R, we can define the branch currents using
conductance as being: I = V*G
In fact we can take this one step further by saying that the supply current to a our parallel resisive network
above is:

 
But we know from above that for a parallel connected circuit, voltage is common to all components and as
voltage equals current times resistance, V = I*R, we can therefore conclude that when using conductance,
the voltage is equal to current divided by conductance. That is V = I/G.
Then we can express the above equations for the current divider rule in relationship to conductance (G),
instead of the resistance (R) as being:
Current Divider Rule using Conductance
 
Likewise for the currents in parallel resistors R2 and R3 are given as:

 
You may have noticed that unlike the equations above for resistance, each branch current has the same
conductance in its numerator. That is to solve for I 1 we use G1, and to solve for I2 we use G2. This is because
the conductances are the reciprocals of the resistances.
Current Divider Example No3
Using the conductance method, find the individual branch currents, I1, I2 and I3 of the following parallel
resistive circuit.

 
Total conductance GT

 
Total supply current IS
 
Individual branch currents I1, I2 and I3

 
As conductance is the reciprocal or inverse of resistance, the equivalent resistance value of the example
circuit is simply 1/800uS which equals 1250Ω or 1.25kΩ, which is clearly less than the smallest resistor
value of R1 at 2kΩ.
Current Divider Summary
Current dividers or current division is the process of finding the individual branch currents in a parallel
circuit were each parallel element has the same voltage. Kirchhoff’s current law, (KCL) states that the
algebraic sum of the individual currents entering a junction or node will equal the currents leaving it. That is
the net result is zero.
Kirchhoff’s current divider rule can also be used to find individual branch currents when the equivalent
resistance and the total circuit current are known. When only two resistive branches are involved, the current
in one branch will be some fraction of the total current I T. If the two parallel resistive branches are of equal
value, the current will divide equally.
In the case of three or more parallel branches, the equivalent resistance R EQ is used to divide the total current
into the fractional currents for each branch producing a current ratio which is equal to the inverse of their
resistive values resulting in the smaller value resistance having the greatest share of the current. The supply
or total current, IT being the sum of all the individual branch currents. This then makes current dividers
useful for use with current sources.
It is sometimes convenient to use conductance with parallel circuits as it can help reduce the maths required
for determining the branch currents through individual circuit elements that are connected together in
parallel. This is because for parallel circuits the total conductance is the sum of the individual conductance
values. Conductance is the reciprocal or inverse of resistance as G = 1/R. The units for conductance are
Siemens, S. The conductance of an element can also be used even if the supply voltage is DC or AC for
current dividers.

Electrical Energy and Power


Electrical Energy supplies the power required to produce work or an action within an electrical circuit and is
given in joules per second
Electrical Energy is the ability of an electrical circuit to produce work by creating an action. This action
can take many forms, such as thermal, electromagnetic, mechanical, electrical, etc. Electrical energy can be
both created from batteries, generators, dynamos, and photovoltaics, etc. or stored for future use using fuel
cells, batteries, capacitors or magnetic fields, etc. Thus electrical energy can be either created or stored.
We remember from our school science classes that the “The Law of the Conservation of Energy” states that
energy cannot be created or destroyed, only converted. But for energy to do any useful work it must be
converted from one form into something else. For example, a motor converts electrical energy into
mechanical or kinetic (rotational) energy, while a generator converts kinetic energy back into electrical
energy to power a circuit.
That is electrical machines convert or change energy from one form to another by doing work. Another
example is a lamp, light bulb or LED (light emitting diode) which convert electrical energy into light energy
and heat (thermal) energy. Then electrical energy is very versatile as it can be easily converted into many
other different forms of energy.
For electrical energy to move electrons and produce a flow of current around a circuit, work must be done,
that is the electrons must move by some distance through a wire or conductor. The work done is stored in
the flow of electrons as energy. Thus “Work” is the name we give to the process of energy.
We can therefore say that Work and Energy are effectively the same as energy can be defined as “the ability
to do some work”. Note that work done or energy transferred applies equally to a mechanical system or
thermal system as it does to an electrical system. This is because because mechanical, thermal and electrical
energies are interchangeable.
Electrical Energy: The Volt
As we now know that energy is the capacity to do work, with the standard unit used for energy (and work)
being the Joule. A joule of energy is defined as the energy expended by one ampere at one volt, moving in
one second. Electric current results from the movement of electric charge (electrons) around a circuit, but to
move charge from one node to another there needs to be a force to create the work to move the charge, and
there is: voltage.
We tend to think of voltage (V) as existing between two different terminals, points or nodes within a circuit
or battery supply. But voltage is important as it provides the work needed to move the charge from one point
to another, either in a forward direction or a reverse direction. The voltage, or potential difference between
two terminals or points is defined as having a value of one volt, when one joule of energy is used in moving
one coulomb of electric charge between those two terminals.
In other words, the Voltage difference between two points or terminals is the work required in Joules to
move one Coulomb of charge from A to B. Therefore voltage can be expressed as being:
The Voltage Unit

 
Where: voltage is in Volts, J is the work or energy in Joules and C is the charge in Coulombs. Thus if J = 1
joule, C = 1 coulomb, then V will equal 1 volt.
Electrical Energy Example No1
What is the terminal voltage of a battery that expends 135 joules of energy to move 15 coulombs of charge
around an electrical circuit.

 
Then we can see in this example that every coulomb of charge possesses an energy of 9 joules.
Electrical Energy: The Ampere
We have seen that the unit of electrical charge is the Coulomb and that the flow of electrical charge around a
circuit is used to represent a flow of current. However, as the symbol for a coulomb is the letter “C“, this can
be confused with the symbol for Capacitance, which is also the letter “C“.
To avoid this confusion, the common symbol used for electrical charge is the capital letter “Q” or small
letter “q“, basically standing for quantity. Thus Q = 1 coulomb of charge or Q = 1C. Note that charge Q can
be either positive, +Q or negative, -Q, that is an excess of either electrons or holes.
The flow of charge around a closed circuit in the form of electrons is called an electric current. However,
the use of the expression “flow of charge” implies movement, so to produce an electrical current, charge
must move. This then leads to the question of what is making the charge move, and this is done by our old
friend Voltage from above.
So the voltage or potential difference between two points provides the required electrical energy to move
charge around a circuit in the form of an electric curent. Therefore the work done to move charge is
provided by a potential difference, and if there is no potential difference between two points, there is no
movement of charge and therefore no current flow. Infact charge without any flow or movement is called
static electricity.
If the movement of charge is called an electric current, then we can correctly say that current is the rate of
movement (or rate of flow) of the charge, but how much charge represents a current. If we select a point
within a circuit, any point, and measure the amount of charge that flows past this point in exactly one
second, this will give us the strength of the electrical current in Amperes, (A).
Thus one ampere of current is equal to one coulomb of charge which flows past a given point in one unit
second, and the more charge per second which passes this point, the greater will be the current. Then we can
define one ampere (A) of electrical current as being equal to one coulomb of charge per second. So 1A =
1C/s
The Ampere Unit

 
Where: Q is the charge (in coulombs) and t is the interval in time (in seconds) that the charge moves. In
other words, electrical current has both a magnitude (the amount of charge) and a specified direction
associated with it.
Note that the commonly used symbol for electrical current is the capital letter “I“, or small “i” both standing
for intensity. That is the intensity or concentration of charge producing the electron flow. For a constant DC
current, the capital letter “I” is generally used, whereas for a time-varying AC current the lower case letter
“i” is commonly used. The symbol i(t) means an instantaneous current value at that exact instant in time.
It is sometimes easier to remember this relationship by using an image. Here the three quantities of Q, I and t
have been superimposed into a triangle represents the actual position of each quantity within the current
formula.
The Ampere

 
Transposing the standard formula above gives us the following combinations of the same equation:

Electrical Energy Example No2


1. How much current flows through a circuit if 900 coulombs of charge passes a given point in 3 minutes.

 
2. An electric current of 3 Amperes flows through a resistor. How many coulombs of charge will flow
through the resistor in 90 seconds.

Electrical Energy: The Watt


Electrical Power is the product of the two quantities, Voltage and Current and so can be defined as the rate
at which work is performed in expending energy. We said previously that voltage provides the work
required in Joules to move one Coulomb of charge from A to B and that current is the rate of movement (or
rate of flow) of the charge. So how are these two definitions linked together.
If voltage, (V) equals Joules per Coulombs (V = J/C) and Amperes (I) equals charge (coulombs) per second
(A = Q/t), then we can define electrical power (P) as being the sum of these two quantities. This is because
electrical power can also equal voltage times amperes, that is: P = V*I.
The Watt

 
So we can see that electrical power is also the rate at which work is performed during one second. That is,
one joule of energy dissipated in one second. As electrical power is measured in Watts (W), therefore it must
be also be measured in Joules per Second. So we can correctly say that: 1 watt = 1 joule per second (J/s).
Electrical Power
1 watt (W) = 1 joule/second (J/s)
 
So if 1 watt = 1 joule/second, it therefore follows that: 1 Joule of energy = 1 watt per second, that is: Work
equals power times time. So electrical energy (the work done) is obtained by multiplying power by the time
in seconds that the charge (in the form of a current) flows. Thus units of electrical energy depend on the
units used for electric power and time. So if we measure electrical power in kilowatts (kW), and the time in
hours (h), then the electrical energy consumed equals kilowatts*hours or simply: kilowatt-hours.
Electrical Energy Example No3
A 100 Watt light bulb is illuminated on for one hour only. How many joules of electrical energy have been
used by the lamp.

 
Note that when dealing with the joule as a unit of electrical energy, it is more convenient to present them in
kilo-joules. Thus the answer can be given as: 360kJ. As a joule on its own is a small quantity, the kilojoule
(kJ), thousands of joules, the megajoule (MJ), millions of joules, and even the gigajoule (GJ), thousands of
millions of joules, are all practical units of electrical energy. Thus one unit of electricity which is one
kilowatt-hour (kWh) is equivalent to 3.6 megajoules (MJ).
Likewise, since a Watt is such a small amount of electrical power, kilowatts (1 kW = 1,000 watts) and
megawatts (1 MW = 1 million watts) are commonly used to identify the power output of electrical
equipment and appliances. Thus we can see that the kilowatt (or megawatt) is a unit of electrical power,
while the kilowatt-hour is a unit of electrical energy.

Potrebbero piacerti anche