You are on page 1of 30

Strain Gage Types

The basic principle of the resistance strain gage is implemented in


several different ways (with the first two being largely obsolete):

1. Unbonded metal-wire gage

2. Bonded metal-wire gage

3. Bonded metal-foil gage

4. Vacuum-deposited thin-metal-film gage

5. Sputter-deposited thin-metal-film gage

6. Bonded semiconductor gage

7. Diffused semiconductor gauge


The unbonded metal-wire gage employs a set of preloaded resistance wires
connected in a Wheatstone bridge. as in Fig. 4.10.

At initial preload. the strains and resistances of the four wires are nominally
equal, which gives a balanced bridge and eo = 0

Application of a small (full scale = 0.04 mm) input motion increases tension in
two wires and decreases it in two others (wires neither go slack), causing
corresponding resistance changes, bridge unbalance, and output voltage
proportional to input motion.

The wires may be made of various copper-nickel. chrome-nickel,or nickel-iron


alloys, are about 0.03 mm in diameter, can sustain a maximum force of only
about 0.001N, and have a gage factor of 2 to 4,

Electric resistance of each bridge arm is 120Ω to 1,000Ω.

Maximum excitation voltage is 5 to 10 V, and full-scale output typically is 20mV


to 50 mV.
The bonded metal-wire gage (today largely superseded by the bonded metal-foil
construction) has been applied to both stress analysis and transducers.

A grid of fine wire is cemented to the specimen surface, where strain is to be


measured.

Embedded in a matrix of cement, the wires cannot buckle and thus faithfully
follow both the tension and the compression strains of the specimen.

Since materials and wire sizes are similar to those of the unbonded gage, gage
factors and resistances are comparable.
Bonded metal-foil gages using identical or similar materials to wire gages are used
today for most general-purpose stress-analysis tasks and many transducers.

The sensing elements are formed from sheets less than 5 μm thick by photo-etching
processes. which allows great flexibility with regard to shape.

In Fig. 4.11, for example, the three linear grid gages are designed with “fat" end
turns.

This local increase in area reduces the transverse sensitivity, a spurious input since
the gage is intended to measure the strain component along the length of the grid
elements.

In a wire gage, these end turns would necessarily have the same cross-section as
the longitudinal elements, which increases the spurious transverse sensitivity.

The manufacturing process also easily provides convenient soldering


tabs (integral to the sensing grid) on all four gages of Fig. 4.11
Bonded metal-foil gages, are the type most likely to be employed by the individual
engineer, both in stress analysis and for construction of home made transducers.

These gages come mounted on a flexible insulating carrier film (polyimide, glass-
reinforced phenolic, etc.) about 25 μm thick: thus the 5 μm thick metal grid will be
slightly more than 25 μm in above the specimen surface, as a result of adhesive
film thickness.

Gages intended for stress analysis applications use a flexible carrier film, allowing
them to easily conform to curved surfaces and protecting them from damage
during handling and installation.

Gages marketed for transducer applications may use a more brittle film.

This requires more care in installation, but reduces hysteresis effects from the 0.1
percent level to about 0.05 percent, important for high-accuracy transducers.
Some transducer-grade gages use a platinum-tungsten alloy tor the grid, giving a
gage factor (approx. 4.5) about twice that of "ordinary" gages.

This allows the same output voltage for lower strain levels, increasing the fatigue
life, linearity and overload safety.

This displacement is significant, only when we are measuring bending strain in


very thin specimens, since then the gage could be feeling a strain considerably
different from that of the specimen.

ln stress analysis, the goal is measurement of stress at a geometric point: this is


impossible in a strain -gage because the grid covers a finite area and thus the
gage reads an average stress over this area.

If the strain gradient is linear, then this average value also can be associated
with the midpoint of the gage length; if not, then the point at which the gage is
reading applies is somewhat uncertain.

However, this uncertainty diminishes with gage size, which makes small gages
desirable where strain gradients are sharp (stress concentrations, etc.).
Strip (also called chain) gages consist of several gages, all mounted on the
same backing strip, adjacent to each other.

These are used to study local strain distributions- perhaps near a stress
concentration.

A typical strip might have 10 gages, each with 0.5 mm gage length, with a total
length of 10 mm.

Minimum practical gage size is constrained by manufacturing limitations and


handling attachment problems: the smallest gages are about 0.18 mm long.

Gages can be applied to curved surface, the minimum safe bend radius can be
as small as 1.5 mm in some gages.

Typical gage resistances are 120, 350, and 1000Ω. with the allowable gage
current determined by heat-transfer conditions. but typically, 5 to 40 mA; gage
factors are 2 to 4.
Resistance of an individual gage is easily measured, but measurement of the gage
factor requires cementing the gage to a specimen for which strain can be accurately
calculated from theory,

Since gages cannot be removed and reused, the gage-factor number supplied with
a purchased gage has not been measured for that individual gage, rather is an
average value obtained by sample-testing the production of that type of gage.

Thus we rely on the statistical quality control of the manufacturer in maintaining the
accuracy of the gage factor.

A ±1 percent accuracy is typical, and this is a fundamental limit on accuracy in


stress-analysis applications.

Note that this does not limit accuracy of strain-gage transducers, since such
transducers are calibrated "end to end" (pressure in to voltage out, say, in a
pressure transducer) and even one need not know the gage factor.

The maximum measurable strain varies from 0.5 to1 percent; however,
Special "postyield"gage deveces allow measurement upto 0.1 mm/mm.
Fatrgue life of gages varies with conditions:
However, 10 million cycles at ±1.500 microstrain 310 MPa stress in steel, a
common full scale design value for foil-gage transducers) is typical.

In addition to single-element cases. gage combinations called rossettes,(Fig.4.12)


are available in many configurations for specific stress-analysis or transducer
applications.

While individual gages conceivably could be cemented down in the same


patterns, precise relative orientation of the several gages is critical in most of
these applications, and this is much more easily obtained in rosette manufacture
than by the user with single gages.

One rosette commonly used in stress analysis solves the problem of a surface
stress analysis whose magnitude and direction are totally unknown.

Theory shows that measurements with a three-gage rossette allows calculation of


all the desired information.

Since such measurements are intended to define stress at a point, ideally the
three gages should be superimposed on that point.
This "sandwich" construction (called stacked rosette) is feasible; however, it
places the top gage farther from the specimen surface and increases its self-
heating, since it is better insulated from the underlying metal specimen which
acts as a heat sink.

If these disadvantages outweigh the advantage of point measurement, we can


use the planar rosette design, which usually is available, (see Fig. 4.12)
Temperature is an important interfering input for strain gages, since resistance
changes with both strain and temperature.

Since strain-induced resistance changes are quite small, the temperature effect
can assume major proportions.

Another aspect of temperature sensitivity is found in the possible differential


thermal expansion of the gage and the underlying material.

This can cause a strain and resistance change in the gage, even though the
material is not subjected to an external load.

These temperature effects can be compensated in various ways.

In Fig. 4.13, a "dummy" gage (identical to the gage) is cemented to a piece of the
same material as is the active gage and placed so as to assume the same
temperature.

The dummy and active gages are placed in adjacent legs of a Wheatstone bridge;
thus resistance changes due to the temperature coefficient of resistance and
differential thermal expansion will have no effect on the bridge output voltage,
whereas resistance changes due to an applied load will unbalance the bridge
in the usual way.

Another approach to this problem involves special, inherently temperature-


compensated gages.

These gages are designed to be used on a specific material and have


expansion and resistance properties such that the two effects very nearly
cancel and no dummy gage is required.
Temperature also can act as a modifying input in that it may change the gage
factor.

With metallic gages, this effect usually is quite small, except at extremely low
or high temperatures.

Semiconductor gages are more seriously affected in this way; however,


compensation is possible.

Although the above temperature problems must be considered carefully in


each application, strain gages have been successfully employed from liquid-
helium temperature -269oC to the order of 1,1000C.

However, these extreme (especially the high temperature) applications


require special techniques and yield results of lower accuracy, than are
obtained in routine room-temperature situations.
Used directly, the bonded strain gage is useful for measuring only very
small displacements (strains).

However, larger displacements may be measured by bonding the gage to


a flexible element, such as a thin cantilever beam, and applying the
unknown displacement to the end of the beam. as in Fig. 4.14a.

For such an application, the gage factor need not be accurately known,
since the overall system can be calibrated by applying known
displacements to the end of the beam and measuring the resulting bridge
output voltage.

The configuration shown is temperature-compensated without the need


for dummy gages and has four times the sensitivity of a single gage
because of judicious application of bridge-circuit properties.

Such transducers may be accurate to 0.1 percent of full scale.


Fig.4.14(a) Strain gage displacement transducers and extensometers.
Weldable Strain Gages:

To meet the needs of severe environments difficult installation conditions, and/or


high temperatures as typified by pipelines, offshore drilling platforms, tunnels,
dams, nuclear containment vessels, steam lines, gas turbines, etc., strain gages
enclosed permanently in protective metal housings, which are attached by spot
welding to the specimen have been developed.

These gages incorporate a number of unique features (see Fig.4.16).

The strain filament and heavier(178 μm diameter) leadout wire are of unitized
construction with a carefully controlled taper; this minimizes joint fatigue
problems and erratic electric connections.

Active and dummy filaments are in close proximity; the dummy gage
experiences no strain since the helix angle of its winding matches Poisson's
ratio.

Both filaments are encased in a strain tube made by welding a tubular shell to a
flat flange.
The filament is mechanically coupled to the strain tube, but electrically isolated
from it by compacted magnesium oxide powder, by using a high-speed
centrifuge and swaging operation.

When the flange is spot-welded to the specimen. the specimen strain is


faithfully transmitted to the strain tube and in turn by the powder to the gage
filament for both tensile and compressive stresses.
Evaporation and Sputtered Deposited Thin-metal-film Gages:
Evaporation-deposited, thin-metal-film gages are used mostly for transducers, as
are the sputter-deposited variety.

Both processes begin with a suitable elastic metal element to transduce the
measured quantity to local strain, just as if bonded foil gages were to be used.

To use a pressure transducer as an example, this would be a thin, circular, metal


diaphragm.

Both the evaporation and the sputtering processes form all the gage elements
directly on the strain surface: they are not separately attached, as with bonded
gages.

In the evaporation process, the diaphragm is placed in a vacuum chamber with some
insulating material.

Heat is applied until the insulating material vaporizes and then condenses, forming a
thin dielectric film on the diaphragm.

Then suitably shaped templates are placed over the diaphragm, and the
evaporation/condensation process is repeated with the metallic gage material,
forming the desired gage pattern on top of the insulating substrate.
In the sputtering process, a thin dielectric layer is again deposited in vacuum over
the entire diaphragm surface: however, the detailed mechanism of deposition is
different from the evaporation method.

Then a complete layer (no templates) of metallic gage material is sputtered on top
of the dielectric substrate.

Now the diaphragms are removed from the vacuum chamber, and microimaging
techniques using photosensitive masking materials are used to define the gage
patterns.

Returning the diaphragms to the vacuum chamber, we now use sputter-etching to


remove all the unmasked metal layer, leaving behind the completed gage pattern.

Resistances and gage factors of film gages usually are similar to those of bonded
foil.

Since no organic cements, as used with bonded foil gages are employed, thin-film
gages exhibit improved time and temperature stability.

Recent developments in the sputtering technique have provided useful strain,


temperature and corrosion sensors for difficult jet-engine turbine blade
measurements.
Bonded Semiconductor Gages: Bonded semiconductor gages are used mainly in
transducers; however, they may find occasional application in stress analysis if the
available strain is very small.

They are made by slicing small sections from specially processed silicon crystals and
are available in both N- and P-type.

The P-type gages increase resistance with applied tensile strain while the N-type
decrease resistance.‘
Their main feature is a very high gage factor--as much as 150.

Most of this gage factor must come from piezoresistance effects, whereas
dimensional change explains most of the gage factor of metallic gages.

Transducers based on semiconductor gages are often called piezoresistive


transducers.

Unfortunately, the high gage factor is accompanied by high temperature sensitivity,


nonlinearity, and mounting difficulties.

Solutions for these problems exist, and they can be accepted in the transducer
manufacturing environment, but they are difficult to live with in routine stress analysis.
In diffused semiconductor gages (used exclusively in transducers), the diffusion
process employed in integrated-circuit manufacture is utilized.

In a pressure transducer, for example, the diaphragm would be silicon rather than
metal, and the strain-gage effect would be realized by depositing impurities in the
diaphragm to form intrinsic strain gages at the desired locations.

This type of construction allows lower manufacturing costs since a large number of
diaphragms can be made on a single silicon wafer.

You might also like