You are on page 1of 47

CamScanner

CamScanner
CamScanner
CamScanner
CamScanner
CamScanner
CamScanner
CamScanner
CamScanner
CamScanner
III. Internet Related Articles
DIRECT CURRENT INSTRUMENTS AND MEASUREMENTS

What are DC Circuits?


Let us start by defining an electric circuit, also called an electric network. A simple circuit
can be described as a closed loop connected to a power source. The power source
provides a voltage (electric potential) that pushes electrons around the loop.
The electrons flow through the whole circuit, which is usually made of conductors and
devices like resistors, capacitors, or inductors.
Electron flow is called electric current. Electrons flow in the opposite direction of
current, and we usually think about current flow, not electron flow.

This is where the name Direct Current (DC) comes from. A direct current circuit is one in
which the current only flows in one direction. In other words, in a DC circuit, the flow of
electricity goes in one direction; from the positive terminal to the negative terminal. In a
DC circuit, the value of the voltage might change. For instance, the voltage across the
circuit might alternate from .5 to 1.5 volts. But the polarity of voltage, or direction of
current flow never changes; it always goes in the same direction.

To put this into perspective, AC circuits feature an Alternating Current, in which the
direction of current switches from one side to the other. Direct Current (DC) always
pushes the electrons along in the same direction, but alternating current (AC) pushes
electrons one way and then pulls them back in the other direction. We will explore more
about AC circuits in the next module. For now, just keep in mind that in a DC circuit, the
current always flows in one direction; in an AC circuit the current alternates and flows
back and forth within a circuit.

Direct current (DC) is a fundamental form of electrical power characterized by a


continuous flow of electric charge in a single direction. Unlike alternating current (AC),
which periodically changes direction, DC maintains a constant polarity, making it a
crucial component in various applications across industries. The concept of direct current
traces back to the pioneering work of scientists like Thomas Edison in the late 19 th
century, who advocated for the use of DC in early electric power systems. While AC has
become predominant in power distribution due to its efficiency in long-distance
transmission, DC remains indispensable in specific technologies and niche applications.

One of the primary advantages of DC is its simplicity, as it involves a straightforward


flow of electrons from a positive terminal to a negative terminal. This simplicity makes
DC suitable for certain electronic devices, such as batteries and electronic components,
where a stable and unidirectional flow of electricity is essential. Batteries, for instance,
generate DC power through chemical reactions, providing a reliable source of energy for
portable devices and vehicles.

Moreover, direct current plays a pivotal role in modern power systems through
technologies like rectifiers and inverters. Rectifiers convert AC to DC, enabling the
utilization of DC power in applications where it is more practical or efficient. Inverters,
on the other hand, perform the opposite function, converting DC to AC, allowing
integration with the broader electrical grid and facilitating the use of renewable energy
sources, such as solar panels and wind turbines.

DC power is also prevalent in electronic circuits and semiconductor devices, where a


consistent flow of current is crucial for their proper operation. Computers,
telecommunication equipment, and many electronic devices rely on DC power for their
internal components, ensuring stable and reliable operation. Additionally, DC motors find
widespread use in various industries due to their efficiency, reliability, and precise control
capabilities, making them suitable for applications ranging from industrial machinery to
transportation.

In recent years, advancements in power electronics and high-voltage DC transmission


have renewed interest in DC as a viable option for long-distance power transmission.
High-voltage direct current (HVDC) systems offer reduced transmission losses and
improved efficiency over long distances compared to traditional AC transmission. This
resurgence underscores the continued importance and versatility of direct current in
shaping the future of electrical power systems.

For electricity to flow through a circuit, it must have somewhere to come from and
somewhere to go. If a circuit is connected to a power source but does not give the
electrons somewhere to go, electricity cannot flow. For instance, a wire in the circuit is no
longer making contact on one side. This is called an open circuit. Open circuits are
otherwise complete circuits that have a break somewhere, preventing current from
flowing. Open circuits are not really circuits at all because current cannot flow within
them. However, the term ‘open circuit’ describes a condition in which current would flow
through the circuit.

Direct Current (DC) is a one-directional flow of electric current. Direct current circuits
are a type of electric circuit in which current flows in only one direction. DC circuits
are incredibly important to our lives. Most electronics and logic-based systems (i.e.,
computers) use DC. Everything that has a battery uses DC; laptops, cell phones,
flashlights all use DC. Even cars use DC (supplied by a car battery). Desktop computers
have power supplies to convert alternating current (AC) from a wall outlet into DC.

Learning about DC circuits is a great way to begin to understand electric circuits and
electronics. That is because DC is the system used in most electronics. Direct current
systems are also easier to understand than systems using alternating current (AC) but AC
systems rely on many of the same principles and circuit components as DC systems. We
can learn about important principles with DC circuits and then use what we have learned
to analyze circuits with other types of current.

That is why DC circuits are the best way to learn about the three components that are
considered the building blocks of classical electrical engineering: resistors, capacitors,
and inductors.

DC Circuits are the basic building block of many electronics, and are the best way to
begin to understand electrical circuits. Most electronics are DC and others use a DC
power supply, so DC circuits is a topic that is of great importance.
There are two ‘tracks’ of knowledge when learning about any circuits; theoretical and
practical. For electronics, there are fundamental (theoretical) principles, as well as
components that function by making practical use of the theory.
For example, resistors are circuit components that rely on the principle of electrical
resistance in order to function. It is important to understand both in order to have a well-
rounded approach to circuits utilizing resistors.

How do DC Circuits Work?

DC Circuits are made up of two main things: components, and conductors.

1. Components are the ‘blocks’ that make up the circuit. Power sources,
resistors, capacitors, LEDs, and transistors are examples of common
components.
2. Conductors provide paths for current between all the components of the
circuit.
Components

Perhaps the most important component is the power source. Without a power source,
there would be nothing to drive current in the circuit at all.

DC circuits are powered by direct current sources, which are power sources that generate
a potential difference (voltage) between two points.
If we were to separate some electrons from their atoms and then connect the electrons,
through conductors, back to their original atoms, we would have a DC circuit. The
separation between the negatively charged electrons and positively charged atoms (ions)
results in a difference of electric potential (voltage). Which is almost exactly what
batteries do.

Batteries are common DC sources, generating a voltage that is close to their stated
voltage, at least until they begin to be depleted. For example, a common battery in the US
is a 1.5V ‘AA’ battery, which can be used to power a DC circuit. When you put batteries
back-to-back (as in many devices like remote controls), you add the voltage of each one
to the total. The total voltage represents the ‘push’ against the electrons in the circuit, like
pressure in a pipe.

Another common component is a resistor, which is a device that restricts the flow of
current. Why would we ever want to do that?

Let us say we want to use a battery to power a DC circuit. We will want to include at
least one resistor so the battery does not short circuit! We connect the positive and
negative terminals of the battery to each end of the resistor.

It also helps to have a visual confirmation that current is flowing. We can connect an
LED in to the circuit, which will light up when current is flowing. Make sure that the
negative terminal of the LED is connected to the negative terminal of the battery.

The battery’s voltage will push electrons out of the negative terminal, through the
resistor, and into the positive terminal of the battery. As the electrons flow, they
generate current in the opposite direction.

The battery has an excess negative charge available inside its negative terminal, and
positive ions that are trying to attract electrons in its positive terminal. Between the
negative and positive terminals is a force that pushes electrons from the negative to the
positive terminals if we connect them to each other. This force is what drives a DC
circuit.

Conductors

Circuits need something to connect the different components together. This is done
with conductors, which are often copper or aluminum wires. The most important thing
about conductors is that they allow current to flow with minimal resistance. Resistance
creates a power loss so it is always best to minimize it whenever it is not wanted.

Over short distances, aluminum or copper provide a low enough resistance for circuits to
function well.
Combined with the circuit components, conductors provide a complete path for electric
current to follow through the circuit.

Direct current does not change with time. It is a flat line that does not go up or down.
This means that the movement of electrons does not speed up or slow down. The current
does not change direction, either. The electrons leave the power source, go through the
circuit, and go back into the power source on the other terminal. Direct current means
that the current in real life, this is not possible; there is always some variation (however
small) in the current supplied by a man-made machine. The definition of a DC circuit is
one in which the current changes direction. It may shift values, but it is always positive (a
negative value indicates a current moving in the opposite direction). does not change.

How are DC Sources Powered?

In general, we will use a battery as our source of electrical energy for DC circuits, and we
will cover batteries in the next lesson. A battery can be thought of as a pump for electric
charge, as it pushes electrons through a circuit via its’ negative terminal, and pulls
electrons into its’ positive terminal. In doing so, it powers the circuit to function; without
a battery or other power source, an electric circuit would just be a bunch of connected
devices that would like a pipe containing water but without a pump to circulate or push
the water.

However, the type of source does not actually matter. It could be a battery, battery array,
power supply, or even something called a function generator. Function generators are
machines that can produce a wide range of voltage sources to help us build and test
circuits. A DC circuit generally contains at least one power source and at least one
resistor, and this is the simple system that we will use to begin our exploration of
electrical circuits. If you can understand such a system, you will have an intuitive
knowledge of what a circuit is an important concept like current, voltage, and resistance,
and Ohm’s Law. Along the way we will learn the basics of electrical components like
batteries and resistors. We will discuss resistors in series and parallel, voltage dividers,
Kirchoff’s Current and Voltage Laws, combinations of resistors and capacitors, and
finish this module by learning about a cool circuit, the RC circuit. If you can understand
the RC circuit, you will be well on your way to a mastery of basic electrical theory!

The direct current meaning is when the electrical current flows consistently in a single
direction. Direct current is produced by rectifiers, batteries, generators with commutators,
and fuel cells. For example, the current flowing in appliances running on batteries or in a
flashlight is a direct current. The most common use and essentiality of direct current is
electroplating. Direct current was supplanted by AC currents (alternating current) for
typical commercial power.
AC DC current is one of the most important topics covered in Physics. Alternating
Current or AC currents is defined as a type of electrical current in which the electron flow
direction alternates back and forth at regular cycles or time intervals. The typical example
of current AC is the current flowing household electricity and power lines from a wall
outlet. You can use any electrical appliance in India if it meets the standard voltage of
230 Volts and the frequency of 50 Hertz.

Difference Between Direct Current and Alternating Current

The primary difference between AC and DC currents are-


The current which changes its direction at regular cycle or time intervals are defined as
current AC or Alternating current. Direct current meaning states unidirectional or that
current DC flows in only one direction.
A significant difference between AC current and DC current is that the alternating current
frequency is between 50 to 60 Hertz. In contrast, Direct current's frequency remains zero
as per the country's standard guidelines.
An alternator generates the Alternating current while the Direct current is generated by
cells, generator, and battery.
A few substations require AC for the generation and transmission of electricity, while
extra substations require electricity transmission through Direct currents.
The DC and AC current is often inter-converted based on the need. The alternating
current is converted into a Direct current using a rectifier, whereas the Direct current is
transformed into an Alternating current using an inverter.

AC-DC current holds large applications. AC currents are used in factories, household
purposes, and industries, while DC mainly uses flash lighting, electrolysis, electronic
equipment, hybrid vehicles, etc.

Similarities Between DC And AC Current

Both AC current DC current results from the induced charge that moves through wires to
transmit electrical energy and then use it to power various other devices.

In AC-DC current, negative electron flow produces the electrical current, and both the
currents are harnessed energy that can be tapped. The current flow through a conductor,
such as power lines between the device and the power source, uses the power. The
negative electrons flow through the line as they are attracted towards the positive charge.
In both AC current and DC current, a voltage source initiates the current flow in the
circuits, converting from one type of current to the other and is considered more
accessible.
Application of Direct Current and Alternating Current

The Application of DC is in-.


Batteries: Both non-rechargeable and rechargeable batteries can only supply current DC.
However, the rechargeable batteries need regular recharging when using the DC.

Electronic Equipment: All major equipment like cellphones, computers, radios, and all
electronic equipment use DC to power the electronic circuits.
Specific Electrical Equipment: Although most electrical equipment uses AC, a few of
them use direct current.
Solar Panels: Solar panels mainly produce direct current. However, when used with an
AC main to supply local AC power, an inverter is required to enable the direct current,
DC from the solar panels and convert it into AC.

The practical applications of electricity began with DC. The basic discoveries of Galvani,
Volta, Ampere or Ohm was on DC. The first widespread practical application was DC
telegraphy, but electric lighting also began with DC powered by dynamos. First came
carbon arc lamps operated in series at constant current and fed from series-wound
generators. Later came carbon-filament incandescent lamps operated in parallel at
constant voltage and supplied from shunt-wound generators.

The first electric central station in the world was built on Pearl Street in New York by
Thomas A. Edison in 1882. It supplied DC at 110 V to an area roughly of 1.6 km in
radius, and it had DC generators driven by steam engines. Within a few years similar
stations were in operation in the central districts of most large cities throughout the
world. By the mid-1880s alternating current (AC) systems began to compete with DC
systems. Inventors such as Nicola Tesla, William Stanley, Michael von Dolivo-
Dobrowolsky, Elihu Thomson, Lucien Gaillard, John Gibbs, and others working in
Europe and North America all contributed to AC technology.

In 1889, René Thury developed the first commercial system for high-voltage DC
transmission in Europe, supplying Genoa, Italy from Guarente River hydro turbines. The
system was composed by generators in series to attain high transmission voltages. When
loads were added to the system, other generators were added to maintain the voltage in
the load.

In 1889, the first long distance transmission of DC electricity in the United States was
switched on at Willamette Falls Station, in Oregon. In 1890 a flood destroyed the
Willamette Falls DC power station. This unfortunate event paved the way for the first
long distance transmission of AC electricity in the world when Willamette Falls Electric
Company installed experimental AC generators from Westinghouse in 1890. In 1896, the
first AC generation and transmission system was finished in the Niagara Falls using
Westinghouse equipment.
This race between AC and DC systems was faced by great personalities of the time, in
what would become the first standards war. Much has been written about the so called
“war of currents,” but this conflict was more a media fight than a conflict that had real
importance in the selection of the winning system. Final decisions on the type of system
to be applied always ended up being based on technical reasons, and AC systems of the
time offered greater advantages than DC systems given the needs and available
technology of the time. N. Tesla's work contributed greatly to demonstrate the benefits of
the use of AC systems, like the invention of the induction motor in 1888. Increasingly,
promoted by industrialists such as George Westinghouse, the advantages of AC electric
utility service became obvious, and by the end of the 19th century DC systems began an
inevitable decline.

The advent of the transformer, tree-phase circuits, and the induction motor promoted the
strengthening of AC electric systems as the world standard.

The transformer made possible the use of different voltage levels for generation,
transmission, distribution, and use; particularly important for the high-voltage power
transmission over long distances.

With the use of three-phase networks it was possible to ensure a smooth, non-pulsating
flow of power and bring an easy way to interrupt current on high-voltage equipment. The
induction motor is rugged, cheap, and serves most industrial and residential purposes.
Also, the advent of steam turbines, which are best at high speeds, gave a great advantage
to AC generators since the commutators of DC motors and generators impose limitations
on the voltage, size, and especially in speed of these machines.

The victory of AC over DC was almost complete. But some vestiges of DC distribution
can be found in the electric traction system (trolley bus, railways, or subway). Also, some
cities continued to use DC well into the 20th century. For example, in Europe, Helsinki
had a DC network until the late 1940s, Stockholm lost its dwindling DC network as late
as the 1970s, and London had some loads on DC as late as 1981. In USA, certain
locations in Boston still used 110 volts DC in the 1960s. In 2007, the last DC circuit, a
vestige of 19th century DC system of New York City was shut down. In the early 20th
century, AC systems were imposed over DC systems due to the benefits they offered and
because these benefits could not be reproduced by DC systems with the available
technology of the time. Some of these benefits have to do with the ease of changing
voltages and current interruption.

These benefits allowed AC systems the transmission of energy over long distances, but
with the growth of electrical networks, these AC systems are operated to their limit and
DC systems have proven to be of great help to solve the problems that large electrical
systems have, including distance limitations.

The natural descendants of DC systems of the early 20th century are the current HVDC
systems. The HVDC systems have evolved considerably since their inception in the mid-
20th century based on mercury-arc technology to become an ideal solution for large
block transmission over long distances. This ensures a great future for the HVDC systems
and a great development potential for large electrical systems.

The development of power electronics has enabled the growth that HVDC systems have
had in recent years, and have helped to demonstrate the great value that DC systems have
for modern networks.

As more than one hundred years ago, today's DC technology remains a fundamental part
of modern electrical systems.

There are, however, two circuit analysis rules that can be used to analyze any circuit,
simple or complex. These rules are special cases of the laws of conservation of charge
and conservation of energy. The rules are known as Kirchhoff’s rules, after their
inventor Gustav Kirchhoff (1824–1887).

KIRCHHOFF’S RULES

 Kirchhoff’s first rule—the junction rule. The sum of all currents entering a
junction must equal the sum of all currents leaving the junction.
 Kirchhoff’s second rule—the loop rule. The algebraic sum of changes in potential
around any closed-circuit path (loop) must be zero.
 Explanations of the two rules will now be given, followed by problem-solving
hints for applying Kirchhoff’s rules, and a worked example that uses them.
 Kirchhoff’s First Rule
 Kirchhoff’s first rule (the junction rule) is an application of the conservation of
charge to a junction. Current is the flow of charge, and charge is conserved; thus,
whatever charge flows into the junction must flow out. Kirchhoff’s first rule
requires that I1=I2+I3. Equations like this can and will be used to analyze circuits
and to solve circuit problems.

MAKING CONNECTIONS: CONSERVATION LAWS

Kirchhoff’s rules for circuit analysis are applications of conservation laws to circuits.
The first rule is the application of conservation of charge, while the second rule is the
application of conservation of energy. Conservation laws, even used in a specific
application, such as circuit analysis, are so basic as to form the foundation of that
application.

Kirchhoff’s Second Rule

Kirchhoff’s second rule (the loop rule) is an application of conservation of energy. The
loop rule is stated in terms of potential, V, rather than potential energy, but the two are
related since Peele=peeled=Qu. Recall that emf is the potential difference of a source
when no current is flowing. In a closed loop, whatever energy is supplied by emf must be
transferred into other forms by devices in the loop, since there are no other ways in which
energy can be transferred into or out of the circuit.

Kirchhoff’s second rule requires emf−Ir−IR1−IR2=0emf−Ir−IR1−IR2=0. Rearranged,


this is emf=Ir+IR1+IR2emf=Ir+IR1+IR2, which means the emf equals the sum of
the IRIR (voltage) drops in the loop.

Applying Kirchhoff’s Rules

By applying Kirchhoff’s rules, we generate equations that allow us to find the unknowns
in circuits. The unknowns may be currents, emfs, or resistances. Each time a rule is
applied, an equation is produced. If there are as many independent equations as
unknowns, then the problem can be solved. There are two decisions you must make when
applying Kirchhoff’s rules. These decisions determine the signs of various quantities in
the equations you obtain from applying the rules.

1. When applying Kirchhoff’s first rule, the junction rule, you must label the current
in each branch and decide in what direction it is going
2. When applying Kirchhoff’s second rule, the loop rule, you must identify a closed
loop and decide in which direction to go around it, clockwise or
counterclockwise.

Instrumentation is a field of study and work centering on measurement and control of


physical processes. These physical processes include pressure, temperature, flow rate,
and chemical consistency. An instrument is a device that measures and/or acts to control
any kind of physical process. Since electrical quantities of voltage and current are easy to
measure, manipulate, and transmit over long distances, they are widely used to represent
such physical variables and transmit the information to remote locations.
A signal is any kind of physical quantity that conveys information. Audible speech is
certainly a kind of signal, as it conveys the thoughts (information) of one person to
another through the physical medium of sound. Hand gestures are signals, too, conveying
information by means of light.
This text is another kind of signal, interpreted by your English-trained mind as
information about electric circuits. In this chapter, the word signal will be used primarily
in reference to an electrical quantity of voltage or current that is used
to represent or signify some other physical quantity.
Analog vs. Digital

An analog signal is a kind of signal that is continuously variable, as opposed to having a


limited number of steps along its range (called digital). A well-known example of analog
vs. digital is that of clocks: analog being the type with pointers that slowly rotate around
a circular scale, and digital being the type with decimal number displays or a “second-
hand” that jerks rather than smoothly rotates.
The analog clock has no physical limit to how finely it can display the time, as its
“hands” move in a smooth, pause less fashion. The digital clock, on the other hand,
cannot convey any unit of time smaller than what its display will allow for. The type of
clock with a “second-hand” that jerks in 1-second intervals are a digital device with a
minimum resolution of one second.
Both analog and digital signals find application in modern electronics, and the
distinctions between these two basic forms of information are something to be covered in
much greater detail later in this book. For now, I will limit the scope of this discussion to
analog signals, since the systems using them tend to be of simpler design.
For the most basic overview of this subject, see this video tutorial on analog and digital
electronics.
With many physical quantities, especially electrical, analog variability is easy to come
by. If such a physical quantity is used as a signal medium, it will be able to represent
variations of information with almost unlimited resolution.
Industrial Instrumentation System

In the early days of industrial instrumentation, compressed air was used as a signaling
medium to convey information from measuring instruments to indicating and controlling
devices located remotely. The amount of air pressure corresponded to the magnitude of
whatever variable was being measured. Clean, dry air at approximately 20 pounds per
square inch (PSI) was supplied from an air compressor through tubing to the measuring
instrument and was then regulated by that instrument according to the quantity being
measured to produce a corresponding output signal.
Aside from certain practical limits imposed by the mechanics of air pressure devices, this
pneumatic signal is infinitely variable, able to represent any degree of change in the
water’s level and is, therefore, analog in the truest sense of the word.
Crude as it may appear, this kind of pneumatic signaling system formed the backbone of
many industrial measurement and control systems around the world, and still sees use
today due to its simplicity, safety, and reliability. Air pressure signals are easily
transmitted through inexpensive tubes, easily measured (with mechanical pressure
gauges), and are easily manipulated by mechanical devices using bellows, diaphragms,
valves, and other pneumatic devices. Air pressure signals are not only useful
for measuring physical processes, but for controlling them as well.
With a large enough piston or diaphragm, a small air pressure signal can be used to
generate a large mechanical force, which can be used to move a valve or other controlling
device. Complete automatic control systems have been made using air pressure as the
signal medium. They are simple, reliable, and relatively easy to understand. However, the
practical limits for air pressure signal accuracy can be too limiting in some cases,
especially when the compressed air is not clean and dry, and when the possibility for
tubing leaks exist.
With the advent of solid-state electronic amplifiers and other technological advances,
electrical quantities of voltage and current became practical for use as analog instrument
signaling media. Instead of using pneumatic pressure signals to relay information about
the fullness of a water storage tank, electrical signals could relay that same information
over thin wires (instead of tubing) and not require the support of such expensive
equipment like air compressors to operate. Analog electronic signals are still the primary
kinds of signals used in the instrumentation world today (January of 2001), but it is
giving way to digital modes of communication in many applications (more on that
subject later). Despite changes in technology, it is always good to have a thorough
understanding of fundamental principles, so the following information will never really
become obsolete.

Live Zero

One important concept applied in many analog instrumentation signal systems is that of
“live zero,” a standard way of scaling a signal so that an indication of 0 percent can be
discriminated from the status of a “dead” system. Take the pneumatic signal system as an
example: if the signal pressure range for transmitter and indicator was designed to be 0 to
12 PSI, with 0 PSI representing 0 percent of process measurement and 12 PSI
representing 100 percent, a received signal of 0 percent could be a legitimate reading of 0
percent measurement or it could mean that the system was malfunctioning (air
compressor stopped, tubing broken, transmitter malfunctioning, etc.). With the 0 percent
point represented by 0 PSI, there would be no easy way to distinguish one from the other.
If, however, we were to scale the instruments (transmitter and indicator) to use a scale of
3 to 15 PSI, with 3 PSI representing 0 percent and 15 PSI representing 100 percent, any
kind of a malfunction resulting in zero air pressure at the indicator would generate a
reading of -25 percent (0 PSI), which is clearly a faulty value. The person looking at the
indicator would then be able to immediately tell that something was wrong.
Not all signal standards have been set up with live zero baselines, but the more robust
signals standards (3-15 PSI, 4-20 mA) have, and for good reason.
REVIEW:

 A signal is any kind of detectable quantity used to communicate information.


 An analog signal is a signal that can be continuously, or infinitely, varied to
represent any small amount of change.
 Pneumatic, or air pressure, signals were once common in industrial
instrumentation signal systems. These have been largely superseded by
analog electrical signals such as voltage and current.
 A live zero refers to an analog signal scale using a non-zero quantity to
represent 0 percent of real-world measurement so that any system
malfunction resulting in a natural “rest” state of zero signal pressure, voltage,
or current can be immediately recognized.

Standard measurements of voltage and current alter the circuit being measured,
introducing uncertainties in the measurements. Voltmeters draw some extra current,
whereas ammeters reduce current flow. Null measurements balance voltages so that
there is no current flowing through the measuring device and, therefore, no alteration of
the circuit being measured.

Null measurements are generally more accurate but are also more complex than the use
of standard voltmeters and ammeters, and they still have limits to their precision. In this
module, we shall consider a few specific types of null measurements, because they are
common and interesting, and they further illuminate principles of electric circuits.

The Potentiometer

Suppose you wish to measure the emf of a battery. Consider what happens if you connect
the battery directly to a standard voltmeter. (Once we note the problems with this
measurement, we will examine a null measurement that improves accuracy.) As
discussed before, the actual quantity measured is the terminal voltage V, which is related
to the emf of the battery by V=emf−Ire=emf−Ire, where I is the current that flows and r is
the internal resistance of the battery.

The emf could be accurately calculated if r were very accurately known, but it is usually
not. If the current I could be made zero, then V=emf=emf, and so emf could be directly
measured. However, standard voltmeters need a current to operate; thus, another
technique is needed.

A potentiometer is a null measurement device for measuring potentials (voltages). A


voltage source is connected to a resistor R, R, say, a long wire, and passes a constant
current through it. There is a steady drop in potential (an IRIR drop) along the wire, so
that a variable potential can be obtained by making contact at varying locations along the
wire.

Note that enfeoff opposes the other voltage source. The location of the contact point (see
the arrow on the drawing) is adjusted until the galvanometer reads zero. When the
galvanometer reads zero, elf=IRxemfx=IRx, where Rxx is the resistance of the section of
wire up to the contact point. Since no current flows through the galvanometer, none flows
through the unknown emf, and so emfxemfx is directly sensed. Because a long uniform
wire is used for R, the ratio of resistances Rx/Rsx/s is the same as the ratio of the lengths
of wire that zero the galvanometer for each emf. The three quantities on the right-hand
side of the equation are now known or measured, and emfxemfx can be calculated. The
uncertainty in this calculation can be considerably smaller than when using a voltmeter
directly, but it is not zero. There is always some uncertainty in the ratio of
resistances Rx/Rsx/s and in the standard emfsemfs. Furthermore, it is not possible to tell
when the galvanometer reads exactly zero, which introduces error into both Rxx and Rss,
and may also affect the current I.

Resistance Measurements and the Wheatstone Bridge

There is a variety of so-called ohmmeters that purport to measure resistance. What the
most common ohmmeters do is to apply a voltage to a resistance, measure the current,
and calculate the resistance using Ohm’s law. Their readout is this calculated resistance.
Such configurations are limited in accuracy, because the meters alter both the voltage
applied to the resistor and the current that flows through it.

The Wheatstone bridge is a null measurement device for calculating resistance by


balancing potential drops in a circuit. The device is called a bridge because the
galvanometer forms a bridge between two branches. A variety of bridge devices are used
to make null measurements in circuits.

Ohm’s law and Kirchhoff’s method are useful to analyze and design electrical circuits,
providing you with the voltages across, the current through, and the resistance of the
components that compose the circuit. To measure these parameters, require instruments,
and these instruments are described in this section.

DC Voltmeters and Ammeters

Whereas voltmeters measure voltage, ammeters measure current. Some of the meters in
automobile dashboards, digital cameras, cell phones, and tuner-amplifiers are voltmeters
or ammeters. The internal construction of the simplest of these meters and how they are
connected to the system they monitor give further insight into applications of series and
parallel connections.

Measuring Current with an Ammeter

To measure the current through a device or component, the ammeter is placed in series
with the device or component. A series connection is used because objects in series have
the same current passing through them.

Ammeters need to have a very low resistance, a fraction of a milliohm. If the resistance is
not negligible, placing the ammeter in the circuit would change the equivalent resistance
of the circuit and modify the current that is being measured. Since the current in the
circuit travels through the meter, ammeters normally contain a fuse to protect the meter
from damage from currents which are too high.
Measuring Voltage with a Voltmeter

A voltmeter is connected in parallel with whatever device it is measuring. A parallel


connection is used because objects in parallel experience the same potential difference.

Since voltmeters are connected in parallel, the voltmeter must have a very large
resistance. Digital voltmeters convert the analog voltage into a digital value to display on
a digital readout.

Analog and Digital Meters

You may encounter two types of meters in the physics lab: analog and digital. The term
‘analog’ refers to signals or information represented by a continuously variable physical
quantity, such as voltage or current. An analog meter uses a galvanometer, which is
essentially a coil of wire with a small resistance, in a magnetic field, with a pointer
attached those points to a scale. Current flows through the coil, causing the coil to rotate.
To use the galvanometer as an ammeter, a small resistance is placed in parallel with the
coil. For a voltmeter, a large resistance is placed in series with the coil. A digital
meter uses a component called an analog-to-digital (A to D) converter and expresses the
current or voltage as a series of the digits and , which are used to run a digital display.
Most analog meters have been replaced by digital meters.

Ohmmeters

An ohmmeter is an instrument used to measure the resistance of a component or device.


The operation of the ohmmeter is based on Ohm’s law. Traditional ohmmeters contained
an internal voltage source (such as a battery) that would be connected across the
component to be tested, producing a current through the component. A galvanometer was
then used to measure the current and the resistance was deduced using Ohm’s law.
Modern digital meters use a constant current source to pass current through the
component, and the voltage difference across the component is measured. In either case,
the resistance is measured using Ohm’s law , where the voltage is known and
the current is measured, or the current is known and the voltage is measured.

The component of interest should be isolated from the circuit; otherwise, you will be
measuring the equivalent resistance of the circuit. An ohmmeter should never be
connected to a “live” circuit, one with a voltage source connected to it and current
running through it. Doing so can damage the meter.

Ohm’s law and Kirchhoff’s method are useful to analyze and design electrical circuits, pr
oviding you with the voltages across, the current through, and the resistance
of the components that compose the circuit. To measure these parameters, require instrum
ents, and these instruments are described in this section.
DC Voltmeters and Ammeters
Whereas voltmeters measure voltage,
ammeters measure current. Some of the meters in automobile dashboards, digital cameras
, cell phones, and tuner
amplifiers are voltmeters or ammeters. The internal construction of the simplest of these
meters and how they are connected to the system they monitor give further insight into ap
plications of series and parallel connections.

Measuring Current with an

Ammeter

To measure the current through a device or component, the ammeter is placed in series wi
th the device or component. A series connection is used because objects in series have the
same current passing through them.

Ammeters need to have a very low resistance, a fraction of a milliohm. If the resistance is
not negligible, placing the ammeter in the circuit would change the equivalent resistance
of the circuit and modify the current that is being measured. Since the current in the circu
it travels through the meter, ammeters normally contain a fuse to protect the meter from d
amage from currents which are too high.

Measuring Voltage with a Voltmeter

A voltmeteris connected in parallel with whatever device it is measuring. A parallel conn


ection is used because objects in parallel experience the same potential difference.

Since voltmeters are connected in parallel, the voltmeter must have a very large
resistance. Digital voltmeters convert the analog voltage into a digital value to display on
a digital readout.
You may encounter two types of meters in the physics lab: analog and digital. The term ‘
analog’ refers to signals or information represented by a continuously variable physical q
uantity, such as voltage or current. An analog meter uses a galvanometer, which is essenti
ally a coil of wire with a small resistance, in a magnetic field, with a pointer attached thos
e
points to a scale. Current flows through the coil, causing the coil to rotate. To use the galv
anometer as an ammeter, a small resistance is placed in parallel with the coil. For a voltm
eter, a large resistance
is placed in series with the coil. A digital meter uses a component called an analog-to-
digital (A to D) converter and expresses the current or voltage as a series of the digits 0 a
nd 1, which are used to run a digital display. Most analog meters have been replaced by d
igital meters.
The physical quantities of interest in continuous current (DC) circuits are the voltage of
nodes and the current flowing through branches. In addition, for the circuit analysis, the
knowledge of the resistance value of its components is necessary. The instruments used to
measure current intensity are the ammeters; those to measure voltage are
the voltmeters while the ohmmeters provide a direct measurement of component
resistance. Instruments capable of all these three functions, and many others needed for
circuits with time-dependent currents, are known as multimeters and are available in all
laboratories. These devices can be realized in many ways and, broadly speaking, are
categorized as analog or digital instruments. The first kind provides the value of the
physical quantity through an index that can move continuously on a graduated scale, the
second kind provides the result of measurements as a number appearing on a digital
display.

A multimeter is a handy tool that you use to measure electricity, just like you would use a
ruler to measure distance, a stopwatch to measure time, or a scale to measure weight. The
neat thing about a multimeter is that unlike a ruler, watch, or scale, it can
measure different things — kind of like a multi-tool. Most multimeters have a knob on
the front that lets you select what you want to measure. Below is a picture of a typical
multimeter. There are many different multimeter models; visit the multimeter gallery for
labeled pictures of additional models.

What can multimeters measure?

Almost all multimeters can measure voltage, current, and resistance. See the next
section for an explanation of what these terms mean, and click on the Using a
Multimeter tab, above, for instructions on how to make these measurements.

Some multimeters have a continuity check, resulting in a loud beep if two things are
electrically connected. This is helpful if, for instance, you are building a circuit and
connecting wires or soldering; the beep indicates everything is connected and nothing has
come loose. You can also use it to make sure two things are not connected, to help
prevent short circuits.

Some multimeters also have a diode check function. A diode is like a one-way valve that
only lets electricity flow in one direction. The exact function of the diode check can vary
from multimeter to multimeter. If you are working with a diode and cannot tell which
way it goes in the circuit, or if you are not sure the diode is working properly, the check
feature can be quite handy. If your multimeter has a diode check function, read the
manual to find out exactly how it works.

Advanced multimeters might have other functions, such as the ability to measure and
identify other electrical components, like transistors or capacitors. Since not all
multimeters have these features, we will not cover them in this tutorial. You can read
your multimeter's manual if you need to use these features.
What is voltage, current, and resistance?

If you have not heard of these terms before, we will give a very simple introductory
explanation here. You can read more about voltage, current, and resistance in
the References tab, above. Remember that voltage, current, and resistance are measurable
quantities that are each measured in a unit that has a symbol, just like distance is a
quantity that can be measured in meters, and the symbol for meters is m.

 Voltage is how hard electricity is being "pushed" through a circuit. A higher


voltage means the electricity is being pushed harder. Voltage is measured in volts.
The symbol for volts is V.
 Current is how much electricity is flowing through the circuit. A higher current
means more electricity is flowing. Current is measured in amperes. The symbol
for amperes is A.
 Resistance is how difficult it is for electricity to flow through something. A
higher resistance means it is more difficult for electricity to flow. Resistance is
measured in ohms. The symbol for ohms is Ω (the capital Greek letter omega).
 Technical Note
 The symbol that is used for a unit is usually different than the symbol for
a variable in an equation. For example, voltage, current, and resistance are related
by Ohm's law (see the References tab to learn more about Ohm's law):
 Voltage=Current Resistance
 which is usually expressed as
 V=IR
 In this equation, V represents voltage, I represents current, and R represents
resistance. When referring to the units’ volts, amps, and ohms, we use the
symbols V, A, and Ω, as explained above. So, "V" is used for both voltage and
volts, but current and resistance have different symbols for their variables and
units.

This is very common in physics. For example, in many equations, "position" and
"distance" are represented by the variable’s "x" or "d," but they are measured in the unit
meters, and the symbol for meters is m.

A simple analogy to better understand voltage, current, and resistance: imagine water
flowing through a pipe. The amount of water flowing through the pipe is like current.
More water flow means more current. The amount of pressure making the water flow is
like voltage; a higher pressure will "push" the water harder, increasing the flow.
Resistance is like an obstruction in the pipe. For instance, a pipe that is clogged with
debris or objects will be harder for water to flow through, and will have a higher
resistance than a pipe that is free of obstruction.
What do all the symbols on the front of the multimeter mean?

You might be confused by all the symbols on the front of your multimeter, especially if
you do not actually see words like "voltage," "current," and "resistance" spelled out
anywhere. Do not worry! Remember from the "What are voltage, current, and
resistance?" section that voltage, current, and resistance have units of volts, amps, and
ohms, which are represented by V, A, and Ω respectively. Most multimeters use these
abbreviations instead of spelling out words. Your multimeter might have some other
symbols, which we will discuss below.

Most multimeters also use metric prefixes. Metric prefixes work the same way with
units of electricity as they do with other units you might be more familiar with, like
distance and mass. For example, you probably know that a meter is a unit of distance,
a kilometer is one thousand meters, and a millimeter is one thousandth of a meter. The
same applies to milligrams, grams, and kilograms for mass. Here are the common metric
prefixes you will find on most multimeters (for a complete list, see the References tab):

 µ (micro): one millionth


 m (milli): one thousandth
 k (kilo): one thousand
 M: (mega): one million

These metric prefixes are used in the same way for volts, amps, and ohms. For example,
200kΩ is pronounced "two hundred kilo-ohms," and means two hundred thousand
(200,000) ohms.

Some multimeters are "auto-ranging," whereas others require you to manually select the
range for your measurement. If you need to manually select the range, you should always
pick a value that is slightly higher than the value you expect to measure. Think about it
like using a ruler and a yardstick. If you need to measure something that is 18 inches
long, a 12-inch ruler will be too short; you need to use the yardstick. The same applies to
using a multimeter. Say you are going to measure the voltage of an AA battery, which
you expect to be 1.5V. The multimeter on the left in Figure 3 has options for 200mV, 2V,
20V, 200V, and 600V (for direct current). 200mV is too small, so you would pick the
next highest value that works: 2V. All the other options are unnecessarily large, and
would result in a loss in accuracy (it would be like using a 50-foot tape measure that only
has markings every foot, and no inch markings; it is not as accurate as using a yardstick
with 1-inch markings).

What do the other symbols on the multimeter mean?

You might have noticed some other symbols besides V, A, Ω, and metric prefixes on the
front of your multimeter. We will explain some of those symbols here, but remember, all
multimeters are different, so we cannot cover every possible option in this tutorial. Check
your multimeter's manual if you still cannot figure out what one of the symbols means.

What are the red and black wires (probes)? Where do I plug them in?

Your multimeter probably came with red and black wires that look something like the
ones in Figure 4. These wires are called probes or leads (pronounced "Leeds"). One end
of the lead is called a banana jack; this end plugs into your multimeter (Note: some
multimeters have pin jacks, which are smaller than banana jacks; if you need to buy
replacement probes, be sure to check your multimeter's manual to find out which kind
you need). The other end is called the probe tip; this is the end you use to test your
circuit. Following standard electronics convention, the red probe is used for positive, and
the black probe is used for negative.

Although they come with two probes, many multimeters have more than two places in
which to plug the probes, which can cause some confusion. Exactly where you plug the
probes in will depend on what you want to measure (voltage, current, resistance,
continuity test, or diode test) and the type of multimeter you have. We have provided one
example in the images below—and you can check our gallery for a multimeter similar to
yours—but since all multimeters are slightly different, you might need to consult the
manual for your multimeter.

Most multimeters (except for very inexpensive ones) have fuses to protect them from too
much current. Fuses "burn out" if too much current flows through them; this stops
electricity from flowing, and prevents damage to the rest of the multimeter. Some
multimeters have different fuses, depending on whether you will be measuring high or
low current, which determines where you plug the probes in. For example, the multimeter
shown in Figure 5 has one fuse for 10 amps (10A) and one fuse for 200 milliamps
(200mA).

Using a Multimeter

Do you have a multimeter but are confused about how to use it or are getting unexpected
readings? If so, the sections below will help you sort through what to do. If there are
words or concepts you do not understand, or symbols on your multimeter that puzzle you,
return to the Multimeter Overview tab. If you are looking for multimeter usage ideas or
labeled photographs of assorted multimeter models, then visit the other tabs in this
multimeter tutorial.

How do I measure voltage?

To measure voltage, follow these steps:

1. Plug your black and red probes into the appropriate sockets (also referred to as
"ports") on your multimeter. For most multimeters, the black probe should be
plugged into the socket labeled "COM," and the red probe into the socket labeled
with a "V" (it might also have some other symbols). Remember to check out our
image gallery, the Multimeter Overview tab, or your multimeter's manual if you
have trouble identifying the right socket.
2. Choose the appropriate voltage setting on your multimeter's dial. Remember that
most battery-powered circuits will have direct current, but the setting you select
will depend on the science project you are doing. If you are working with a
manual-ranging multimeter, you can estimate the range you need based on the
battery (or batteries) powering your circuit. For example, if your circuit is
powered by a single 9V battery, it probably does not make sense to select the
setting for 200V, and 2V would be too low. If available, you would want to select
20V.
3. Touch the probe tips to your circuit in parallel with the element you want to
measure voltage across (refer to the Multimeter Overview tab for an explanation
of series and parallel circuits). Be sure to use the red probe on the side connected
to the positive battery terminal, and the black probe on the side connected to the
negative battery terminal (nothing will be harmed if you get this backwards, but
your voltage reading will be negative).
4. If your multimeter is not auto-ranging, you might need to adjust the range. If your
multimeter's screen just reads "0," then the range you have selected is probably
too high. If the screen reads "OVER," "OL," or "1" (these are different ways of
saying "overload"), then the range you have selected is too low. If this happens,
adjust your range up or down as necessary. Remember that you might need to
consult your multimeter's manual for specifics about your model.

How do I measure current?

To measure current, follow these steps:

1. Plug your red and black probes into the appropriate sockets (also referred to as
"ports") on the multimeter. For most multimeters, the black probe should be
plugged into the socket labeled "COM." There might be multiple sockets for
measuring current, with labels like "10A" and "mA". Note: It is always safer to
start out with the socket that can measure a larger current. Plug the red socket into
the high-current port.
2. Choose the appropriate current setting on your multimeter. Remember to check if
your circuit is direct current or alternating current, and that almost all battery-
powered circuits will be direct current. If your meter is not auto-ranging, you
might need to guess at the scale to use (you can change this later if you do not get
a good reading).
3. Connect the multimeter probes in series to the current you want to measure (refer
to the Multimeter Overview tab for an explanation of series and parallel circuits).
For example, Figure 7 shows how to measure the current through a lightbulb that
is powered by a battery. Be sure to use the red probe toward the battery's positive
side, otherwise your current reading will be negative.
4. If your multimeter is not auto-ranging, you might need to adjust the range. If your
multimeter's screen just reads "0," then the range you have selected is probably
too high. If the screen reads "OVER," "OL," or "1" (these are different ways of
saying "overload"), then the range you have selected is too low. If this happens,
adjust your range up or down as necessary. Remember that you might need to
consult your multimeter's manual for specifics about your model.

How do I measure resistance?

To measure resistance, follow these steps:

1. Plug your red and black probes into the appropriate sockets on your multimeter.
For most multimeters, the black probe should be plugged into the socket labeled
"COM," and the red probe should be plugged into the socket labeled with an "Ω"
symbol.
2. Choose the appropriate resistance measurement setting on your multimeter's dial.
If you have an estimate for the resistance, you will be measuring (for example, if
you are measuring a resistor with a known value), that will help you pick the
range.
3. Important: Turn off the power supply to your circuit before measuring
resistance. If your circuit has a power switch, you can do this by turning the
switch "off." If there is no switch, you can remove the batteries. If you do not do
this, your reading might be incorrect. If your circuit has multiple components, you
might need to remove the component you want to measure in order to accurately
determine its resistance. For example, if your circuit has two resistors in parallel,
you will have to remove one resistor to measure their resistances individually.

Connect one of your multimeter's probes to each side of the object whose
resistance you want to measure. Resistance is always positive and the same in
both directions, so it does not matter if you switch the black and red probes in this
case (unless you are dealing with a diode, which acts like a one-way valve for
electricity, so it has a high resistance in one direction and a low resistance in the
other direction). Figure 8 shows how to measure the resistance of a lightbulb.

4. If your multimeter is not auto-ranging, you might need to adjust the range. If your
multimeter's screen just reads "0," then the range you have selected is probably
too high. If the screen reads "OVER," "OL," or "1" (these are different ways of
saying "overload"), then the range you have selected is too low. If this happens,
adjust your range up or down as necessary. Remember that you might need to
consult your multimeter's manual for specifics about your model.

How do I do a continuity check?

To do a continuity check (which ensures that there is a conductive path between two
points in your circuit), follow these steps:
1. Set your multimeter to the continuity check symbol. Remember that this symbol
might not look the same on all multimeters (and some multimeters do not have it
at all), so check out the Multimeter Overview tab or our multimeter
image gallery to see examples.
2. Plug your probes into the appropriate sockets. On most multimeters, the black
probe should go into the socket labeled "COM," and the red probe should go into
the same socket you would use to measure voltage or resistance (not current),
labeled with a V and/or an Ω.
3. Important: Turn off the power supply to your circuit before doing a continuity
check. If your circuit has a power switch, you can do this by turning the switch
"off." If there is no switch, you can remove the batteries.

Touch two parts of your circuit with the probes. If the two parts of the circuit are
electrically connected with very little resistance between them, your multimeter should
beep. If they are not connected, it will not make a noise and might display something on
the screen such as "OL," "OVER," or "1," which all stand for "overload." The easiest way
to test this function with your multimeter is to check it with a single piece of conductive
material (most metals) and a piece of non-conducting material, like wood or plastic. How
do I do a diode check?

The diode check feature is useful to determine in which direction electricity flows
through a diode. The exact operation of the "diode check" function will vary for different
multimeters, and some multimeters do not have a diode check feature at all. Because of
this variety, and because the feature is not required for most Science Buddies projects, we
have not included directions here. If you need to do a diode check, consult the manual for
your multimeter.

How do I know which scale to pick for voltage, current, or resistance, and how do I read
the numbers at different scales?

If your multimeter is not auto-ranging, knowing which scale to pick can be tricky,
especially if you are not very familiar with metric prefixes. Here are two rules of thumb
you can follow for measuring voltage, current, and resistance:

 Voltage: Many manual-ranging multimeters have settings for 200mV, 2V, and
20V. It is very unlikely that battery-powered circuits will exceed 20V (for
example, two 9V batteries connected in series will provide a maximum of 18V).
A single AA or AAA battery supplies 1.5V. Two AA or AAA batteries combined
in a battery pack will provide 3V, four will provide 6V, and eight will provide
12V. So, if you know what type of batteries (and how many), are powering your
circuit, you can pick a starting range to measure voltage. Remember that you want
to pick the next highest voltage setting (just like with measuring distance; you
would need a yardstick— not a 12-inch ruler— to measure something that is 18
inches long). So, for a circuit powered by a single AA battery (1.5V), you would
select the 2V setting. For a circuit powered by a 9V battery, you would select
20V.
 Current: When measuring current, it is always a good idea to start out with the
highest possible current setting (and the appropriate high-current socket, if your
multimeter has multiple sockets to measure current), in order to avoid blowing a
fuse. If the current you measure is low enough to safely use your low-current
settings and socket, then you can take a new reading to get a more-accurate
measurement. For example, say your multimeter has a socket with a 10A fuse and
one with a 200mA fuse. Using the 10A socket, you measure a current of 150mA.
Then it would be safe to measure again with the 200mA socket (and a lower
setting on the knob).
 Resistance: If you are measuring an object with a known resistance, you can use
that value to choose the appropriate resistance setting. As with current and
voltage, you need to pick the next largest resistance value on your scale. For
example, to measure a 4.7kΩ resistor, you would select 20kΩ. If you are
measuring an object with unknown resistance, you will just have to guess, but it is
difficult to damage your multimeter or the object you are testing when measuring
resistance, so this is not a big problem.

The same value might appear differently when measured with a different scale selected
on the multimeter dial. For an example, let us use measuring the DC voltage from an AA
battery—which we expect to be 1.5V—using a multimeter that has settings for 200mV,
2V, 20V, 200V, and 600V. When measuring the battery with each setting.

The "1." is this multimeter's way of saying that it is "overloaded"—the value of 1.6V is
outside the selected range of 200mV. Other multimeters might display "OVER" or "OL"
when this happens. Notice that as the range increases, the accuracy decreases. At the 2V
setting, the reading displays 3 decimal places. At the 200V setting, the reading only
displays one decimal place.

You might also need to take metric prefixes into account when reading the number from
the multimeter screen. For example, suppose your screen reads "6.1" when you are
measuring current with the "10A" setting. This means your current measurement is 6.1
amps. However, if the screen reads "6.1" when you have the current dial set to 20mA,
this means you are measuring 6.1 milliamps.

My multimeter is not working! What is wrong?

Do not panic! There are several common mistakes that can be easily fixed.

 Make sure your multimeter has fresh batteries.


 Some multimeters have an auto power-saving feature, and will turn off after a
certain period of inactivity. If this happens, turn your multimeter's dial to "off"
and then turn it on again.
 Make sure you have your probes plugged into the correct ports for what you want
to measure (see the "How do I measure..." sections above).
 Make sure you are connecting your probes to your circuit in the correct manner
(series or parallel) for what you want to measure (see the "How do I measure..."
sections above).
 Make sure you have the correct setting chosen on your multimeter dial for what
you want to measure; for example, if you need to measure DC voltage, make sure
you do not have current, resistance, or AC voltage selected on the dial.
 If your multimeter is not auto-ranging, you might need to manually adjust your
range. If your multimeter screen always reads "0," this might mean the range you
have selected is too high. If it reads "OL," "OVER," or "1," the range you have
selected could be too low. Each multimeter is different, so you might need to read
your multimeter's manual to find out what the display on the screen means. You
can then adjust the range accordingly.
o For example, if you are trying to measure the voltage of a 9V battery, but
have your multimeter set to 2 DCV, this range is too small and you would
have to increase it to a higher value, such as 20 DCV.

How do I know if I need to change the fuse?

Some multimeters have a fuse (or multiple fuses) that will "burn out" when too much
current flows through them, which then prevents more electricity from flowing, and
hopefully saves the rest of the multimeter from damage. In some multimeters, these fuses
can be replaced if they burn out, but instructions for replacing them (and figuring out if
they need to be replaced at all) will vary for different multimeter models.

You will probably need to open your multimeter to access the fuses (Important: Always
disconnect the probes before you do this). Some multimeters have covers that will pop or
slide off, and some have screws that must be removed first. Fuses usually look like small,
glass cylinders with metal caps on the end and a thin wire running down the middle:

If a fuse has burned out, it might be visibly blackened or charred. The wire on the inside
might have completely burned away, and no longer be visible.

How do I change the fuse?

Important: Always disconnect the leads from your multimeter before opening the cover
to change the fuse.

Instructions for changing the fuse vary with each multimeter model, so you will need to
check your multimeter's manual for instructions. This tutorial from Spark Fun provides
directions for changing a fuse on their brand of multimeter, but remember that these
directions might not apply to your model. Note that in some multimeters-especially in
inexpensive ones-you might not be able to change the fuse.

Multimeter Fun
In this multimeter tutorial, we have already covered what a multimeter is and how to use
it. This section will provide some suggestions of practical and interesting things to do
with a multimeter.

What can I do with a multimeter around the house?

Remember that the primary purpose of a multimeter is to test circuits and electrical
components in an experiment or science project that involves electronics. What if you do
not have a circuit to test? Here are a couple suggestions for quick experiments you can do
with a multimeter around your house.

 Test batteries! Have you ever wondered if a device or toy stopped working
because the batteries were dead? With a multimeter, you can make sure batteries
are dead before recycling them by testing their voltage. Remember that fresh AA
and AAA batteries should supply about 1.5V (a fresh alkaline cell will measure
about 1.6V with nothing else attached). However, batteries will effectively be
"dead" long before they reach 0V. For example, if a battery is supplying 0.7V, that
probably is not enough to power most household devices that are expecting 1.5V
from the battery.
 Do you have rechargeable batteries? You can test their voltage over time as they
recharge, and then make a plot of voltage vs. charge. How long does it take for the
voltage to stop increasing? Is the graph a straight line?
 Do you have anything you can take apart with a circuit board inside, like an old
toy, or a TV remote? Use the continuity check on your multimeter (if it has one)
to test which parts of the circuit are directly connected to each other.
(Warning: Old circuit boards are constructed with a lead-based solder, which is
toxic. Always wash your hands carefully after handling solder, and check your
local waste disposal guidelines to see if there are special rules about disposing of
lead as hazardous waste.)
 Set your multimeter to measure resistance, and have everyone in your house take
turns grabbing the metal tips of the probes (one in each hand). Who has the
highest resistance? The lowest?
 Use the continuity check or resistance measurement to test different materials in
your house. Which ones are conductors and which ones are insulators?
 Hook the multimeter leads up to a speaker using speaker wire, and set the
multimeter to measure AC amps (or AC volts if AC amps is not available). Who
can get the multimeter to display the biggest number by yelling into the speaker?
(In this case, the speaker is working like a microphone, generating current when it
detects sound.)

What shouldn’t I do with my multimeter?

 Do not use a multimeter to test electricity from the wall outlets in your home.
Electricity from wall outlets is very dangerous and can be fatal.
 Do not connect the probes directly to a battery or other power supply when you
have a "current" measurement setting selected. This will cause a "short circuit"
across the battery terminals, and a very high current will flow through your
multimeter. This will probably blow a fuse, or possibly damage the multimeter.

Which Science Buddies projects require a multimeter?

A lot of Science Buddies projects require a multimeter; it is a handy tool! In fact, there
are too many to list, but here is a sampling of a few that cover a variety of topics:

 Electrolyte Challenge: Orange Juice Vs. Sports Drink. This project compares the
number of electrolytes in a sports drink to the amount in typical orange juice. Do
sports drinks live up to their promise? Use a multimeter to find out how well these
different liquids conduct electricity.
 How to Turn a Potato into a Battery. Did you know that you can make a battery
and even light up LEDs with everyday fruits and vegetables? Use a multimeter to
test the current and voltage that can be supplied by your afternoon snack.
 Water to Fuel to Water: The Fuel Cycle of the Future. Help save the planet with
this green energy project! Use fuel cells to store energy produced by solar panels
for use when it is dark out, and use your handy multimeter to keep track of your
power generation.
 How Bright Is Your Glow Stick? Measure It!. Using a light sensor, a simple
circuit, and a multimeter, you can measure the amount of light given off by glow-
in-the-dark objects.
 Spice Up the Power of a Microbial Fuel Cell with a Dash of Salt. Does generating
electricity from mud sound crazy? Believe it or not, with help from some friendly
bacteria, you can turn an everyday scoop of dirt into a battery. Use a multimeter
to help design this green-energy fuel cell.
 Wily Waves: Build an Oscillating Water Column to Extract Energy from Ocean
Waves. Harnessing the power of the ocean is yet another green energy technology
that could help reduce our dependence on fossil fuels. Use a multimeter to
measure the electrical power generated by your oscillating water column.
 A Battery That Makes Cents. Did you ever think you could make a battery out of
pocket change? In this project you will make a battery out of pennies and nickels,
and use a multimeter to measure the current and voltage that it can produce.

Multimeter Gallery

While most multimeters can perform the same basic functions, different models that are
made by different manufacturers might not all look the same. In the gallery below, we
have provided a series of images of different multimeters with different measurement
settings and sockets for the probes labeled. Note that most of the multimeters have basic
features in common, including settings for measuring voltage, current, and resistance. All
of them have a single "ground" socket for the black probe. Most of them have separate
sockets for measuring high and low current. The low-current socket is also used to
measure voltage and resistance. However, some multimeters only have two sockets total,
or only have one socket for measuring current. Some also have additional features that
we did not label.
Remember, this gallery is meant to be a general guide; if you do not see your model of
multimeter pictured here, your best bet is to consult your specific multimeter's manual. If
you need help getting introduced to multimeters in general, refer back to our Multimeter
Overview section. If you need to know how to take a specific type of measurement, refer
to the Using a Multimeter section.

How to measure current with a clamp accessory


Multimeters, Clamp meters

Often viewed principally as a voltage-measuring device, a digital multimeter equipped


with a clamp accessory can quickly measure current.

1. Determine if the current to be measured is ac or dc.


2. Select a clamp accessory for your digital multimeter that is designed to measure
that specific current or one that can measure both ac and dc.

Note: Look at the accessory clamp’s specifications and determine whether the
clamp outputs a current level or a voltage level.

3. Determine the circuit’s anticipated maximum current by checking the nameplate


of a component or the breaker rating. Plug-in clamp accessories are available in a
variety of preset ranges. Determine if the range of your multimeter or clamp
accessory is high enough to measure it. If not, select an instrument equipped for
higher ranges.

Note: If a meter includes fused current terminals, verify that its fuses are good.

4. Set up your DMM as follows:


o To measure ac current with a current output clamp, turn the dial to
man/Ã.
o Plug the black test lead into the COM jack.
o For plug-in clamp accessories that produce an ac current output, plug the
red test lead into the man/Ã jack. These clamps are designed to measure ac
current only and, depending on the clamp’s scaling factor, deliver 1 mA to
the DMM for every 1 A of measured current (1 mA/A).
o Follow steps 6-8 below.
o To measure ac/dc current with a voltage output clamp, turn the dial to
man for ac current, or to medic for dc current.
o Plug the black test lead into the COM jack.
o For plug-in clamp accessories that produce a voltage output, plug the red
test lead into the V jack. These clamps are designed to deliver 1 mV, 10
mV or 100 mV to the DMM for every 1 A of measured current.
o Follow steps 6-8 below.
5. Open the jaws by pressing the tool’s trigger.
6. Enclose a single conductor inside the jaws. Make certain the jaws are completely
closed before taking readings.
7. View the reading in the display.

Tip: Clamps measure current in a circuit by measuring the strength of the


magnetic field around a single conductor. Whenever possible, separate the test
conductor from surrounding conductors by a few inches. The goal: Prevent the
clamp from picking up stray magnetic fields. If separation is not possible, take
several readings at different locations along the same conductor. Do not measure
shielded conductors, as the magnetic fields are greatly diminished or even
eliminated.
Current measurement analysis
Knowing the current consumption in a system, component or circuit is very helpful
during troubleshooting.
Electrical components such as motors often carry a nameplate that displays the
component’s circuit rating. A current measurement can be compared to that rating to
determine the health of the component’s operating condition.
Take current measurements: to determine how much a load (a component, such as a
motor) is drawing from the system. You can also measure the total load on a circuit.
A motor, for example, is overloaded if it draws more than the rated current, underloaded
if it draws less.
When troubleshooting, a technician can take a baseline measurement and be on the
lookout for overload, overcurrent, or a current imbalance between phases.
Generally, higher-than-rated currents usually indicate a problem which can cause
additional problems. Higher current produces higher temperature, and that may cause
insulation breakdown and component failure.
Most digital multimeters can only measure dc or ac current up to 10 A. Higher current
must be scaled down with a current clamp accessory, which can measure current in a
circuit from .01 A to 1000 A by measuring the strength of the electric field around a
conductor.
For maximum efficiency, it is recommended that a current measurement be taken when
equipment is first installed and during normal operation. These measurements can be
used to provide a baseline comparison when troubleshooting a future problem.

Circuit breakers are the standard safety devices installed in your electrical service panel
to protect your electrical system from overloads and other hazards. They are also used as
convenient shut-off points for electrical circuits, allowing you to shut off the power when
you need to make repairs or upgrades to a circuit. But unless your electrical
service panel (circuit breaker box) is very carefully mapped, it can be difficult to identify
which breaker controls which outlets, switches, or light fixtures.

Although it sounds easy, mapping all the breakers in a service panel can take two people
several hours. A great little device that simplifies this task is an electronic circuit breaker
finder, which easily locates a circuit in a breaker box and makes this all a one-person job.
This device works for either circuit breakers or fuses and is very easy to use.

What Is a Circuit Breaker Finder?

A circuit breaker finder is a small battery-operated electronic tool with two parts—a
transmitter and receiver. It is designed for single purpose—to identify which circuit
breaker in a main service panel serves a particular outlet or light fixture to which the
tool's transmitter is connected. It makes the task of matching circuit breakers with their
wiring circuits much easier.

Parts of a Circuit Breaker Finder

An electronic circuit breaker finder includes two parts: a transmitter and a receiver. The
transmitter plugs into a household outlet (or light socket, using an adapter) for which you
are trying to identify the controlling circuit breaker or fuse. A faint electronic signal is
sent through the circuit wires.

At the breaker box, you use the electronic receiver that is paired with the transmitter. When
the receiver passes over the circuit breaker that carries the electronic signal from the
transmitter, the receiver rapidly beeps and flashes. It is as simple as that.

How to Use a Circuit Breaker Finder

1. Install the Battery

The receiver on circuit breaker finders is powered by a battery—often a 9-volt


battery installed in the handle. The battery compartment is easily accessed by
sliding a cover back on the bottom of the receiver.

After installing the battery, adjust the receiver for maximum sensitivity. With the
model shown here, you simply rotate the receiver's wheel back until it clicks and
the LED lights up.

2. Test the Receiver and Transmitter

The next step is to confirm that the receiver will pick up the transmitter's signal.
Plug the transmitter into a wall outlet. The outlet power should be on, and the
transmitter should light up, indicating power. Place the receiver near the
transmitter. The receiver should light up and/or beep to indicate it has picked up
the transmitter signal. You can now use the device to locate the matching circuit
breaker in your breaker box.

3. Match the Circuit to Its Breaker

Open to the door to the main circuit breaker box. Hold the receiver so the sensor
tip is at a right angle and directly on the face of a circuit breaker. Slowly move the
receiver up and down over the rows of circuit breakers while continually lowering
the receiver’s sensitivity until only one breaker or fuse causes the receiver to
beep. (Note: Sometimes an adjacent breaker or fuse may cause a beep due to the
routing of the wires in the panel.)

After you have located the correct breaker or fuse, you can turn off the circuit by
switching the breaker to the "off" position. The receiver will stop beeping.

Return to the outlet where the transmitter is plugged in to check that the
transmitter light is off, indicating the circuit is no longer energized. This confirms
you have turned off the correct breaker.

4. Using a Circuit Breaker Finder with a Light Fixture

Because the circuit breaker finder's transmitter must plug into an outlet, you can
use an inexpensive adapter, called a keyless socket adapter, to plug the transmitter
into a light fixture socket.

To use an adapter, turn off the light fixture at the wall switch, and remove one of
the light bulbs on the fixture. Screw the adapter into the light bulb socket. Plug
the transmitter into the adapter, and turn on the light switch. The transmitter will
function just as if it were plugged into an outlet. Follow the same steps listed
above for wall outlets to check for power, transmitter, and breaker.

In conclusion, direct current (DC) stands as a foundational and versatile form of


electrical power that has played a pivotal role in the evolution of modern
technology. While alternating current (AC) dominates in power distribution
systems for its efficiency in long-distance transmission, DC remains essential in
numerous applications where a constant and unidirectional flow of electricity is
required. The simplicity and reliability of DC make it a preferred choice for
electronic devices, batteries, and specific industries, showcasing its enduring
significance.
The continued relevance of DC is evident in its integration with emerging
technologies. Rectifiers and inverters, for instance, enable the seamless
conversion between AC and DC, allowing for the efficient utilization of power
from diverse sources, including renewable energy. The precise control capabilities
of DC motors contribute to their widespread use in various industrial applications,
underscoring the adaptability and efficiency of direct current in meeting the
demands of modern machinery.

Furthermore, the recent resurgence of interest in high-voltage direct current


(HVDC) transmission highlights DC’s potential to address challenges in long-
distance power transmission. As technology advances, DC continues to play a
critical role in shaping the landscape of electrical power systems, fostering
innovation, and contributing to sustainability efforts. The coexistence of AC and
DC in today’s interconnected world demonstrates the complementary nature of
these two forms of electrical power, each serving specific needs and driving
progress in their respective domains.

In essence, direct current’s journey from its early roots in electrical engineering to
its contemporary applications reflects its enduring importance. As our dependence
on electricity grows and technologies evolve, the role of DC is likely to expand,
influencing the way we generate, transmit, and utilize electrical power in the years
to come.
CamScanner
CamScanner
CamScanner
CamScanner

You might also like