You are on page 1of 27

ELECTRICAL RESISTANCE STRAIN GAGES

José L.F. Freire


Pontifical Catholic University of Rio de Janeiro – PUC-Rio

Key Words: strain, measurement, circuits, adhesives, Wheatstone bridge, potenciometer circuit,
sensors, transducers, load cells, applications.

Contents
1. Introduction
2. Sensitivity of a thin metallic conductor to strain
3. Strain gage conductor materials
4. Characteristics of modern strain gages
5. Other topics concerning modern strain gages
5.1 Conductor and strain gage sensitivities
5.2. Temperature response of strain gage installations
5.3. Gage response in non-uniform strain fields
5.4. Gage response in dynamic strain fields
6. Strain gage circuits
6.1 Potenciometer circuit
6.2. Constant voltage Wheatstone bridge circuits
6.3. Constant current Wheatstone bridge circuits
6.4. Constant voltage Wheatstone bridge circuits – specific application cases
6.5. Shunt Calibration
6.6 Quarter bridge: three-wire connection of a strain gage
6.7 Null balance bridges
7. Strain gage measurement systems
8. Examples of strain gage applications
9. Conclusions
Glossary
Bibliography

Summary

Electrical resistance strain gages are sensors fabricated from thin foil or wire-type conductors that
respond to variations in their length with variations in their electrical resistance. Strain gages are used
to measure linear strains that occur at surface points of an object when it responds to an actuating load,
Strain gages are either bonded with adhesives to the surfaces of structures or are welded on. Strain
gages are used to determine strains in localized areas of structural components in the laboratory, in the
field or as sensors in transducers such as resistive accelerometers or load cells. They are accurate
“point” measuring elements and can be used in cases of static or dynamic loading. The measurement
systems employ conventional and in-shelf circuitry elements, and have the important advantage of
leading to electrical and treatable output. Circuitries commonly used are the potentiometer and the
Wheatstone bridge circuit.
1. Introduction

Electrical resistance strain gages are sensors made of thin foil or wire-type conductors that respond to
variations in length with variations in electrical resistance.

Strain gages are used to measure linear strains that occur at surface points of an object when it
responds to some actuating load, as shown in Figure 1. This figure shows a surface point on the object
before and after a load was applied. The strain gage is bonded to the surface with an adhesive.
Deformation of the surface element forces the strain gage to change its length. For the special
conductor gage materials, the variation in length of the parallel segments of the wire-type conductor
will be directly proportional to the variation in electrical resistance of the conductor. Equation (1)
shows the relationship between the variation in resistance of gage ∆R and strain ε to be determined,
where K is the gage factor. The static or dynamic variation in resistance is measured and registered by
an auxiliary circuitry. Notation of variables used in equation (1) is presented in Figure 1.

Figure 1: Sketch of a strain gage bonded to the surface point of an object before and after loading was
applied

R f − Ri L f − Li ∆R
= K. ∴ = K .ε (1)
Ri Li R

Figure 1 and equation (1) show that strain gages measure the relative displacement between two points
located at the extremities of their grids with changes in resistance. These displacement changes can be
so small that other mechanical devices would not be able to measure them with the necessary
resolution and accuracy.

An example of the necessary resolution is given in Figure 2. It shows a prismatic object loaded with an
axial force P, as occurs, for example, in a tensile test used to measure the elastic and plastic mechanical
properties of a structural material. Inside the elastic range, the uniaxial stress-strain relationships given
in equations (2) hold, where E and µ are the Young modulus and Poisson coefficient, respectively. For
low carbon steel – for example, ASTM A 36 – the yield strength is 250MPa or greater. Using
E=200GPa, µ=0.3, load P=12.5 kN and assuming that the cross section area of the prismatic object is
square with sides equal to w=10mm, it can be calculated that the tensile stress is σ=125MPa and the
linear strain measured by the strain gage is ε=612.5 x 10-6. The relative displacement occurring
between two points, A and B, located at the surface of the prismatic object and far apart at L=10mm
will be ∆L=0.0006125mm. These numbers show that working stresses as high as half the yield strength
of the material of a structural member will cause very small displacements in a relatively large gage
length of 10mm. Any device or method employed to measure this displacement will need a resolution
of 1µm or at least 10µm to indicate strain steps of 1 to 10x10-6. Electrical resistance strain gages can be
used with this resolution and with an accuracy varying from numbers as low as 1% to as high as 0.1%,
depending on the set ups and circuitry used in each measurement.

Figure 2: Elastic stress-strain uniaxial mechanical behavior relations

⎧ σx
⎪⎪ε x= E
⎨ (2)
⎪ε = − µε = − µ σ x = ε
⎪⎩ y x
E
z

2. Sensitivity of a thin metallic conductor to strain

Figure 3 shows a thin wire-type conductor subjected to a normal positive or negative change in length.

Figure 3: Thin wire-type conductor being strained


Resistance R depends on resistivity ρ, length L and area A, as shown in equation (3). The relative
change in resistance is given by equation (4) in terms of relative changes of variables ρ, L and A.
L
R = ρ. (3)
A
dR dρ dL dA (4)
= + −
R ρ L A

Term dL is the instantaneous strain.


L

dL
dε = (5)
L

If the conductor has a circular cross section with diameter D, the infinitesimal variation of the area is:

π .D 2 dA dD
A= → = 2.
4 A D

The relative variation of the area can be written in terms of the infinitesimal strain and the Poisson
coefficient:

dD dA
= dε y = − µ .dε x ∴ = −2.µ .dε
D A

The relative change in resistivity is assumed to be proportional to the relative change in volume, the
proportionality coefficient designated as c, the Bridgeman constant.

dρ dV
= c. (6)
ρ V

The change in volume of the conductor is written as:

V = L. A
dV dL dA
= + = dε − 2.µ .dε
V L A

Therefore, it is possible to determine the dependence of the relative change in resistance R in terms of ε
as:

= [c(1 − 2.ν ) + 2.µ + 1].dε


dR
R (7)
dR
= K .dε
R
Integration of (7) gives (8) if K is considered constant:

Rf Lf
ln = K . ln (8)
Ri Li

Equation (9) will be valid if the total strain value is low and if coefficients c and µ remain constant.

dL Lf ∆L ∆L
ε = ∫ dε = ∫ = ln ≅ ≅
L Li Li Lf

∆R
= K .ε (9)
R

In the case of c being equal to unity, which happens for a conductor made of the alloy Constantan
(55%Cu + 45%Ni), proportionality constant K will be equal to 2 and will not depend on the change in
value of µ in the elastic-plastic transition.

3. Strain gage conductor materials

Electrical conductor materials for strain gages must possess the capability of being formed as thin
wires or very thin metal foils, and must have constant K values over a wide range of elastic (and plastic
if possible) strains and temperature. The ones that are used the most are listed in Table 1. Constantan
and Karma alloys are the ones used the most in general applications within temperature ranges of [-30
o
C,200oC] and [-50 oC,250oC], respectively.

Table 1: Stain gage conductor materials


Material Alloys K Application
Advance or 45% Ni – 55% Cu General use up to 8% strain. Widely
2.1
Constantan used
74% Ni - 20% Cr – 3% Al – Wide range for temperature
Karma 2.0
3% Fe compensation, high fatigue strength
36% Ni – 8% Cr – 0.5%Mo –
Isoelastic 3.6 General use. Temperature sensitive.
55.5% Fe
Nichrome V 80% Ni – 20% Cr 2.1
Platinum- 92% Pt – 8% W High temperature applications (may be
4.0
Tungsten over 250oC)
Armour D 70% Fe – 20% Cr – 10%Al 2.0

4. Characteristics of modern strain gages

The first strain gages were made from thin wire conductors that were bent to form several rows of
parallel sensing legs to produce the sensing grids. These grids were cemented to thin special china
paper backings to allow for electrical isolation and easy handling and mounting.

Modern strain gages are made from thin conductor foils (e.g. Constantan foils) backed by flexible and
non-conducting thin foils of polyamide, phenol or epoxy resins. The strain sensitive grids are produced
by a photoresistive etching process. The backings give some stiffness to the foil grids to make their
handling process relatively easy and to electrically isolate the grids from the prototype metallic
materials.

The adhesives used to bond the strain gages to the prototype play an important role in the measurement
system. They must be easy to apply, be compatible with both the prototype and backing materials, have
a linear response to strain along the entire range of the measurement, and be time independent.

Figure 4 and Table 2 summarize most of the technological firsthand information needed to give sound
and important information to a beginner. Circuitry for electrical measurements of resistance variation
∆R are commented on in section 3 of this topic.

Figure 4: Sketch of a strain gage layout

The process of bonding a gage is described in Table 3. Of course, there are other processes and
adhesives for more specific applications such as strain gage bonding for transducers, high temperature
applications, or large strain measurements.
Table 2: Characteristics of modern strain gages
Topic Description
Lord Kelvin worked on the basics in 1856.
History First wire grid gages were presented in mid-1930, by Ruge and Simmons independently of each other. Foil
gages were first presented by Sanders and Roe after 1952.
Grid materials Advance or Constantan, Karma alloy, Isoelastic, Nichrome V, Armour D and Pt-W (see Table 1)
Backing or carrier
Thin paper, polyimide film, epoxy film, glass fiber epoxy-phenolic reinforced polymer film
materials
Epoxy (more stable, used in transducer applications), cyanoacrylate (very easy to use, general application up
Adhesive Materials
to 65oC), polyester, ceramic (high temperature applications)
Uniaxial Biaxial rosette Triaxial stacked Triaxial 45 degree
rosette rosette

Strain gage
configurations

Weldable Linear array High temperature Residual stress rosette


(weldable)

Special types of strain


gages

Embedded (concrete) Concrete

Combined effect of time of exposure to strain, temperature and humidity on the gage installation and
Stability measurement system can affect stability. Zero drift can be as low as a few micro-strains to several hundred,
depending on foil and carrier material, adhesive, installation cleanliness, and stability of instrumentation.
Gage Factor Constantan gages have gage calibration factor K around 2.0.
Depends on gage geometry and type.
Transverse sensitivity
Kt = St/Sa = 0 to 5%, usually < 0.5%
Temperature
Available for prototype materials with expansion coefficients varying from zero to 26x10-6/oC
compensation

Grid dimensions vary in such a way that the most suitable geometry
for a specific application will be found most of the time. There are
Grid dimensions
different W/L ratios. Minimum and maximum L range from 1.0mm
(stress concentration) to 150mm (concrete) applications.

Common: 120 and 350Ω


Initial resistance
Possible: 500, 1000, 2000, 5000Ω
Gage resistance = +0.3%, Gage factor= +1%. Transducers that employ strain gages have a much better
Accuracy
accuracy due to calibration procedures and temperature corrections
Resolution 1µε. Can be lower or higher, depending on the circuitry and measurement range and devices
Range + 1%, special gages with annealed foils may work up to 20%
Circuitry Potenciometer and Wheatstone circuits. Four wire resistance direct reading circuit
Static to dynamic range. Dynamic measurements may require several MHz acquisition rate. Circuitry,
Acquisition rate and
amplifiers and reading-recording devices limit strain gage installation capabilities to measuring 100 ns
dynamic range
pulses.
Temperature
Humidity
-40oC<θ<60oC general use- polyimide carriers and
Must be protected to Radiation
cyanoacrylate adhesive.
o o avoid degradation of Large apparent strains
-195 C<θ<230 C Karma alloy fully encapsulated
backing and adhesive are induced with time
and epoxy adhesive.
and undesirable loss of (zero drift)
Influence of other Special high temperature gages - PtW alloy -
o isolation
parameters weldable or bonded with ceramic adhesive (~600 C)
Magnetic fields
Fatigue and strain cycling Hydrostatic pressure
5 High field gradients
Fatigue strength varies from +1500x10 to Minimal influence
influence the response of
+2200x108, depending on the gage type. Fatigue (20ηε) up to about
the gage and associated
strength depends on the quality of the installation 20MPa pressure
cabling
Widespread: experimental stress analysis, transducers (accelerometers with DC response and load cells),
Application application to integrity monitoring, control of mixtures in chemical processes, weight control and
measurement
Table 3: Steps for bonding, cabling installation, quality control and protection of strain gage
installations

Step Description
Selection of points where to measure strains, characteristics of measurement:
Measurement system static or dynamic, low or high temperature, low or high strain gradient,
design general stress analysis or transducer application, etc. Selection of gage,
selection of adhesive and protection.
Locate point at surface of prototype.
Degreasing entire bonding and adjacent area. Check compatibility of
prototype exposed surface with degreaser to avoid undesirable chemical
attack.
Thorough cleaning of surface, taking off all paint and oxidation. Expose
surface to bare metal.
Use hand sand paper up to number 220.
Clean bonding area completely with clean cotton and Freon, alcohol or
acetone.
Example: bonding With hard lead pencil, mark position where to set gage.
strain gage with Place gage with adhesive tape and check for position.
cyanoacrylate Pull back adhesive tape just enough to expose back side of gage that will
receive adhesive.
Put small drop of adhesive on surface of gage and reposition adhesive tape to
previous well-positioned place. Press gage against prototype surface using
tip of finger on adhesive tape for about two minutes. A small silicone rubber
pad may be used to spread pressure over gage and protect finger skin from
excess adhesive.
Peel off adhesive tape. Clean surrounding gage area of excess adhesive.
Cement solder auxiliary tabs with cyanoacrylate.
Solder lead wires. Use 20 to 30W soldering iron.
Measure resistance of gage and circuit using ohmmeter.
Check electrical
Measure resistance between gage and surface of metallic prototype.
contacts
Resistance should be above 500 MΩ.
Using one or more coats, protect gage, electrical contacts and portion of lead
Provide protection to
wire that reaches installation. These should be selected to protect against
installation
electrical contact with environment, humidity and touch.

The high resistance of the gages is achieved with the use of relatively long and thin etched conductors.
To make them relatively small in size, the gages are constructed with several parallel thin conductors
that form a grid of dimensions L and W as showed in Figure 5. The initial resistance of normally
marketed strain gages is 120 or 350Ω. Strain gages may also be found with resistances as high as 1,000
or 2,000Ω to be applied in special measurements; for example, when direct measurement of ∆R is
required.
Figure 5: Sketch of strain gage showing grid dimensions, axial (measurement direction) and transverse
direction.

The terminal tabs used for soldering the cables or for connecting the parallel conductors that form the
grid are made much thicker and wider. This procedure helps to decrease their resistance and, therefore,
their variation under strain. As a consequence, almost 100% of the resistance variation is caused by the
straining of the thin parallel conductors, and the spurious response of the gages to transverse strains is
usually very low.

Figure 6: Constantan foil strain gage with polyimide backing. Original size is L = 75 mm to be applied
in concrete structures. Note the large, thick tabs for soldering the connecting cables and the 180o
turning corner connections of the wire-type parallel foil conductors.

Manufacturers of commercial strain gages state the strain gage calibration factor K and the transverse
sensitivity Kt (see item 5.1). Approximate values of K are given in Table 1. The calibration of gage
factor K follows standardized procedures. The value of K is determined with the help of an
instrumented cantilever steel beam with a 0.285 Poisson coefficient.

5. Other topics concerning modern strain gages

This section discusses topics that deserve attention concerning measurements where modern strain
gages are used.

5.1 Conductor and strain gage sensitivities

The relative variation in the gage resistance is given in equation (10), where Sa, St and Sat are the
sensitivities of the gage to the strain state (see Figure 5 for directions a and t).

∆R
= Sa.ε a + St.ε t + Sat.γ at Sat << St << Sa (10)
R
In light of the fact that Sat is very small when compared with Sa and St, and defining the transverse
sensitivity as Kt = St/Sa, equation (11) is obtained.

∆R ⎛ St ε ⎞ ⎛ ε ⎞
= Sa.ε a .⎜⎜1 + . t ⎟⎟ = Sa.ε a .⎜⎜1 + K t . t ⎟⎟ (11)
R ⎝ Sa ε a ⎠ ⎝ εa ⎠

In the beam calibration test, the apparent axial strain indicated by the gage ε aa will be given by
equation (12) as

∆R ⎛ ε ⎞
= Sa.ε a .⎜⎜1 + K t . t ⎟⎟ = K .ε aa (12)
R ⎝ εa ⎠

Noting that the uniaxial stress state in the beam gives ε t = −0.285.ε a , the gage factor K will be given
by (13) as

K = Sa.(1 − 0.285K t ) (13)

In actual experiments, the transverse sensitivity is neglected most of the time because the Kt value of
modern gages is very low (<0.5%). In cases where the errors associated with the existence of the
transverse sensitivity are considered relevant, two strain gages mounted along orthogonal directions
must be used to determine the apparent strains in the axial and transverse directions. The actual strains
will be determined by using equations (14) and the Kt value of the strain gages.

1 − 0.285K t a
εa =
1 + Kt
(
ε a − K t .ε tt )
(14)
1 − 0.285K t t
εt =
1 + Kt
(
ε t − K t .ε aa )

The error associated with the measurement of strain along the prescribed gage axial direction is given
by equation (15).

⎛ε ⎞
K t ⎜⎜ t + 0.285 ⎟⎟
⎛ε −ε ⎞ a
ε
Error ⎜⎜ a ⎟= ⎝ a
a ⎠ (15)

⎝ εa ⎠ 1 − 0.285 K t

As an example of the error measurement given, assume a thin walled cylindrical pressure vessel
(external diameter D and wall thickness t) loaded with internal pressure p. Stresses and strains in the
axial (longitudinal l) and transversal (circunferencial c) directions and their ratios are given by
equations (16), where a Poisson coefficient µ=0.3 is used.
pD pD σl
σc = , σl = ∴ = 0.5
2t 4t σc
(16)
εc =
pD
(1 − 0.5µ ) , εl =
pD
(0.5 − µ ) ∴ ε l = 0.5 − µ = 0.23
2tE 2tE ε c 1 − 0.5µ

Considering a transverse sensitivity equal to 5% (a value that may be considered very high for modern
strain gages), the errors generated in considering the apparent strains as the actual ones in the
circumferential and longitudinal directions will be 2.6% and 23%, respectively. The errors generated in
calculating the circumferential and longitudinal stresses considering the apparent strains will be 3.6%
and 12%, respectively.

5.2. Temperature response of strain gage installations

Strain gages installations are temperature dependent and their response depends on the variation of
temperature ∆θ that occurs during the test time period. Equation (17) shows that an apparent strain
arises due to ∆θ, where α and β are expansion coefficients of the gage and prototype materials,
respectively, and γ reflects the dependence of the material’s resistivity to temperature.

R = R (ε , θ ), ρ = ρ (ε , θ ), l = l (ε , θ ), A = A(ε , θ )
dR dρ dl dA
= + −
R ρ l A
⎛ ∆R ⎞
∴ ε aparent = ⎜ ⎟ = (β − α ).K .∆θ + (γ − α ).∆θ (17)
⎝ R ⎠θ

Modern strain gages are fabricated to be temperature compensated for the most-used prototype
materials such as carbon steel, stainless steel, aluminum alloy, glass and polymer, to name a few. Even
if constructed in this way, they can still show some apparent strain when large variations in
temperature occur. Even small changes in temperature can cause large apparent strains if the initial
temperature is relatively far from room temperature (temperature compensation setup). Strain gage
dealers furnish charts or appropriate equations to correct for the apparent strains generated by
temperature variations. Sometimes, the user can calibrate the measurement system beforehand by
varying the temperature conditions of the test setup. On other occasions it is possible to take advantage
of the circuitry used to allow for electrical temperature compensation (see section 6).

Strain gages are resistors that generate heat due to the flow of an electrical current. The circuitry used
for resistance measurement increases the local temperature. Therefore, the heat dissipation of the local
gage installation (including prototype material and gage protection) must be checked and taken into
consideration in the measurement design.

Suggestions for maximum allowable power density of heat generated in the strain gage grids are given
in Table 4 for some prototype materials. The values depend on the ability that each material has to
dissipate the heat generated in the installation, helping to keep temperature as uniform as possible.
These power densities are calculated as the ratio between the heat generated by the strain gage and its
grid area. The maximum allowable circuit voltage measurement can be selected using the power
densities given in Table 4. The last column of this table gives allowable source voltages from
calculations that took into consideration equation (18), developed from the theory of the Wheatstone
bridge circuit presented in section 6. Voltages in Table 4 were calculated using the mid range value of
the power densities Pd, 120 Ω strain gages and grid areas equal to 10 mm2.

Table 4: Allowable maximum power densities Pd


Voltage from equation (18)
2 (kwatts/m2
Power density Pd = P/A (watts/in ) in volts, A = 10 mm2,
)
R=120 Ω
Thick sections of aluminum and
5 - 10 8 - 16 8
cooper alloy
Thick sections of steel parts 2-5 3-8 5
Thin sections of steel parts 1-2 1.5 - 3.0 3
Ceramic, glass, composite
0.2 - 0.5 0.3 - 0.8 1.6
materials
Polymers 0.02 - 0.05 0.03 - 0.08 0.5

V = 2 Pd . A.R (18)

5.3. Gage response in non-uniform strain fields

The resistance variation response given by the gage encompasses all the parallel sensing elements of its
grid. Therefore, the response is related to the average strain occurring in the prototype material
underneath the grid. This means that the effect of the strain gradient is averaged. Strain gages with
small grids should be used to give accurate strain information in sharp strain gradients, as illustrated in
Figure 7. Strips of small gages may be used to determine the strain gradient and the strain at the hot
spot in a welded structure.

Figure 7: Example of an inappropriately large strain gage placed in the stress concentration area to
measure the peak strain at the root of a U-notch. The average strain results from the total response of
the grid which includes elements positioned in the x and y directions.

5.4. Gage response in dynamic strain fields


Strain gage response to dynamic loading can unfold in three parts. The first is related to the influence
of the gage installation on the inertia and stiffness of the component. In general, relatively heavyweight
and large-size prototypes are not influenced by gage installations, but some care has to be taken with
the associated gage cabling. The second part has to do with the time required by the conducting grid
and associated backing carrier and adhesive to respond to the prototype strain wave. This time is on the
order of 100ns. The third part relates the grid length with the size of the wave pulse. Figure 8 shows the
distorted response given by a gage to a rectangular pulse that has a travel speed equal to c.

Figure 8: Distorted response of a strain gage to a rectangular wave

The bibliography cited at the end of the paper also shows the influence of different environments (such
as humidity, extremely low or high temperature, hydrostatic pressure, radiation) or type of loading
(creep, cyclic deformation or fatigue) on the strain gages and associated installations.

6. Strain gage circuits

Two types of signal conditioning circuits are commonly used to measure the variations in resistance of
strain gages: the potenciometer circuit (Figures 9 and 10) and the Wheatstone circuit (Figure 11), both
with configurations that use constant voltage or constant current energy supply sources.

These circuits measure the drop in voltage between specific points (Figures 9 and 11). The constant
current potenciometer circuit is commonly used with a direct resistance measurement circuit, as shown
in Figure 10.

6.1 Potenciometer circuit

The potentiometer circuit (Figure 9) can be activated by constant voltage or constant current sources. It
is recommended that the potenciometer circuit be coupled with highly accurate and very fast response
voltmeters since it does not have the hardware capability of zeroing the initial voltage readings
between the strain gage terminals. The voltage variation response of both potenciometer circuits is
given by equations (20) and (21), respectively, for the constant voltage and constant current circuits.
The constant voltage circuit presents a nonlinear term, given in equation (20).
Figure 9: The potenciometer circuit: constant voltage on left and constant current on right

Their response and connection with the strain gage response are respectively given by equations (20)
and (21).

1 ⎛ ∆R1 ∆R2 ⎞ 1 R2 (20)


∆VR = ⎜⎜ − ⎟(1 − η ) η = 1− r=
(1 + r ) 2
⎝ R1 R2 ⎟⎠ 1 ⎛ ∆R1 ∆R2 ⎞ R1
1+ ⎜⎜ +r ⎟
1 + r ⎝ R1 R2 ⎟⎠

∆R1 (21)
∆V R = I .R1 .
R1

Two advantages of the potenciometer constant current circuit are: the absence of the nonlinear term
and the absence of influence on the voltage output of the resistance of the circuit’s auxiliary cables.

Figure 10: The four-wire direct resistance measurement circuit

The so-called four-wire direct resistance measurement circuit is a constant current potenciometer
circuit that uses a highly accurate ohmmeter to measure the strain gage resistance. It is shown in Figure
10. This circuit is essentially the same as the one in the right of Figure 9. The ohmmeter needs a
resolution of at least 10-3 Ω in order to measure 1 to 10 µε accurately if low resistance (120 Ω) strain
gages are used. Application of this circuit is optimized if strain gages with higher initial resistance (e.g.
1,000 or 2,000 Ω) are used.

In any case, the accuracy and resolution required for the ohmmeters imply static measurements due to
the need for several measurements (around a thousand) to be automatically taken by the instrument
and later repeated by the measurement dedicated system (around 30 measurements). The mean
response of these measurements is used to avoid reading errors.

Constant current circuits are able to measure with an acquisition rate of one reading per second per
channel of measurement. This type of circuit can be very useful for measuring static phenomena that
span small variations during long duration measurements.

6.2. Constant voltage Wheatstone bridge circuit

The Wheatstone bridge is the most-used strain gage signal conditioning circuit. Due to its capability of
initial circuit balancing (initial output voltage zeroing), it has the great advantage of allowing the
measurement of small variations of resistance. The circuit is shown in Figure 11.

Figure 11: The Wheatstone bridge

Basically, the Wheatstone bridge consists of four resistances placed in the four arms of the circuit, a
constant voltage or current source, and a voltage reading and recording system. Equations (22) and (23)
present the development and the final response of output voltage E of the constant voltage bridge.

i = i1 + i 2
V
V AC = V = i1 .( R1 + R2 ) → i1 =
R1 + R2
V (22)
V AC = V = i 2 .( R3 + R4 ) → i2 =
R3 + R 4
V BD = E = R1 .i1 − R4 .i 2
V .R1 V .R 4
E= −
R1 + R2 R3 + R4

R1 .R3 − R2 .R4
E = V. (23)
(R1 + R2 ) × (R3 + R4 )
The numerator of (23) will be zero when the bridge is balanced. Using this capability, small variations
in resistance occurring in any of the arms will be included in the total variation of the output voltage
reading by the voltage measurement device. Equation (24) is arrived at under the assumption of initial
bridge balance, E = 0. Development of equation (24) leads to the basic bridge relation (25) and to the
nonlinear term η depicted in equation (26). This term is usually neglected. Generally, it is seen as being
very small when compared to the unity if the variations in resistances ∆R1, ∆R2, ∆R3 and ∆R4 are small
(as they are expected to be if elastic strains are being measured).

E=0
(R1 + ∆R1 )(. R3 + ∆R3 ) − (R2 + ∆R2 )(. R4 + ∆R4 ) (24)
∆E = V .
(R1 + ∆R1 + R2 + ∆R2 ) × (R3 + ∆R3 + R4 + ∆R4 )

⎛ ∆R ∆R2 ∆R3 ∆R4 ⎞ R2


⎟⎟.(1 − η )
r
∆E = V . .⎜⎜ 1 − + − r= (25)
(1 + r ) 2
⎝ 1R R 2 R3 R 4 ⎠ R1

1
η= (26)
1+ r
1+
∆R 1 ∆R4 ⎛ ∆R ∆R3 ⎞
+ + r.⎜⎜ 2 + ⎟
R1 R4 ⎝ R2 R3 ⎟⎠

For example, using r = 1 and variations of ∆R of around 1% (usually very large in strain gage
measurements) the nonlinear term η of equation (26) becomes approximately 0.02 <<1.00. Therefore,
equation (25) is written without the nonlinear term as

r ⎛ ∆R ∆R2 ∆R3 ∆R4 ⎞


∆E = V . .⎜⎜ 1 − + − ⎟ (27)
(1 + r ) 2
⎝ R1 R2 R3 R4 ⎟⎠

6.3. Constant current Wheatstone bridge circuit

When compared with the constant voltage power supply circuit, the Wheatstone bridge circuit can also
employ a constant current power supply with the advantage of reducing the nonlinear effect.

The output voltage from an initially balanced constant current bridge is given by equation (28). It
shows that nonlinearity is caused by the sum of the resistance variations, located in the denominator of
the expression, and by the products of the ∆R terms, located inside the second bracket in the
numerator.
R1 .R3 ⎡⎛ ∆R ∆R2 ∆R3 ∆R4 ⎞ ⎛ ∆R 1 ∆R3 ∆R2 ∆R4 ⎞⎤
∆E = I . .⎢⎜⎜ 1 − + − ⎟⎟ + ⎜⎜ − ⎟⎟⎥ (28)
∑ Ri + ∑ ∆Ri ⎣⎝ 1R R 2 R3 R 4 ⎠ ⎝ 1
R R3 R2 R4 ⎠⎦

6.4. Constant voltage Wheatstone bridge circuits – specific cases

The relationship between the output voltage and the sensed strain is given for three installation cases,
illustrated in Figures 12 to 14.

The first case (Figure 12) represents the majority of installations used in field measurements where
only one arm of the bridge is used. It is known as the “quarter bridge” configuration. This case is used
here to show that an amplification in output is required to couple the bridge with ordinary accuracy and
resolution voltmeters. For example, a relatively average to high strain value of around 1,000 µε will
produce an output of only 1mv if the excitation voltage is equal to 2v, gage factor K is 2 and ratio r is
1.

Figure 12: The Wheatstone bridge - quarter bridge configuration

The second case uses two arms of the bridge and is commonly known as the “half bridge”
configuration (Figure 13). The gages are mounted in arms R1 and R2 or R4. Generally, it is used when
one wants to increase the output signal and minimize the temperature output. The employment of such
configurations requires knowledge of the strain distribution in order to place the gages in the most
appropriate positions.

Two examples may be given for this case. One is the use of two gages on opposite surfaces of the cross
section of a beam under flexure loading. The output of the bridge will be proportional to twice the
strain on those surfaces if the gages are placed in arms R1 and R2 or R4. The second example is the
installation of two gages in the longitudinal and transverse directions of a prismatic bar under axial
loading, where the output will be given by the longitudinal strain increased by factor (1+µ), where µ is
the Poisson coefficient. In both cases, it can be seen, using equation (27), that any undesirable spurious
temperature output from the gages will be cancelled out due to the opposite signal in the adjacent arms
of the bridge.

Figure 13: The Wheatstone bridge - half bridge configuration

The third case employs four gages placed in the arms of the bridge (Figure 14). This configuration is
called the “full bridge”. It is used to cancel out spurious temperature outputs from the gages, while at
the same time amplifying the output signal by multiplying the existing strains by four or by 2(1+ µ),
respectively, in instances such as a beam under bending or a bar under axial loading. This case is
encountered in most load cell circuits. It is also used in field measurements when certain types of
loadings are to be measured in a structure but other loads are not.

Figure 14: The Wheatstone bridge - full bridge configuration


6.5. Shunt Calibration

The shunt calibration is used to calibrate bridge output. It employs the placement of a parallel or shunt
resistance along one of the arms of the bridge. A switch is also set in place to connect or disconnect the
parallel resistance. When the switch is off, the arm resistance is R. When the switch is on, the arm
resistance will change to a lower resistance. The parallel resistance can be calculated to simulate the
output signal equivalent to a given strain; for example, -1000µε. If the gage factor is assumed to be
K=2.0, the shunt calibration required for an output of -1000µε will be 59880Ω or 174650Ω,
respectively, for strain gage resistances of 120 Ω and 350Ω. The development for the cases of 120Ω
and 350Ω strain gages that have resistances Rs placed in parallel with the arm is illustrated in Figure
15. Equation (29) gives the values of Rs as a function of gage factor K and of desired calibration strain
ε, which must be entered as a negative value since ∆R caused by the parallel resistance is negative.

Figure 15: Example of application of parallel resistance Rs to strain gage instrumented arm of
Wheatstone bridge to give electrical output equivalent to strain equal to -1000µε

⎛ 1 ⎞
Rs = − Rg .⎜1 + ⎟ (29)
⎝ K .ε ⎠

6.6 Quarter bridge: three-wire connection of a strain gage

A common practice used in quarter bridge installations is the use of the three-wire connection instead
of the two-wire connection. Two-wire connections are sensitive to temperature output from the cabling
and they also decrease the sensitivity output if the cables are very long. Figure 16 compares both types
of connections with long cables. It can be seen that the resistance of the R1 arm of the bridge is
composed of the strain gage resistance and the resistance of both connecting wires. The resistance
variation caused by the temperature variation occurring in both cables will aid the strain output sensed
by the gage and will increase the denominator of ∆R/R. The three-wire arrangement relocates the
resistances of the cables in the adjacent arms of the bridge. One cable resistance is placed in arm R1,
the other in arm R4, and the third cable resistance is placed in the voltage source arm of the bridge,
causing a minimum interference in the bridge output. The temperature output generated in the wires
placed in arms R1 and R4 will be canceled out in the measurement of ∆E, as can be seen by application
of equation (27).

Figure 16: Quarter bridges with two and three-wire connections

6.7 Null balance bridges

The term “balancing” means the adjustment of the resistances of the arms of the bridge so that voltage
output ∆E becomes zero. In most cases, a precision resistor with total resistance R5 + R6 is placed in
parallel with arms R2 and R3, as depicted in Figure 17. Resistances R5 and R6 are varied (∆R5 = -
∆R6) to achieve identity of products R1 and R3eq, and R4 and R2eq.

In the null balancing bridge, the variation of the strain gage resistance in arm 1 is determined by the
measurement of the variation of the precision resistor placed in arm 5-6. The measurement is
accomplished by counting the number of turns given in the resistor, which nulls the electrical current
transiting galvanometer G, placed between points B and D of the bridge.

Presently, in many applications balancing is achieved by software. The measuring circuits use highly
accurate voltage meters that measure the initial imbalance of the circuit and subtract it from the voltage
readings whenever a strain measurement is required.
Figure 17: Balancing the Wheatstone bridge with a precision variable resistor

7. Strain gage measurement systems

There is a reasonable number of strain gage conditioner makers world wide. It is recommended that a
thorough search be made in order to select the most adequate conditioner for a given application.
Figure 18 and Table 5 illustrate and list the most important features and functions a measurement
system should have.

Table 5: Features and functions of a strain gage measurement system

Basic features Static measurement capability


Use of at least one strain gage in a quarter bridge configuration
Constant voltage source, generally equal to 2V. Other maximum voltages are possible; for example, 2.5V, 5V, 10V,
or continuous variable voltage source from 0 to 10 or 15V
Battery or electrical DC or AC energy feed
Built-in amplifier with gain equal to 100, 200 or 500
Output voltage terminal to connect to output voltage reading device, or built-in voltage reading device and recorder
Bridge completion resistors - possibility of selecting circuit to work with quarter, half or full bridge
Adjustment of gage factor K circuit
Calibration circuit (shunt calibration)
Initial balancing circuit
Advanced Channel multiplexing - manual or automatic devices.
features Static or dynamic measurements with acquisition rates of 2kHz or more per channel
Data recording for 1, 2, 8, …, 1028 channels or more
Noise filters
Software-based recorder and data analyzers with functions ranging from rosette stress analysis to spectral analysis
and FFT
Figure 18: Main features of a strain gage signal conditioner with multiplexing capability

7. Examples of strain gage applications

Basically, strain gages are used as sensor elements in a large variety of transducers in field or
laboratory applications for direct strain measurements.

Strain gages are applied as sensor elements in transducers to measure force, such as in load cells used
in weight scales and other measurement devices such as elongation clip gages, inclinometers and
accelerometers to measure, respectively, displacement, inclination angles and acceleration. In all these
applications, one or more full Wheatstone bridges are used. The gages are located and bonded in
appropriate geometric configurations of spring elements in order to amplify as much as possible the
signal to be measured and to avoid other possible spurious signals generated by unpredictable or
undesirable loads. Examples of spring elements for load cells are given in Figures 19a to f.
(b)

(a)

(d)
(c)

(e)

(f)

Figure 20: Illustrations of spring elements for load cells. Geometric arrangements depend on type of
load effort related to load to be measured: a) normal load, b) basic geometry for bending load; c)
advanced geometry for bending load, d) advanced binocular geometry for bending load; e) cantilever
geometry for shear load; f) short pin geometry for shear load

In the case of precision transducers, there are certain practices for avoiding small uncertainties caused
by the temperature response of the spring material and strain gages. Specifically, calculated or
experimentally dimensioned Constantan, Cupper and Balco (a nickel-iron alloy) resistors are placed to
adjust the sensitivity of the installation, to keep response at zero with the variation of temperature, and
to give a null response under zero load. Figure 20 presents the general arrangement of the complete full
bridge for load cells.
Figure 20: General arrangement of complete full bridge for load cells, including resistors to adjust for
sensitivity, bridge balance and influence of temperature.

Laboratory and field use for direct measurement of strains with the objective of understanding the
mechanical behavior of large or small structures represents a very important application of strain gages.

Figure 21 shows an example of the instrumentation of a steel pipe with a machined defect simulating
metal thickness loss caused by corrosion. The steel pipe had closing caps and was hydrostatically
tested to rupture. The strain gage was used to measure the circumferential strain at the instant of
rupture. In this case, a special uniaxial strain gage capable of reaching and measuring large strains was
used.

Figure 21: Bonding a strain gage rosette at the center region of a pipe with external machined defect to
determine the maximum circumferential strain in a laboratory burst pressure test.

Figure 22 presents pictures of two sections of a stacker machine that was tested to confirm assumptions
made at the design stage in order to make an increase in its load capability possible.
Location of structural
members that were
monitored

(a)

(c) rosette at
stress
concentration

(b) uniaxial gage at a


member under nominal
loading conditions

(d) signal conditioner and


data acquisition system

Figure 22: Instrumentation of a large iron-ore pile stacker machine

8. Conclusions

This article has presented the principal concepts regarding strain gages. These gages are used to
determine strains in localized areas of structural components in the laboratory or field, or as sensors in
transducers such as resistive accelerometers or load cells. They are accurate “point” measuring
elements and can be used in cases of static or dynamic loading. The measurement systems employ
conventional and on-shelf circuitry elements, and they present the important advantage of leading to
electrical and treatable output.

Glossary

Measurement system: Consists of a device or set of devices associated with a protocol, and are used
to make a measurement.
Sensor: The basic sensing element of a transducer. Transforms an input into an output.
Constantan: 55% Cu, 45% Ni alloy used in the form of a thin wire or thin foils as the sensor element
in strain gages.
Advance: Same as Constantan
Karma: 74% Ni, 20% Cr, 3% Al, 3% Fe alloy used as the sensor element in strain gages.
Cyanoacrylate: An adhesive compound that polymerizes quickly when spread as a very thin film.
Transducer: The measurement system element that is affected by the quantity to be measured. Helps
the sensor to receive input and to transmit output.
Potenciometer circuit: Electrical circuit that results from connecting the strain gage to a voltage meter
and a power source. Used to determine the variation in the resistance of the strain gage. Can be
powered by a constant voltage or constant current source.
Wheatstone circuit: Electrical circuit composed of four resistors, voltage meter and power source.
Used to determine variation in the resistance of the one, two or four strain gages that are placed in its
four arms. Can be powered by a constant voltage or constant current source.
Shunt calibration: Placement of a known resistor in parallel with a strain gage or arm of the
measuring system circuit in order to cause a known resistance variation which will be used for
calibration purposes.
Null balance bridge: Wheatstone bridge circuit with adjustable resistors placed in parallel with two of
its four arms in order to make its voltage output equal to zero.

Bibliography

1. “Experimental Stress Analysis”, J.W. Dally and W.F. Riley, 4th ed., College House Enterprises,
LLC, 5713 Glen Cove Drive, Knoxville Tennessee, USA, 2006, jdally@collegehousebooks.com
[a comprehensive approach to Experimental Mechanics. The topic strain gages is discussed with
depth]
2. “Instrumentation for Engineering Measurements”, J.W. Dally, W.F. Riley and K.G. MacConnell,
John Wiley & Sons, 2nd ed., 1993 [A comprehensive approach to measurement in mechanics. The
topic circuits for strain gages is discussed in depth].
3. Strain Gage User’s Handbook, edited by R.L. Hannah and S.E. Reed, Society for Experimental
Mechanics, Elsevier Applied Science [The whole book thoroughly discusses all topics regarding
strain gages].
4. Handbook of Experimental Mechanics, edited by A.S. Kobayashi, Society for Experimental
Mechanics, Prentice-Hall, 1987 [A comprehensive approach to Experimental Mechanics. The
topic strain gages is discussed in depth in one of its chapters].
5. Springer Handbook of Experimental Solid Mechanics, 1st Edition, edited by William N. Sharpe,
Society for Experimental Mechanics, Springer, 2008 [A comprehensive approach to Experimental
Mechanics. The topic strain gages is discussed in depth].
6. Manual on Experimental Stress Analysis, edited by, J.F. Doyle and J.W. Phillips, Society for
Experimental Mechanics, 7 School Street, Bethel, CT, USA, 1989 [An introductory approach to
Experimental Mechanics. The topic strain gages is discussed in one of its chapters].
7. “The Strain Gage Primer”, C.C. Perry and H.R. Lissner, 2nd edition, McGraw-Hill, 1962 [A
comprehensive approach to strain gages is discussed in depth in the book’s several chapters].

Author’s biographical sketch

José L.F. Freire: B.S. (1972) and M.Sc. (1975) degrees in mechanical engineering from the Catholic
University of Rio de Janeiro (PUC-Rio); Ph.D. (1979) in Engineering Mechanics from Iowa State
University of Science and Technology (ISU); associate professor of Mechanical Engineering at PUC-
Rio and chairman of the Structural Integrity Laboratory; member of the Society for Experimental
Mechanics and is its past president; major areas of research: Experimental Stress Analysis, Pipeline
Engineering and Structural Integrity of Equipment and Structures.

You might also like