Sunteți pe pagina 1din 56

Voltage, Current, and Resistance

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. The difference in voltage between any two nodes in a
circuit is known as the Potential Difference, p.d. sometimes called Voltage Drop.

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". 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 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. The diagram at the left
shows the movement of the positive charge (holes) which flows from the positive
terminal of the battery, through the circuit and returns to the negative terminal of the
battery. This was the convention chosen during the
discovery of electricity in which the direction of electric
current was thought to flow in a circuit. In circuit diagrams,
the arrows shown on symbols for components such as
diodes and transistors point in the direction of conventional
current flow. Conventional Current Flow is the opposite
in direction to the flow of electrons.

Electron Flow

The flow of electrons around the circuit is opposite to the direction of the conventional
current flow. The current flowing in a circuit is composed of
electrons that flow from the negative pole of the battery (the
cathode) and return to the positive pole (the anode). This is
because the charge on an electron is negative by definition
and so is attracted to the positive terminal. The flow of
electrons is called Electron Current Flow. Therefore,
electrons flow 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.

Resistance
The Resistance of a circuit is its ability to resist or prevent the flow of current (electron
flow) through it making it necessary to apply a bigger voltage to the circuit to cause the
current to flow again. Resistance is measured in Ohms, Greek symbol ( Ω, Omega ) with
prefixes used to denote Kilo-ohms (kΩ = 103Ω) and Mega-ohms (MΩ = 106Ω).
Resistance cannot be negative only positive.

Resistor Symbols
The amount of resistance determines whether the circuit 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.
Semiconductors are used to make Diodes and Transistors etc.

The electrical resistance of an object is a measure of its opposition to the passage of an


electric current. An object of uniform cross section will have a resistance proportional to
its resistivity and length and inversely proportional to its cross-sectional area.

Discovered by Georg Ohm in 1827,[1] electrical resistance shares some conceptual


parallels with the mechanical notion of friction. The SI unit of electrical resistance is the
ohm (Ω). Resistance's reciprocal quantity is electrical conductance measured in siemens.

The resistance of an object can be defined as the ratio of voltage to current:

R = {V \over I}

Electrical resistance is a measure of the degree to which an electrical component opposes


the passage of current. It is the ratio of the potential difference (i.e. voltage) across an
electric component (such as a resistor) to the current passing through that component:

<math>R = \frac{V}{I}<math>

where

R is the resistance of the component.

V the voltage across the component, measured in volts


I is the current passing through the component, measured in amperes

Here are the standard units of measurement for electrical current, voltage, and resistance:

Relationship between Voltage and Current in a circuit of constant resistance.

Summary
Hopefully by know you have an idea of how voltage, current and resistance are related.
The relationship between Voltage, Current and Resistance forms the basis of Ohm's law
which in a linear circuit states that if we increase the voltage, the current goes up and if
we increase the resistance, the current goes down. 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 D.C. 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 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".
• The effective (rms) value of an AC current has the same average power
loss equivalent to a DC 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.
• 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 Amp A
Resistance R Ohms Ω

In the next tutorial about DC Theory 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: E = I x R.

Ohms Law
The relationship between Voltage, Current and Resistance in any DC electrical circuit
was firstly discovered by the German physicist Georg Ohm, (1787 - 1854). 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 bases 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 Voltage (V)

[V = I x R] V (volts) = I (amps) x R (Ω)

To find Current (I)

[I = V ÷ R] I (amps) = V (volts) ÷ R (Ω)

To find Resistance (R)

[R = V ÷ I] R (Ω) = V (volts) ÷ I (amps)

It is sometimes easier to remember 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 at
the bottom. This arrangement represents the actual position of each quantity in the Ohms
law formulas.

Ohms Law Triangle


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, the less current will flow for
any 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.

Power in Electrical Circuits


Electrical Power, (P) in a circuit is the amount of energy that is absorbed or produced
within the circuit. A source of energy such as a voltage will produce or deliver power
while the connected load absorbs it. 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) with prefixes used to denote milliwatts (mW = 10-3W) or kilowatts (kW = 103W).
By using Ohm's law and substituting for V, I and R the formula for electrical power can
be found as:

To find 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 the
Power Triangle with power at the top and current and voltage at the bottom. Again, this
arrangement represents the actual position of each quantity in the Ohms law power
formulas.

The Power Triangle


One other point about Power, if the calculated power is positive in value for any formula
the component absorbs the power, but if the calculated power is negative in value the
component produces power, in other words it is a source of electrical energy. 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 Horsepower or hp. The relationship between horsepower
and watts is given as: 1hp = 746W.

Ohms Law Pie Chart


We can now take all the equations from above for finding Voltage, Current, Resistance
and Power and condense them into a simple Ohms Law pie chart for use in DC circuits
and calculations.

Ohms Law Pie Chart


Fundamentals: Electric Laws − Formulary −
Equations Formula wheel ▼ Important
formulas Electrical engineering laws Electronic

engineering laws
V comes from "voltage" and E from "electromotive
force". E means also energy, so V is chosen.
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 x 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 x I = 0 so again the power dissipated within the
circuit is 0.

As electrical power is the product of V x 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, power is dissipated in the form of Heat (heaters), Mechanical Work
such as motors, etc Energy in the form of radiated (Lamps) or as stored energy
(Batteries).

Energy in Electrical Circuits


Electrical Energy that is either absorbed or produced is the product of the electrical
power measured in Watts and the time in Seconds with the unit of energy given as Watt-
seconds 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, a single 100 W
light bulb connected for one hour will consume a total of 100 watts x 3600 sec = 360,000
Joules. So prefixes such as kilojoules (kJ = 103J) or megajoules (MJ = 106J) are used
instead. If the electrical power is measured in "kilowatts" and the time is given in hours
then the unit of energy is in kilowatt-hours or kWh which is commonly called a "Unit of
Electricity" and is what consumers purchase from their electricity suppliers.

Now that we know what is the relationship between voltage, current and resistance in a
circuit, in the next tutorial about DC Theory 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 unit.
Kirchoff's Circuit Laws

In 1845, a German physicist, Gustav Kirchoff developed a pair or set of rules or laws
which deal with the conservation of current and energy within electrical circuits, one of
these laws deals with current flow, Kirchoff's Current Law, (KCL) and the other one
which deals with voltage, Kirchoff's Voltage Law, (KVL).

Kirchoff's First Law - The Current Law, (KCL)


Kirchoff's 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 Kirchoff is known as the Conservation of Charge.

Kirchoff's Current Law

Here, the 3 currents entering the node, I1, I2, I3 are all positive in value and the 2 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
Kirchoff's current law when analysing parallel circuits.

Kirchoff's Second Law - The Voltage Law, (KVL)


Kirchoff's 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 Kirchoff is known as the Conservation of
Energy.

Kirchoff's 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 Kirchoff's voltage law when
analysing series circuits.

When analysing either DC circuits or AC circuits using Kirchoff's 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.

• Circuit - a circuit is a closed loop conducting path in which an electrical


current flows.
• Path - a line of connecting elements or sources with no elements or
sources included more than once.
• 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. No
components are inside a mesh.
• Components are connected in series if they carry the same current.
• Components are connected in parallel if the same voltage is across them.

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 Kirchoff's Current Law, KCL the equations are given as;

At node A : I1 + I2 = I3

At node B : I3 = I1 + I2
Using Kirchoff's Voltage Law, KVL the equations are given as;

Loop 1 is given as : 10 = R1 x I1 + R3 x I3 = 10I1 + 40I3

Loop 2 is given as : 20 = R2 x I2 + R3 x 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 = 20I1 + 40(I1 + I2) = 40I1 + 60I2

We now have two "Simultaneous Equations" that can be reduced to give us the value of
both 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 Kirchoff's 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 Kirchoff's Circuit Laws is as
follows:

• 1. Assume all voltage sources 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 Kirchoff's first law equations for each node.
• 4. Find Kirchoff'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 Kirchoff's Circuit Law to calculate the voltages and currents
circulating around a linear circuit, we can also use loop analysis to calculate the currents
in each independent loop helping to reduce the amount of mathematics required using
just Kirchoff's laws. In the next tutorial about DC Theory we will look at Mesh Current
Analysis to do just that.

Circuit Analysis
In the previous tutorial we saw that complex circuits such as bridge or T-networks can be
solved using Kirchoff's Circuit Laws. While Kirchoff´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 circuit from the previous section.

Mesh Analysis Circuit

One simple method of reducing the amount of math's involved is to analyse the circuit
using Kirchoff'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 Kirchoff'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


A more 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 Kirchoff´s method.

For example: : i1 = I1 , i2 = -I2 and I3 = I1 - I2

We now write Kirchoff'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. This
then gives us a matrix of:

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 called the resistance matrix.
and this gives I1 as -0.143 Amps and I2 as -0.429 Amps

As : I3 = I1 - I2

The current I3 is therefore given as : -0.143 - (-0.429) = 0.286 Amps

which is the same value of 0.286 amps, we found using Kirchoff´s circuit law in the
previous 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 about DC Theory we will look at Nodal Voltage
Analysis to do just that.

Nodal Voltage Analysis


As well as using Mesh Analysis to solve the currents flowing around complex circuits it is
also possible to use nodal analysis methods too. 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 Kirchoff's first law to find the voltage potentials around the circuit. 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 Kirchoff'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 Kirchoff'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.
Direct and Alternating Current

There are two different ways that electricity is produced, and they are used in most cases
for very different purposes. They can also be converted from one form to another, as
discussed in this section.

The first and simpler type of electricity is called direct current, abbreviated "DC". This is
the type of electricity that is produced by batteries, static, and lightning. A voltage is
created, and possibly stored, until a circuit is completed. When it is, the current flows
directly, in one direction. In the circuit, the current flows at a specific, constant voltage
(this is oversimplified somewhat but good enough for our needs.) When you use a
flashlight, pocket radio, portable CD player or virtually any other type of portable or
battery-powered device, you are using direct current. Most DC circuits are relatively low
in voltage; for example, your car's battery is approximately 12 V, and that's about as high
a DC voltage as most people ever use.

An idealized 12 V DC current. The voltage is considered positive because its potential is


measured relative to ground or the zero-potential default state of the earth.
(This diagram drawn to the same scale as the AC diagram below.)

The other type of electricity is called alternating current, or "AC". This is the electricity
that you get from your house's wall and that you use to power most of your electrical
appliances. Alternating current is harder to explain than direct current. The electricity is
not provided as a single, constant voltage, but rather as a sinusoidal (sine) wave that over
time starts at zero, increases to a maximum value, then decreases to a minimum value,
and repeats. A representation of an alternating current's voltage over time is shown in the
diagram below.

While simple direct current circuits are generally described only by their voltage,
alternating current

circuits require more detail. First of all, if the voltage goes from a positive value to a
negative value and back again, what do we say is the voltage? Is it zero, because it
averages out to zero? That would seem to imply that there is no energy there at all. But
imagine, if you will, a wave of water flowing across the surface of the sea. The peaks and
troughs of the wave seem to "cancel each other out", but the wave clearly exists and has
energy. The same is true of alternating current.

The way the science world measures the energy in an AC signal is to compute what is
called the root mean square (RMS) average of the voltage. In simple terms, the RMS
value of an electrical current is the number which represents the same energy that a DC
current at that voltage would produce; it is in essence an average of the alternating
current waveform. Whenever you see an AC voltage specification, they are giving you
the RMS number unless they say otherwise specifically. So for example, in North
America, most homes have 115 VAC electricity. This is AC electricity equivalent in
energy to a 115 V DC circuit. (This is an approximate number, and standard household
electricity in North America is also sometimes called 110VAC or 120VAC; it's the same
thing.) Other parts of the world use different voltages ranging from 100VAC to 240VAC,
and of course, heavy equipment anywhere can use much higher voltages.

The other key characteristic of AC is its frequency, measured in cycles per second (cps)
or, more commonly, Hertz (Hz). This number describes how many times in a second the
voltage alternates from positive to negative and back again, completing one cycle. In
North America, the standard is 60 Hz, meaning 60 cycles from positive to negative and
back again in one second. In other parts of the world the standard is 50 Hz.

Three cycles of an idealized North American 115 VAC, 60 Hz alternating current signal
(black curve).
Note that each cycle represents 16.67 milliseconds of time, because that is 1/60th of a
second. The
curve actually goes from -170 V to +170 V in order to provide the average (RMS) value
of 115 V.
The RMS equivalent is shown as a green horizontal line. To demonstrate what RMS
means, look at
the blue shaded area, which shows the total energy in the signal for one cycle. The green
shading is
the area between the RMS line and the zero line for one cycle, and represents the energy
in an equivalent
115 V DC signal. The definition of the RMS value is that which makes the green and blue
areas equal.
(This diagram drawn to the same scale as the DC diagram above.)

Why does standard electricity come only in the form of alternating current? There are a
number of reasons, but one of the most important is that a characteristic of AC is that it is
relatively easy to change voltages from one level to another using a transformer, while
transformers do not work for DC. This capability allows the companies that generate and
distribute electricity to do it in a more efficient manner, by transmitting it at high voltage
for long lengths, which reduces energy loss due to the resistance in the transmission
wires. Another reason is that it may be easier to mechanically generate alternating current
electricity than direct current.

Work, Power and Apparent Power

If we have electricity, what can we do with it? Well, we can do work, which in physics is
defined as transferring energy from one object or another through the application of
force. Essentially, any time you have a circuit with electricity flowing in it and you are
doing something with it, you are accomplishing work. The basic unit of work (or energy)
is a joule ("J").

Now, finally, we get to where we wanted to go in our discussion of electricity: the


definition of power. Power is simply the rate at which work is done. The more power you
have in a system, the more work you can get done in the same period of time. In terms of
electricity, increasing power means the ability to do more electrical work (for example,
running more appliances, or spinning a motor faster, or running a faster CPU, etc.) in the
same number of seconds. Power is measured in watts ("W"). Since power is the rate at
which work is done, one watt equals one joule of energy expended in one second:

Power (W) = Work (J) / Time (seconds)

Conversely, the amount of energy used by a device can be computed as the amount of
power it uses multiplied by the length of time over which that power is applied:

Work (J) = Power (W) * Time (seconds)


Computing electrical power can be very simple or very complicated, depending on the
type of electricity you are looking at. Let's start with direct current. Here, power (in
watts) is just the product of the voltage (in volts) and the current (in amps) of the circuit:

P (W) = V (V) * I (A)

Fairly simple stuff, and it makes sense: you do more work when you have electrons
pushing with more force (higher voltage) and also when you have more of them per
period of time (higher current). Since P = V*I, and I = V/R, another way to express
power is:

P = V² / R

For example, if you have a simple 5 V circuit running through a 20 ohm simple
resistance, you will have 250 mA of current, and the total power is 5*0.250=1.25 W.
Double the voltage to 10 V, and the power doesn't double; it increases by a factor of four,
because doubling the voltage while leaving the resistance the same will also double the
current. The new power is 5 W.

Unsurprisingly, with alternating current the answer is more complicated. To understand


it, it is necessary to introduce the concept of phase, which I will try to do without
overcomplicating things (not an easy task! :^) ). As illustrated on this page, alternating
current is a wave of voltage that swings between a large positive and negative value. The
current also makes this sinusoidal trip the same number of times per second. However,
sometimes the current and voltage don't peak at the exact same time. The timing
relationship between current and voltage of a flow is called its phase, and is expressed in
degrees. Why degrees? Well, a cycle of a sine wave is analogous to a circle. 360 degrees
is a full cycle, 180 degree half a cycle, and so on.

Now, what determines the phasing between the current and voltage? Primarily, it depends
on the kinds of loads being powered. Simple loads, such as light bulbs, heater elements
and the like, are said to be primarily resistive. These loads will cause the phase between
the current and voltage to be close to zero. When the phase angle is zero, the voltage and
current applied to the load is equal to the voltage and current used by the load.
115 VAC current and voltage driving a purely resistive load. The phase angle between
voltage and current
is about 0 degrees. Note that the voltage and current peak together.

Other loads, particularly items such as motors, are said to be reactive. Reactive loads are
caused by more complex opposition to the flow of alternating current such as that
produced by capacitors and inductors. They can cause the current and voltage to be out of
phase, in theory by as much as 90 degrees.
115 VAC current and voltage driving a (theoretical) purely reactive load. The current is
lagging behind
the voltage by about 90 degrees. (It is possible for the current to be leading by 90 degrees
also.)
Note that whenever one of voltage or current hits a peak, the other one is at zero!

If the phase angle between current and voltage is 90 degrees, then whenever voltage is at
its peak (either positive or negative), current is zero, and vice-versa. This is a "worst-
case" situation that doesn't normally arise in the real world because real loads aren't
purely reactive. A more typical situation is where the phase angle is about 45 degrees.
115 VAC current and voltage driving a partly resistive and partly reactive
load. The current is lagging behind the voltage by about 45 degrees, making
this an inductive load. If the load were capacitive, the current would be
leading the voltage. See here for more on inductors and capacitors.

"Alright, alright," you are saying. "Why do I care about all of this?" Well, here's one
important reason: PC power supplies are partly reactive loads, and often exhibit a phase
difference between voltage and current of about 45 degrees. This means that the voltage
and current applied to the load do not equal the voltage and current used by the load, and
you cannot compute the power used by the supply by simply multiplying the current and
voltage. OK, now here's where it gets interesting. :^) The voltage and current applied to
the load can be multiplied together to yield what is called apparent power, measured in
Volt-Amps (VA):

Apparent Power (VA) = V (V) * I (A)

Apparent power represents the voltage and current being sent to the device, and is used to
measure draw from the utility, for determining heat generation by equipment under use,
and for sizing wires and circuit breakers. The actual power used by the load is called
"true" power, or just power, and is measured in Watts. (Even though Watts = Volts *
Amps, apparent power is measured in VA to differentiate it from true power.) The
relationship between power and apparent power is expressed using this formula:

P (W) = cosine(phase) * Apparent Power (VA)


where "cosine" is the trigonometric function. "cosine(phase)" is also called the power
factor of the load. Let's try an example. Let's suppose we are trying to run a power supply
and the power supplied is 115V voltage and 2A of current. The apparent power is 115 * 2
= 230 VA. If the nature of the power supply is that its voltage and current are out of
phase by 50 degrees, then the power factor is cosine(50º) = 0.642 (sometimes expressed
as 64.2%) and the power used by the load is 148 W.

There's a particular place where all of this comes into play, and that is in the capacity and
sizing of uninterruptible power supplies. UPSes are normally specified in terms of
apparent power (VA), whereas PC power supplies are specified in terms of true power
(W). Many people use the numbers interchangeably, when they most definitely are not
the same! Now that you understand the difference between the two, and you know what a
power factor is, you are light years ahead of 95% of the population when it comes to
figuring out how to purchase a properly-sized UPS or similar device.
Combinations of Resistors

Resistors do not occur in isolation. They are almost always part of a


larger circuit, and frequently that larger circuit contains many resistors. It
is often the case that resistors occur in combinations that repeat.

Series Combinations of Resistors

Two elements are said to be in series whenever the same current


physically flows through both of the elements. The critical point is that the
same current flows through both resistors when two are in series. The
particular configuration does not matter. The only thing that matters is
that exactly the same current flows through both resistors. Current flows
into one element, through the element, out of the element into the other
element, through the second element and out of the second element. No
part of the current that flows through one resistor "escapes" and none is
added. This figure shows several different ways that two resistors in series
might appear as part of a larger circuit diagram.

Questions

Here is a circuit you may have seen before. Answer the questions
below for this circuit.
Q1. Are elements #3 and #4 in series?

Q2. Are elements #1 and #2 in series?

Q3. Is the battery in series with any element

You might wonder just how often you actually find resistors in series. The
answer is that you find resistors in series all the time.

An example of series resistors is in house wiring. The leads from the


service entrance enter a distribution box, and then wires are strung
throughout the house. The current flows out of the distribution box,
through one of the wires, then perhaps through a light bulb, back through
the other wire. We might model that situation with the circuit diagram
shown below.
In many electronic circuits series
resistors are used to get a different
voltage across one of the resistors. We'll
look at those circuits, called voltage dividers, in a short while. Here's the
circuit diagram for a voltage divider.

Besides resistors in series, we can also have other elements in series -


capacitors, inductors, diodes. These elements can be in series with other
elements. For example, the simplest form of filter, for filtering low
frequency noise out of a signal, can be built just by putting a resistor in
series with a capacitor, and taking the output as the capacitor voltage.

As we go along you'll have lots of opportunity to use and to expand what


you learn about series combinations as you study resistors in series.

Let's look at the model again. We see that the wires are actually small
resistors (small value of resistance, not necessarily physically small) in series
with the light bulb, which is also a resistor. We have three resistors in
series although two of the resistors are small. We know that the resistors
are in series because all of the current that flows out of the distribution
box through the first wire also flows through the light bulb and back
through the second wire, thus meeting our condition for a series connection.
Trace that out in the circuit diagram and the pictorial representation above.

Let us consider the simplest case of a series resistor connection, the


case of just two resistors in series. We can perform a thought experiment
on these two resistors. Here is the circuit diagram for the situation we're
interested in.
Imagine that they are embedded in an opaque piece of plastic, so that
we only have access to the two nodes at the ends of the series connection,
and the middle node is inaccessible. If we measured the resistance of the
combination, what would we find? To answer that question we need to define
voltage and current variables for the resistors. If we take advantage of the
fact that the current through them is the same (Apply KCL at the interior
node if you are unconvinced!) then we have the situation below.

Note that we have defined a voltage across each resistor (Va and Vb) and
current that flows through both resistors (Is) and a voltage variable, Vs, for
the voltage that appears across the series combination.

Let's list what we know.

• The current through the two resistors is the same.


• The voltage across the series combination is given by:
o Vs= Va + Vb
• The voltages across the two resistors are given by Ohm's Law:
o Va = Is Ra
o Vb = Is Rb

We can combine all of these relations, and when we do that we find the
following.

• Vs= Va + Vb
• Vs= Is Ra + Is Rb
• Vs= Is (Ra + Rb)
• Vs= Is Rseries

Here, we take Rseries to be the series equivalent of the two resistors in


series, and the expression for Rseries is:
Rseries = Ra + Rb
What do we mean by series equivalent? Here are some points to
observe.

• If current and voltage are proportional, then the device is a resistor.


• We have shown thatVs= Is Rseries, so that voltage is proportional to
current, and the constant of proportionality is a resistance.
• We will call that the equivalent series resistance.

There is also a mental picture to use when considering equivalent series


resistance. Imagine that you have two globs of black plastic. Each of the
globs of black plasic has two wires coming out. Inside these two black
plastic globs you have the following.

• In the first glob you have two resistors in series. Only the leads of
the series combination are available for measurement externally. You
have no way to penetrate the box and measure things at the interior
node.
• In the second box you have a single resistor that is equal to the series
equivalent. Only the leads of this resistor are available for
measurement externally.

Then, if you measured the resistance using the two available leads in the two
different cases you would not be able to tell which black plastic glob had the
single resistor and which one had the series combination.

Here are two resistors. At the top are two 2000W resistors. At the
bottom is single 4000W resistors. (Note, these are not exactly standard
sizes so it took a lot of hunting to find a supply store that sold them!). You
can click the green button to grow blobs around them.

Parallel Resistors

The other common connection is two elements in parallel. Two resistors


or any two devices are said to be in parallel when the same voltage physically
appears across the two resistors. Schematically, the situation is as shown
below.
Note that we have defined the voltage across both resistor (Vp) and the
current that flows through each resistor (Ia and Ib) and a voltage variable,
Vp, for the voltage that appears across the parallel combination.

Let's list what we know.

• The voltage across the two resistors is the same.


• The current through the parallel combination is given by:
o Ip= Ia + Ib
• The currents through the two resistors are given by Ohm's Law:
o Ia = Vp /Ra
o Ib = Vp /Rb

We can combine all of these relations, and when we do that we find the
following.

• Ip= Ia + Ib
• Ip= Vp /Ra + Vp /Rb
• Ip= Vp[ 1/Ra + 1/Rb]
• Ip= Vp/Rparallel

Here, we take Rparallel to be the parallel equivalent of the two resistors in


parallel, and the expression for Rparallel is:
1/Rparallel = 1/Ra + 1/Rb

There may be times when it is better to rearrange the expression for


Rparallel. The expression can be rearranged to get:

Rparallel = (Ra*Rb)/(Ra + Rb)

Either of these expressions could be used to compute a parallel


equivalent resistance. The first has a certain symmetry with the expression
for a series equivalent resistance.
Problems

P8. What is the equivalent resistance of this resistance combination?

Your grade is:

P9. What is the equivalent resistance of this resistance combination?

Your grade is:

P10. What is the equivalent resistance of this resistance combination?


Here all three resistors are 33 kΩ . Remember to input your answer in
ohms.

Your grade is:

P11. Here is an operational amplifier circut, a Wien-bridge oscillator. The


circuit is taken from Wojslaw and Moustakas' book Operational Amplifiers
(John Wiley & Sons, 1986, p100). Assuming that the amplifiers take no
current at the "+" and "-" terminals are resistors, R3 and R4 in series?

What If You Have A More Complex Circuit

Here's a circuit with resistors that has them connected in a different


way. For a short while we're going to work on the question of how to analyze
this circuit. For a start we're going to assume that this is a resistor. It has
two leads at the left (marked here with red dots) and we'll assume that we
want to find the equivalent resistance you would have at those leads.

We will use the following numerical values for the resistors in this
example, and we will work through using these values.

• Ra = 1500 Ω
• Rb = 3000 Ω
• Rc = 2000 Ω
• Rd = 1000 Ω
• Vs = 12 v
We need to figure out where we can start. We can start by trying to
find any of the combinations we've learned about. So let's think about
whether there are any series or parallel combinations and if there are let's
see if we can identify them. Then we can apply what we know about series
and parallel combinations. There's no guarantee that approach will work, but
it is worth a try. Let's look at two resistors at a time.

The first question is are there any series or parallel combinations?


Click the red button below to see two resistors in series.

Search

Introduction to Combination of Resistors - Series and


Parallel
A Resistor (R) is the most commonly used in all the electronic components. There are
different types of resistors that are available in the market to resist the flow of current in
an electrical circuit to act as the voltage dividers or voltage droppers.

The Resistors produce a voltage drop across themselves when the electrical current flows
through them to obey the Ohm's Law (V=IR) where V is the Voltage, I is the current and
R is the resistance, Different values of resistance produces different values of current or
voltage.

The Symbol of Resistor :

Equations of the Equivalent Resistance when the


Resistors are Combined in Series and Parallel
Connecting the Resistors in Series :

The Resistors can be connected together in a series connection or in a parallel connection


or in the combinations of both series and parallel to generate a complex networks whose
overall resistance will be the combination of the individual resistors. In any combination
all the resistors must obey the Ohm's Law and Kirchoff's Circuit Laws.
The equation of the equivalent resistance when all the resistors connected in series is

The Resistors are connected together in Parallel when both the terminals are respectively
connected to both the terminals of the other resistor or group of resistors. The voltage
drop across all the resistors will be same in the resistors that are connected in parallel
connection. All the Resistors in Parallel have the Common Voltage.

The equation of the equivalent resistance when all the resistors connected in parallel is

Example for the Combination of Resistors Combined in


Series and Parallel
1) Find the Equivalent resistance of the circuit

The values of the resistors given are:

In the above circuit, All the resistors are connected in series,The total current flows
through each resistor as all the resistors are connected in series.

The Equivalent resistance of the above circuit is

Therefore the Total resistance of the circuit is

2) Find the Equivalent resistance of the circuit


The values of the resistors given are:

In the above circuit, All the resistors are connected in parallel. So the current that is
supplied by the battery splits up and the amount of current going through each resistor
will depend on the resistance of the resistor.

The Equivalent resistance of the above circuit is

Therefore the Total resistance of the circuit is

Series circuits
Series circuits are sometimes called current-coupled or daisy chain-coupled. The current
in a series circuit is through every component in the circuit. Therefore, all of the
components in a series connection carry the same current.

[edit] Resistors
[edit] Inductors

Inductors follow the same law, in that the total inductance of non-coupled inductors in
series is equal to the sum of their individual inductances:

However, in some situations it is difficult to prevent adjacent inductors from influencing


each other, as the magnetic field of one device couples with the windings of its
neighbours. This influence is defined by the mutual inductance M. For example, if two
inductors are in series, there are two possible equivalent inductances depending on how
the magnetic fields of both inductors influence each other.

When there are more than two inductors, the mutual inductance between each of them
and the way the coils influence each other complicates the calculation. For a larger
number of coils the total combined inductance is given by the sum of all mutual
inductances between the various coils including the mutual inductance of each given coil
with itself, which we term self-inductance or simply inductance. For three coils, there are
six mutual inductances M12, M13, M23 and M21, M31 and M32. There are also the three self-
inductances of the three coils: M11, M22 and M33.

Therefore

Ltotal = (M11 + M22 + M33) + (M12 + M13 + M23) + (M21 + M31 + M32)

By reciprocity Mij = Mji so that the last two groups can be combined. The first three terms
represent the sum of the self-inductances of the various coils. The formula is easily
extended to any number of series coils with mutual coupling. The method can be used to
find the self-inductance of large coils of wire of any cross-sectional shape by computing
the sum of the mutual inductance of each turn of wire in the coil with every other turn
since in such a coil all turns are in series.
[edit] Capacitors

Capacitors follow the same law using the reciprocals. The total capacitance of capacitors
in series is equal to the reciprocal of the sum of the reciprocals of their individual
capacitances:

The working voltage of a series combination of identical capacitors is equal to the sum of
voltage ratings of individual capacitors. This simple relationship only applies if the
voltage ratings are equal as well as the capacitances. However, the division of DC voltage
between the capacitors is dominated by the leakage resistance of the capacitors, rather
than their capacitances, and this has considerable variation. To counter this equalising
resistors may be placed in parallel with each capacitor which effectively add to the
leakage current. The value of resistor chosen (perhaps a few megohms) is as large as
possible, but low enough to ensure that the capacitor leakage current is insignificant
compared to the current through the resistor. At DC, the circuit appears as a chain of
series identical resistors and equal voltage division between the capacitors is ensured. In
high-voltage circuits, the resistors serve an additional function as bleeder resistors.[4]

[edit] Switches

Two or more switches in series form a logical AND; the circuit only carries current if all
switches are 'on'. See AND gate.

[edit] Cells and batteries

A battery is a collection of electrochemical cells. If the cells are connected in series, the
voltage of the battery will be the sum of the cell voltages. For example, a 12 volt car
battery contains six 2-volt cells connected in series.

[edit] Parallel circuits


If two or more components are connected in parallel they have the same potential
difference (voltage) across their ends. The potential differences across the components
are the same in magnitude, and they also have identical polarities. The same voltage is
applicable to all circuit components connected in parallel. The total current I is the sum of
the currents through the individual components, in accordance with Kirchhoff’s current
law.

[edit] Resistors

The current in each individual resistor is found by Ohm's law. Factoring out the voltage
gives

To find the total resistance of all components, add the reciprocals of the resistances Ri of
each component and take the reciprocal of the sum. Total resistance will always be less
than the value of the smallest resistance:

For only two resistors, the unreciprocated expression is reasonably simple:

This sometimes goes by the mnemonic "product over sum".

For N equal resistors in parallel, the reciprocal sum expression simplifies to:

and therefore to:


.

To find the current in a component with resistance Ri, use Ohm's law again:

The components divide the current according to their reciprocal resistances, so, in the
case of two resistors,

An old term for devices connected in parallel is multiple, such as a multiple connection
for arc lamps.

[edit] Inductors

Inductors follow the same law, in that the total inductance of non-coupled inductors in
parallel is equal to the reciprocal of the sum of the reciprocals of their individual
inductances:

If the inductors are situated in each other's magnetic fields, this approach is invalid due to
mutual inductance. If the mutual inductance between two coils in parallel is M, the
equivalent inductor is:
If L1 = L2

The sign of M depends on how the magnetic fields influence each other. For two equal
tightly coupled coils the total inductance is close to that of each single coil. If the polarity
of one coil is reversed so that M is negative, then the parallel inductance is nearly zero or
the combination is almost non-inductive. It is assumed in the "tightly coupled" case M is
very nearly equal to L. However, if the inductances are not equal and the coils are tightly
coupled there can be near short circuit conditions and high circulating currents for both
positive and negative values of M, which can cause problems.

More than three inductors becomes more complex and the mutual inductance of each
inductor on each other inductor and their influence on each other must be considered. For
three coils, there are three mutual inductances M12, M13 and M23. This is best handled by
matrix methods and summing the terms of the inverse of the L matrix (3 by 3 in this
case).

The pertinent equations are of the form:

[edit] Capacitors

Capacitors follow the same law using the reciprocals. The total capacitance of capacitors
in parallel is equal to the sum of their individual capacitances:

The working voltage of a parallel combination of capacitors is always limited by the


smallest working voltage of an individual capacitor.

[edit] Switches
Two or more switches in parallel form a logical OR; the circuit carries
current if at least one switch is 'on'. See OR gate.

[edit] Cells and batteries

If the cells of a battery are connected in parallel, the battery voltage will be the same as
the cell voltage but the current supplied by each cell will be a fraction of the total current.
For example, if a battery contains four cells connected in parallel and delivers a current of
1 ampere, the current supplied by each cell will be 0.25 ampere. Parallel-connected
batteries were widely used to power the valve filaments in portable radios but they are
now rare.

Combination Circuits

Previously in Lesson 4, it was mentioned that there are two different ways to connect two
or more electrical devices together in a circuit. They can be connected by means of series
connections or by means of parallel connections. When all the devices in a circuit are
connected by series connections, then the circuit is referred to as a series circuit. When all
the devices in a circuit are connected by parallel connections, then the circuit is referred
to as a parallel circuit. A third type of circuit involves the dual use of series and parallel
connections in a circuit; such circuits are referred to as compound circuits or combination
circuits. The circuit depicted at the right is an example of the use of both series and
parallel connections within the same circuit. In this case, light bulbs A and B are
connected by parallel connections and light bulbs C and D are connected by series
connections. This is an example of a combination circuit.

When analyzing combination circuits, it is critically important to have a solid


understanding of the concepts that pertain to both series circuits and parallel circuits.
Since both types of connections are used in combination circuits, the concepts associated
with both types of circuits apply to the respective parts of the circuit. The main concepts
associated with series and parallel circuits are organized in the table below.

Series Circuits Parallel Circuits

• The current is the same in every • The voltage drop is the same across
resistor; this current is equal to that in each parallel branch.
the battery. • The sum of the current in each
• The sum of the voltage drops across individual branch is equal to the
the individual resistors is equal to the current outside the branches.
voltage rating of the battery. • The equivalent or overall resistance
• The overall resistance of the collection of the collection of resistors is
of resistors is equal to the sum of the given by the equation
individual resistance values,
1/Req = 1/R1 + 1/R2 + 1/R3 ...
Rtot = R1 + R2 + R3 + ...
Each of the above concepts has a mathematical expression. Combining the mathematical
expressions of the above concepts with the Ohm's law equation ( V = I • R) allows one to
conduct a complete analysis of a combination circuit.

Analysis of Combination Circuits

The basic strategy for the analysis of combination circuits involves using the meaning of
equivalent resistance for parallel branches to transform the combination circuit into a
series circuit. Once transformed into a series circuit, the analysis can be conducted in the
usual manner. Previously in Lesson 4, the method for determining the equivalent
resistance of parallel are equal, then the total or equivalent resistance of those branches is
equal to the resistance of one branch divided by the number of branches.

This method is consistent with the formula

1 / Req = 1 / R1 + 1 / R2 + 1 / R3 + ...

where R1, R2, and R3 are the resistance values of the individual resistors that are
connected in parallel. If the two or more resistors found in the parallel branches do not
have equal resistance, then the above formula must be used. An example of this method
was presented in a previous section of Lesson 4.

By applying one's understanding of the equivalent resistance of parallel branches to a


combination circuit, the combination circuit can be transformed into a series circuit. Then
an understanding of the equivalent resistance of a series circuit can be used to determine
the total resistance of the circuit. Consider the following diagrams below. Diagram A
represents a combination circuit with resistors R2 and R3 placed in parallel branches. Two
4- resistors in parallel is equivalent to a resistance of 2 . Thus, the two branches can be
replaced by a single resistor with a resistance of 2 . This is shown in Diagram B. Now
that all resistors are in series, the formula for the total resistance of series resistors can be
used to determine the total resistance of this circuit: The formula for series resistance is
Rtot = R1 + R2 + R3 + ...

So in Diagram B, the total resistance of the circuit is 10 .

Once the total resistance of the circuit is determined, the analysis continues using Ohm's
law and voltage and resistance values to determine current values at various locations.
The entire method is illustrated below with two examples.

Example 1:

The first example is the easiest case - the resistors placed in parallel have the same
resistance. The goal of the analysis is to determine the current in and the voltage drop
across each resistor.

As discussed above, the first step is to simplify the circuit by replacing the two parallel
resistors with a single resistor that has an equivalent resistance. Two 8 resistors in series
is equivalent to a single 4 resistor. Thus, the two branch resistors (R2 and R3) can be
replaced by a single resistor with a resistance of 4 . This 4 resistor is in series with R1
and R4. Thus, the total resistance is

Rtot = R1 + 4 + R4 = 5 + 4 + 6

Rtot = 15
Now the Ohm's law equation ( V = I • R) can be used to determine the total current in the
circuit. In doing so, the total resistance and the total voltage (or battery voltage) will have
to be used.

Itot = Vtot / Rtot = (60 V) / (15 )

Itot = 4 Amp

The 4 Amp current calculation represents the current at the battery location. Yet, resistors
R1 and R4 are in series and the current in series-connected resistors is everywhere the
same. Thus,

Itot = I1 = I4 = 4 Amp

For parallel branches, the sum of the current in each individual branch is equal to the
current outside the branches. Thus, I2 + I3 must equal 4 Amp. There are an infinite
number of possible values of I2 and I3 that satisfy this equation. Since the resistance
values are equal, the current values in these two resistors are also equal. Therefore, the
current in resistors 2 and 3 are both equal to 2 Amp.

I2 = I3 = 2 Amp

Now that the current at each individual resistor location is known, the Ohm's law
equation ( V = I • R) can be used to determine the voltage drop across each resistor.
These calculations are shown below.

V1 = I1 • R1 = (4 Amp) • (5 )
V1 = 20 V

V2 = I2 • R2 = (2 Amp) • (8 )

V2 = 16 V

V3 = I3 • R3 = (2 Amp) • (8 )

V3 = 16 V

V4 = I4 • R4 = (4 Amp) • (6 )

V4 = 24 V

The analysis is now complete and the results are summarized in the diagram below.
Example 2:

The second example is the more difficult case - the resistors placed in parallel have a
different resistance value. The goal of the analysis is the same - to determine the current
in and the voltage drop across each resistor.

As discussed above, the first step is to simplify the circuit by replacing the two parallel
resistors with a single resistor with an equivalent resistance. The equivalent resistance of
a 4- and 12- resistor placed in parallel can be determined using the usual formula for
equivalent resistance of parallel branches:

1 / Req = 1 / R1 + 1 / R2 + 1 / R3 ...

1 / Req = 1 / (4 ) + 1 / (12 )
-1
1 / Req = 0.333
-1
Req = 1 / (0.333 )

Req = 3.00

Based on this calculation, it can be said that the two branch resistors (R2 and R3) can be
replaced by a single resistor with a resistance of 3 . This 3 resistor is in series with R1
and R4. Thus, the total resistance is
Rtot = R1 + 3 + R4 = 5 + 3 + 8

Rtot = 16

Now the Ohm's law equation ( V = I • R) can be used to determine the total current in the
circuit. In doing so, the total resistance and the total voltage (or battery voltage) will have
to be used.

Itot = Vtot / Rtot = (24 V) / (16 )

Itot = 1.5 Amp

The 1.5 Amp current calculation represents the current at the battery location. Yet,
resistors R1 and R4 are in series and the current in series-connected resistors is
everywhere the same. Thus,

Itot = I1 = I4 = 1.5 Amp

For parallel branches, the sum of the current in each individual branch is equal to the
current outside the branches. Thus, I2 + I3 must equal 1.5 Amp. There are an infinite
possibilities of I2 and I3 values that satisfy this equation. In the previous example, the two
resistors in parallel had the identical resistance; thus the current was distributed equally
among the two branches. In this example, the unequal current in the two resistors
complicates the analysis. The branch with the least resistance will have the greatest
current. Determining the amount of current will demand that we use the Ohm's law
equation. But to use it, the voltage drop across the branches must first be known. So the
direction that the solution takes in this example will be slightly different than that of the
simpler case illustrated in the previous example.

To determine the voltage drop across the parallel branches, the voltage drop across the
two series-connected resistors (R1 and R4) must first be determined. The Ohm's law
equation ( V = I • R) can be used to determine the voltage drop across each resistor.
These calculations are shown below.

V1 = I1 • R1 = (1.5 Amp) • (5 )
V1 = 7.5 V

V4 = I4 • R4 = (1.5 Amp) • (8 )

V4 = 12 V

This circuit is powered by a 24-volt source. Thus, the cumulative voltage drop of a
charge traversing a loop about the circuit is 24 volts. There will be a 19.5 V drop (7.5 V +
12 V) resulting from passage through the two series-connected resistors (R1 and R4). The
voltage drop across the branches must be 4.5 volts to make up the difference between the
24 volt total and the 19.5-volt drop across R1 and R4. Thus,
V2 = V3 = 4.5 V

Knowing the voltage drop across the parallel-connected resistors (R1 and R4) allows one
to use the Ohm's law equation ( V = I • R) to determine the current in the two branches.

I2 = V2 / R2 = (4.5 V) / (4 )
I2 = 1.125 A

I3 = V3 / R3 = (4.5 V) / (12 )

I3 = 0.375 A

The analysis is now complete and the results are summarized in the diagram below.

Developing a Strategy

The two examples above illustrate an effective concept-centered strategy for analyzing
combination circuits. The approach demanded a firm grasp of the series and parallel
concepts discussed earlier. Such analyses are often conducted in order to solve a physics
problem for a specified unknown. In such situations, the unknown typically varies from
problem to problem. In one problem, the resistor values may be given and the current in
all the branches are the unknown. In another problem, the current in the battery and a few
resistor values may be stated and the unknown quantity becomes the resistance of one of
the resistors. Different problem situations will obviously require slight alterations in the
approaches. Nonetheless, every problem-solving approach will utilize the same principles
utilized in approaching the two example problems above.

The following suggestions for approaching combination circuit problems are offered to
the beginning student:

• If a schematic diagram is not provided, take the time to construct one. Use
schematic symbols such as those shown in the example above.
• When approaching a problem involving a combination
circuit, take the time to organize yourself, writing
down known values and equating them with a symbol
such as Itot, I1, R3, V2, etc. The organization scheme
used in the two examples above is an effective starting
point.
• Know and use the appropriate formulae for the
equivalent resistance of series-connected and parallel-connected resistors. Use of
the wrong formulae will guarantee failure.
• Transform a combination circuit into a strictly series circuit by replacing (in your
mind) the parallel section with a single resistor having a resistance value equal to
the equivalent resistance of the parallel section.
• Use the Ohm's law equation ( V = I • R) often and appropriately. Most answers
will be determined using this equation. When using it, it is important to substitute
the appropriate values into the equation. For instance, if calculating I2, it is
important to substitute the V2 and the R2 values into the equation.

For further practice analyzing combination circuits, consider analyzing the problems in
the Check Your Understanding section below.

Check Your Understanding


1. A combination circuit is shown in the diagram at the right. Use the diagram to answer
the following questions.

a. The current at location A is _____ (greater than, equal to, less than) the current at
location B.

b. The current at location B is _____ (greater than, equal to, less than) the current at
location E.

c. The current at location G is _____ (greater than, equal to, less than) the current at
location F.

d. The current at location E is _____ (greater than, equal to, less than) the current at
location G.

e. The current at location B is _____ (greater than, equal to, less than) the current at
location F.
f. The current at location A is _____ (greater than, equal to,
less than) the current at location L.

f. The current at location H is _____ (greater than, equal to,


less than) the current at location I.

2. Consider the combination circuit in the diagram at the right. Use the diagram to answer
the following questions. (Assume that the voltage drops in the wires themselves in
negligibly small.)

a. The electric potential difference (voltage drop) between points B and C is _____
(greater than, equal to, less than) the electric potential difference (voltage drop) between
points J and K.

b. The electric potential difference (voltage drop) between points B and K is _____
(greater than, equal to, less than) the electric potential difference (voltage drop) between
points D and I.

c. The electric potential difference (voltage drop) between points E and F is _____
(greater than, equal to, less than) the electric potential difference (voltage drop) between
points G and H.

d. The electric potential difference (voltage drop) between points E and F is _____
(greater than, equal to, less than) the electric potential difference (voltage drop) between
points D and I.

e. The electric potential difference (voltage drop) between points J and K is _____
(greater than, equal to, less than) the electric potential difference (voltage drop) between
points D and I.

f. The electric potential difference between points L and A is _____ (greater than, equal
to, less than) the electric potential difference (voltage drop) between points B and K.
3. Use the concept of equivalent resistance to determine the unknown resistance of the
identified resistor that would make the circuits equivalent.
4. Analyze the following circuit and determine the values of the total resistance, total
current, and the current at and voltage drops across each individual resistor.

5. Referring to the diagram in question #4, determine the ...

a. ... power rating of resistor 4.

b. ... rate at which energy is consumed by resistor 3.

Resistance of resistor combinations


Resistors can be placed in series or parallel. When placed in series the total resistance is
equal to the sum of the individual resistors:

Resistors in series

It is also worth noting that the same current flows through each resistor, but the voltage
across each resistor is proportional to the resistance of that particular resistor.
For resistors placed in parallel, the arithmetic is a little more complicated because the
reciprocal of the total resistance is equal to the sum of the reciprocals of the constituent
resistors:

Resistors in parallel

When there are only two resistors R1 and R2 in parallel this can be simplified to:

If both these resistors have the same value it can be seen that the overall value of the
resistance is half the value for the individual resistor.

When resistors are placed in parallel the voltage across the resistors is the same, but the
current through each one is inversely proportional to its resistance.

S-ar putea să vă placă și