Você está na página 1de 22

Hardness

http://en.wikipedia.org/wiki/Hardness

Hardness
From Wikipedia, the free encyclopedia
This article is about mechanical properties of materials. For other uses, see hard (disambiguation).
"Durezza" redirects here. For the French wine grape, see Durezza (grape).
Hardness is a measure of how resistant solid matter is to various kinds of permanent shape change when a
compressive force is applied. Some materials, such as metal, are harder than others. Macroscopic hardness is
generally characterized by strong intermolecular bonds, but the behavior of solid materials under force is
complex; therefore, there are different measurements of hardness: scratch hardness, indentation hardness,
and rebound hardness.
Hardness is dependent on ductility, elastic stiffness, plasticity, strain, strength, toughness, viscoelasticity,
and viscosity.
Common examples of hard matter are ceramics, concrete, certain metals, and superhard materials, which
can be contrasted with soft matter.
Contents
[hide]

1 Measuring hardness
o

1.1 Scratch hardness

1.2 Indentation hardness

1.3 Rebound hardness

2 Hardening

3 Physics
o

3.1 Mechanisms and theory


4 See also

4.1 Other strengthening mechanisms

5 References

6 Further reading

Hardness

7 External links
Measuring hardness[edit]

A Vickers hardness tester


There are three main types of hardness measurements: scratch, indentation, and rebound. Within each of
these classes of measurement there are individual measurement scales. For practical reasons conversion
tables are used to convert between one scale and another.
Scratch hardness[edit]
Main article: Scratch hardness
Scratch hardness is the measure of how resistant a sample is to fracture or permanent plastic
deformation due to friction from a sharp object.[1] The principle is that an object made of a harder material
will scratch an object made of a softer material. When testing coatings, scratch hardness refers to the force
necessary to cut through the film to the substrate. The most common test is Mohs scale, which is used
in mineralogy. One tool to make this measurement is the sclerometer.
Another tool used to make these tests is the pocket hardness tester. This tool consists of a scale arm with
graduated markings attached to a four wheeled carriage. A scratch tool with a sharp rim is mounted at a
predetermined angle to the testing surface. In order to use it a weight of known mass is added to the scale
arm at one of the graduated markings, the tool is then drawn across the test surface. The use of the weight
and markings allows a known pressure to be applied without the need for complicated machinery. [2]
Indentation hardness[edit]
Main article: Indentation hardness
Indentation hardness measures the resistance of a sample to material deformation due to a constant
compression load from a sharp object; they are primarily used in engineering and metallurgy fields. The tests
work on the basic premise of measuring the critical dimensions of an indentation left by a specifically
dimensioned and loaded indenter.

Hardness
Common indentation hardness scales are Rockwell, Vickers, Shore, and Brinell.
Rebound hardness[edit]
Rebound hardness, also known as dynamic hardness, measures the height of the "bounce" of a diamondtipped hammer dropped from a fixed height onto a material. This type of hardness is related to elasticity. The
device used to take this measurement is known as a scleroscope.[3]
Two scales that measures rebound hardness are the Leeb rebound hardness test and Bennett hardness scale.
Hardening[edit]
Main article: Hardening (metallurgy)
There are five hardening processes: Hall-Petch strengthening, work hardening, solid solution
strengthening, precipitation hardening, and martensitic transformation.
Physics[edit]

Diagram of a stress-strain curve, showing the relationship betweenstress (force applied per unit area)
andstrain or deformation of a ductile metal.
In solid mechanics, solids generally have three responses to force, depending on the amount of force and the
type of material:

They exhibit elasticitythe ability to temporarily change shape, but return to the original shape when
the pressure is removed. "Hardness" in the elastic rangea small temporary change in shape for a given
forceis known as stiffness in the case of a given object, or a high elastic modulusin the case of a
material.

They exhibit plasticitythe ability to permanently change shape in response to the force, but remain
in one piece. The yield strength is the point at which elastic deformation gives way to plastic deformation.
Deformation in the plastic range is non-linear, and is described by the stress-strain curve. This response
produces the observed properties of scratch and indentation hardness, as described and measured in
materials science. Some materials exhibit both elasticity and viscosity when undergoing plastic
deformation; this is called viscoelasticity.

They fracturesplit into two or more pieces.

Strength is a measure of the extent of a material's elastic range, or elastic and plastic ranges together. This is
quantified as compressive strength,shear strength, tensile strength depending on the direction of the forces

Hardness
involved. Ultimate strength is an engineering measure of the maximum load a part of a specific material and
geometry can withstand.
Brittleness, in technical usage, is the tendency of a material to fracture with very little or no detectable
plastic deformation beforehand. Thus in technical terms, a material can be both brittle and strong. In
everyday usage "brittleness" usually refers to the tendency to fracture under a small amount of force, which
exhibits both brittleness and a lack of strength (in the technical sense). For perfectly brittle materials, yield
strength and ultimate strength are the same, because they do not experience detectable plastic deformation.
The opposite of brittleness is ductility.
The toughness of a material is the maximum amount of energy it can absorb before fracturing, which is
different from the amount of force that can be applied. Toughness tends to be small for brittle materials,
because elastic and plastic deformations allow materials to absorb large amounts of energy.
Hardness increases with decreasing particle size. This is known as the Hall-Petch relationship. However,
below a critical grain-size, hardness decreases with decreasing grain size. This is known as the inverse HallPetch effect.
Hardness of a material to deformation is dependent on its microdurability or small-scale shear modulus in any
direction, not to any rigidity or stiffness properties such as its bulk modulus orYoung's modulus. Stiffness is
often confused for hardness.[4][5] Some materials are stiffer than diamond (e.g. osmium) but are not harder,
and are prone to spalling and flaking in squamose or acicular habits.
Mechanisms and theory[edit]

A representation of the crystal lattice showing the planes of atoms.


The key to understanding the mechanism behind hardness is understanding the metallic microstructure, or
the structure and arrangement of the atoms at the atomic level. In fact, most important metallic properties
critical to the manufacturing of todays goods are determined by the microstructure of a material. [6] At the
atomic level, the atoms in a metal are arranged in an orderly three-dimensional array called a crystal lattice.
In reality, however, a given specimen of a metal likely never contains a consistent single crystal lattice. A
given sample of metal will contain many grains, with each grain having a fairly consistent array pattern. At an
even smaller scale, each grain contains irregularities.
There are two types of irregularities at the grain level of the microstructure that are responsible for the
hardness of the material. These irregularities are point defects and line defects. A point defect is an
irregularity located at a single lattice site inside of the overall three-dimensional lattice of the grain. There are
three main point defects. If there is an atom missing from the array, a vacancy defect is formed. If there is a
different type of atom at the lattice site that should normally be occupied by a metal atom, a substitutional
defect is formed. If there exists an atom in a site where there should normally not be, an interstitial defect is
formed. This is possible because space exists between atoms in a crystal lattice. While point defects are

Hardness
irregularities at a single site in the crystal lattice, line defects are irregularities on a plane of
atoms. Dislocations are a type of line defect involving the misalignment of these planes. In the case of an
edge dislocation, a half plane of atoms is wedged between two planes of atoms. In the case of a screw
dislocation two planes of atoms are offset with a helical array running between them. [7]
In glasses, hardness seems to depend linearly on the number of topological constraints acting between the
atoms of the network.[8] Hence, the rigidity theory has allowed predicting hardness values with respect to
composition.

Planes of atoms split by an edge dislocation.


Dislocations provide a mechanism for planes of atoms to slip and thus a method for plastic or permanent
deformation.[6] Planes of atoms can flip from one side of the dislocation to the other effectively allowing the
dislocation to traverse through the material and the material to deform permanently. The movement allowed
by these dislocations causes a decrease in the material's hardness.
The way to inhibit the movement of planes of atoms, and thus make them harder, involves the interaction of
dislocations with each other and interstitial atoms. When a dislocation intersects with a second dislocation, it
can no longer traverse through the crystal lattice. The intersection of dislocations creates an anchor point
and does not allow the planes of atoms to continue to slip over one another [9] A dislocation can also be
anchored by the interaction with interstitial atoms. If a dislocation comes in contact with two or more
interstitial atoms, the slip of the planes will again be disrupted. The interstitial atoms create anchor points, or
pinning points, in the same manner as intersecting dislocations.
By varying the presence of interstitial atoms and the density of dislocations, a particular metal's hardness
can be controlled. Although seemingly counter-intuitive, as the density of dislocations increases, there are
more intersections created and consequently more anchor points. Similarly, as more interstitial atoms are
added, more pinning points that impede the movements of dislocations are formed. As a result, the more
anchor points added, the harder the material will become.
See also[edit]

Hardness comparison

Vickers hardness test

Brinell hardness test

Hardness of ceramics

Toughness

Hardness

Schmidt hammer

Roll hardness tester

Tablet hardness testing

Persoz pendulum

Other strengthening mechanisms[edit]

Grain boundary strengthening

Precipitation hardening

Solid solution strengthening

Work hardening

References[edit]
1.

Jump up^ Wredenberg, Fredrik; PL Larsson (2009). "Scratch testing of metals and polymers:
Experiments and numerics". Wear 266 (12): 76. doi:10.1016/j.wear.2008.05.014.

2.

Jump up^ Hoffman Scratch Hardness Tester. byk.com

3.

Jump up^ Allen, Robert (2006-12-10). "A guide to rebound hardness and scleroscope test".
Retrieved 2008-09-08.

4.

Jump up^ Jeandron, Michelle (2005-08-25). "Diamonds are not forever". Physics World.

5.

Jump up^ San-Miguel, A.; Blase, X.; Mlinon, P.; Perez, A.; Iti, J.; Polian, A.; Reny, E. et al.
(1999-05-19). "High Pressure Behavior of Silicon Clathrates: A New Class of Low Compressibility
Materials".Physical Review 83 (25):
5290. Bibcode:1999PhRvL..83.5290S. doi:10.1103/PhysRevLett.83.5290.

6.

^ Jump up to:a

Haasen, P. (1978). Physical metallurgy. Cambridge [Eng.] ; New York:

Cambridge University Press.


7.

Jump up^ Samuel, J. (2009). Introduction to materials science course manual. Madison,
Wisconsin: University of Wisconsin-Madison.

8.

Jump up^ Smedskjaer, Morten M.; John C. Mauro; Yuanzheng Yue (2010). "Prediction of Glass
Hardness Using Temperature-Dependent Constraint Theory". Phys. Rev. Lett. 105 (11):
2010.doi:10.1103/PhysRevLett.105.115503.

Hardness
9.

Jump up^ Leslie, W. C. (1981). The physical metallurgy of steels. Washington: Hempisphere
Pub. Corp., New York: McGraw-Hill, ISBN 0070377804.

Further reading[edit]

Chinn, R. L. (2009). "Hardness, bearings, and the Rockwells". Advanced Materials & Processes, 167
(10), 2931.

Davis, J. R. (Ed.). (2002). Surface hardening of steels: Understanding the basics. Materials Park, OH:
ASM International.

Dieter, George E. (1989). Mechanical Metallurgy. SI Metric Adaptation. Maidenhead, UK: McGraw-Hill
Education. 0-07-100406-8

Malzbender, J. (2003). "Comment on hardness definitions." Journal of the European Ceramics


Society. 23 (9). DOI 10.1016/S0955-2219(02)00354-0

Revankar, G. (2003). "Introduction to hardness testing." Mechanical testing and evaluation, ASM
Online Vol. 8.

External links[edit]

An introduction to materials hardness

Guidelines to hardness testing

Testing the Hardness of Metals


Categories:

Condensed matter physics

Matter

Solid mechanics

Materials science

Hardness tests

Hardness comparison
From Wikipedia, the free encyclopedia
There are a large number of hardness testing methods available
(e.g. Vickers, Brinell, Rockwell, Meyer and Leeb). Although it is impossible in many cases to give an exact
conversion, it is possible to give an approximate material-specific comparison table e.g. for steels.

Hardness
Hardness comparison table[edit]

Brinell HB
(10 mm Ball, 3000 kg
load)

Vickers H
V
(1 kg)

Rockwell C HRC
(120 degree cone
150 kg)

Rockwell B HRB
(1/16" ball
100 kg)

Leeb HLD[1]

800

72

856

780

1220

71

850

760

1170

70

843

745

1114

68

837

725

1060

67

829

712

1021

66

824

682

940

65

812

668

905

64

806

652

867

63

799

626

803

62

787

Hardness
614

775

61

782

601

746

60

776

590

727

59

770

576

694

57

763

552

649

56

751

545

639

55

748

529

606

54

739

514

587

53

120

731

502

565

52

119

724

495

551

51

119

719

477

534

49

118

709

461

502

48

117

699

451

489

47

117

693

444

474

46

116

688

427

460

45

115

677

415

435

44

115

669

401

423

43

114

660

Hardness
388

401

42

114

650

375

390

41

113

640

370

385

40

112

635

362

380

39

111

630

351

361

38

111

621

346

352

37

110

617

341

344

37

110

613

331

335

36

109

605

323

320

35

109

599

311

312

34

108

588

301

305

33

107

579

293

291

32

106

572

285

285

31

105

565

276

278

30

105

557

269

272

29

104

550

261

261

28

103

542

258

258

27

102

539

Hardness
249

250

25

101

530

245

246

24

100

526

240

240

23

99

521

237

235

23

99

518

229

226

22

98

510

224

221

21

97

505

217

217

20

96

497

211

213

19

95

491

206

209

18

94

485

203

201

17

94

482

200

199

16

93

478

196

197

15

92

474

191

190

14

92

468

187

186

13

91

463

185

184

12

91

461

183

183

11

90

459

180

177

10

89

455

Hardness
175

174

88

449

170

171

87

443

167

168

87

439

165

165

86

437

163

162

85

434

160

159

84

430

156

154

83

425

154

152

82

423

152

150

82

420

150

149

81

417

147

147

80

413

145

146

79

411

143

144

79

408

141

142

78

405

140

141

77

404

135

135

75

397

130

130

72

390

Hardness
114

120

67

365

105

110

62

350

95

100

56

331

90

95

52

321

81

85

41

300

76

80

37

287

References[edit]
1. Jump up^ H.Pollok, Umwertung der Skalen (Conversion of Scales), Qualitt und Zuverlssigkeit,
Ausgabe 4/2008.
External links[edit]

Hardness Conversion Table Brinell, Rockwell,Vickers Various steels

Rockwell to Brinell conversion chart (Brinell, Rockwell A,B,C)

Struers hardness conversion table (Vickers, Brinell, Rockwell B,C,D)

Brinell Hardness HB conversion chart (N/mm2, Brinell, Vickers, Rockwell C)

Schmidt hammer
From Wikipedia, the free encyclopedia
A Schmidt hammer, also known as a Swiss hammer or a rebound hammer, is a device to measure
the elastic properties or strength of concrete or rock, mainly surface hardness and penetration resistance.

Original Schmidt Concrete Test Hammer

Hardness

Testing the compressive strength of a concrete cube using Schmidt hammer


It was invented by Ernst Schmidt, a Swiss engineer. The Schmidt hammer is distributed by Proceq and TQC
worldwide.[1]
The hammer measures the rebound of a spring-loaded mass impacting against the surface of the sample.
The test hammer will hit the concrete at a defined energy. Its rebound is dependent on the hardness of the
concrete and is measured by the test equipment. By reference to the conversion chart, the rebound value
can be used to determine the compressive strength. When conducting the test the hammer should be held at
right angles to the surface which in turn should be flat and smooth. The rebound reading will be affected by
the orientation of the hammer, when used in a vertical position (on the underside of a suspended slab for
example) gravity will increase the rebound distance of the mass and vice versa for a test conducted on a
floor slab. The Schmidt hammer is an arbitrary scale ranging from 10 to 100. Schmidt hammers are available
from their original manufacturers in several different energy ranges. These include: (i) Type L-0.735 Nm
impact energy, (ii) Type N-2.207 Nm impact energy; and (iii) Type M-29.43 Nm impact energy.
The test is also sensitive to other factors:

Local variation in the sample. To minimize this it is recommended to take a selection of readings and
take an average value.

Water content of the sample, a saturated material will give different results from a dry one.

Prior to testing, the Schmidt hammer should be calibrated using a calibration test anvil supplied by the
manufacturer for that purpose. 12 readings should be taken, dropping the highest and lowest, and then take
the average of the ten remaining. Using this method of testing is classed as indirect as it does not give a
direct measurement of the strength of the material. It simply gives an indication based on surface properties,
it is only suitable for making comparisons between samples.
This test method for testing concrete is governed by ASTM C805. ASTM D5873 describes the procedure for
testing of rock.

Roll hardness tester

Hardness
From Wikipedia, the free encyclopedia
A roll hardness tester is a device to measure the roll hardness, hardness profile and hardness variation of
paper rolls.
Contents
[hide]

1 Method

2 Standards

3 Application

4 See also

5 References
Method[edit]
In the preparation phase, the plunger, guide bar and guide disk are pushed forward by the compression
spring. At the end of the movement the hammer mass is hooked by the pawl. During the loading phase the
hammer is pushed towards the surface in a controlled movement. The hammer mass remains locked in place
by the pawl. This has the effect of stretching the impact spring to put it under tension. Impact Rebound: At
the very end of the movement, the pawl spring releases the hammer mass. The impact spring contracts
causing the hammer mass to strike against the plunger. This is the impact. The hammer mass then rebounds
back to the body of the hammer and distance travelled is recorded on the scale. The rebound distance
depends directly on the hardness of the roll under test: A softer roll will absorb more of the impact energy
and the rebound distance will be less. A harder roll will reflect more of the impact energy and the rebound
distance will increase.
Standards[edit]

TAPPI T 834 om-07: Determination of containerboard roll hardness

TAPPI TIP 1004-01: TAPPI Roll Number for inventory/tracking systems and bar codes

Application[edit]
Roll hardness is one of the most important parameters when deciding whether a paper roll is good or bad. A
roll that is wound too softly can go out of round when handled. A roll that is wound too hard, on the other
hand, can crack during transportation.[1] These variations are difficult to detect. Above all, it is typically the
variation in hardness across a given roll that relates most directly to such converting issues with soft edges
being perhaps the biggest contributor.[2]
See also[edit]

Paper mill

Pulp (paper)

Hardness

Tablet hardness testing

References[edit]
1.

Jump up^ Black Clawson Converting Machinery, "The Art of Winding Good Rolls".

2.
Categories:

Jump up^ TAPPI Standard T 834 om-07

Hardness instruments

Tablet hardness testing


From Wikipedia, the free encyclopedia
Tablet hardness testing, is a laboratory technique used by the pharmaceutical industry to test the
breaking point and structural integrity of a tablet "under conditions of storage, transportation, and handling
before usage"[1] The breaking point of a tablet is based on its shape.[2] It is similar to friability testing,[1] but
they are not the same thing.
Tablet hardness testers first appeared in the 1930s.[3] In the 1950s, the Strong-Cobb tester was introduced. It
was patented by Robert Albrecht on July 21, 1953.[4] and used an air pump. The tablet breaking force was
based on arbitrary units referred to as Strong-Cobbs. [3] The new one gave readings that were inconsistent to
those given by the older testers.[3] Later, electro-mechanical testing machines were introduced. They often
include mechanisms like motor drives, and the ability to send measurements to a computer or printer. [3]
There are 2 main processes to test tablet hardness: compression testing and 3 point bend testing. For
compression testing, the analyst generally aligns the tablet in a repeatable way, [2] and the tablet is squeezed
by 2 jaws. The first machines continually applied force with a spring and screw thread until the tablet started
to break.[3] When the tablet fractured, the hardness was read with a sliding scale. [3]
Contents
[hide]

1 List of common hardness testers

2 Units of measurement

3 Sources

4 Further reading
List of common hardness testers[edit]
There are several devices used to perform this task:

Hardness

The Monsanto tester was developed 50 years ago. The design consists of "a barrel containing a
compressible spring held between 2 plungers". The tablet is placed on the lower plunger, and the upper
plunger is lowered onto it.[5][1]

The Strong-Cobb tester forces an anvil against a stationary platform. Results are viewed from a
hydraulic gauge.[5] The results are very similar to that of the Monsanto tester.[6]

The Pfizer tester compresses tablet between a holding anvil and a piston connected to a force-reading
gauge when its plier-like handles are gripped.[5]

Erweka tester tests a tablet placed on the lower anvil and a weight moving along a rail transmits
pressure slowly to the tablet.[5]

The Dr.Schleuniger Pharmatron tester operates in a horizontal position. An electric motor drives an
anvil to compress a tablet at a constant rate. The tablet is pushed against a stationary anvil until it
fractures. A reading is taken from a scale indicator.[5]

Units of measurement[edit]
The units of measurement of tablet hardness mostly follows standards used in materials testing
the International System of Units.

Kilogram (kg) The kilogram is recognized by the SI system as the primary unit of mass.

Newton (N) The Newton is the SI unit of force; the standard for tablet hardness testing. 9.807
Newtons = 1 kilogram.

Pound (lb) Technically a unit of mass but can also be used for force and should be written as pound
force or lbf in this case. Sometimes used for tablet strength testing in North America, but it is not an SI
unit. 1 kilogram = 2.204 pounds.

Kilopond (kp) Not to be confused with a pound. A unit of force also called a kilogram of force. Still
used today in some applications, but not recognized by the SI system. 1 kilopond = 1 kgf.

Strong-Cobb (SC) An ad hoc unit of force which is a legacy of one of the first tablet hardness testing
machines.[4] Although the SC is arbitrary, it was recognized as the international standard from the 1950s
to the 1980s. 1 Strong-Cobb represented roughly 0.7 kilogram of force or about 7 newtons.[7] Although the
Strong-Cobb unit is arbitrarily based on the dial reading of a hardness tester, it became an international
standard for tablet hardness in the 1950s until it was superseded by testers using SI units in the 1980s.
[6]

The Strong-Cobb is a unit with a very unusual name for a unit of measurement since it is named after

the company, Strong-Cobb Inc. The inventor of the hardness tester was Robert Albrecht, [4] the plant
engineer for the Strong-Cobb Company. He sold the patent to the company for $1.00.

Hardness
Sources[edit]
1.

^ Jump up to:a

b c

Joseph Price Remington (2006). Remington: The Science And Practice Of

Pharmacy. Lippincott Williams & Wilkins. ISBN 0781746736.


2.

^ Jump up to:a

3.

^ Jump up to:a

b c d e f

"Tablet hardness testing". Sotax. Retrieved 16 February 2013.


"Some Information on Tablet Hardness Testing". Engineering Systems.

Retrieved 16 February 2013.


4.

^ Jump up to:a

b c

Robert Albrecht (Jul 21, 1953). "Tablet Hardness Testing Machine". United

States Patent Office. Retrieved 16 February 2013. US 2645936


5.

^ Jump up to:a

b c d e

6.

^ Jump up to:a

"Quality control of solid dosage form". Scribd. Retrieved 16 February 2013.

McCallum, A., Buchter, J. and Albrecht, R. (1955). "Comparison and correlation

of the Strong Cobb and the Monsanto tablet hardness testers". Journal of the American
Pharmaceutical Association 44 (2): 8385. doi:10.1002/jps.3030440208. PMID 14353719.
7.

Jump up^ Russ Rowlett (September 1, 2004). "How Many? A Dictionary of Units of
Measurement". University of North Carolina. Retrieved 16 February 2013.

Further reading[edit]

J. E. Rees and P. J. Rue (1978). "Work Required to Cause Failure of Tablets in Diametral
Compression". Drug Development & Industrial Pharmacy 4 (2): 131156. Retrieved16 February 2013.

American Society for the Testing of Materials (ASTM), Designation: E407, 'Standard Practices for Force

Verification of Testing Machines'.


Categories:

Hardness tests

Measuring instruments

Laboratory techniques

Pharmacology

Tablet hardness testing


From Wikipedia, the free encyclopedia
Tablet hardness testing, is a laboratory technique used by the pharmaceutical industry to test the
breaking point and structural integrity of a tablet "under conditions of storage, transportation, and handling
before usage"[1] The breaking point of a tablet is based on its shape.[2] It is similar to friability testing,[1] but
they are not the same thing.

Hardness
Tablet hardness testers first appeared in the 1930s.[3] In the 1950s, the Strong-Cobb tester was introduced. It
was patented by Robert Albrecht on July 21, 1953.[4] and used an air pump. The tablet breaking force was
based on arbitrary units referred to as Strong-Cobbs. [3] The new one gave readings that were inconsistent to
those given by the older testers.[3] Later, electro-mechanical testing machines were introduced. They often
include mechanisms like motor drives, and the ability to send measurements to a computer or printer. [3]
There are 2 main processes to test tablet hardness: compression testing and 3 point bend testing. For
compression testing, the analyst generally aligns the tablet in a repeatable way, [2] and the tablet is squeezed
by 2 jaws. The first machines continually applied force with a spring and screw thread until the tablet started
to break.[3] When the tablet fractured, the hardness was read with a sliding scale. [3]
Contents
[hide]

1 List of common hardness testers

2 Units of measurement

3 Sources

4 Further reading
List of common hardness testers[edit]
There are several devices used to perform this task:

The Monsanto tester was developed 50 years ago. The design consists of "a barrel containing a
compressible spring held between 2 plungers". The tablet is placed on the lower plunger, and the upper
plunger is lowered onto it.[5][1]

The Strong-Cobb tester forces an anvil against a stationary platform. Results are viewed from a
hydraulic gauge.[5] The results are very similar to that of the Monsanto tester.[6]

The Pfizer tester compresses tablet between a holding anvil and a piston connected to a force-reading
gauge when its plier-like handles are gripped.[5]

Erweka tester tests a tablet placed on the lower anvil and a weight moving along a rail transmits
pressure slowly to the tablet.[5]

The Dr.Schleuniger Pharmatron tester operates in a horizontal position. An electric motor drives an
anvil to compress a tablet at a constant rate. The tablet is pushed against a stationary anvil until it
fractures. A reading is taken from a scale indicator.[5]

Units of measurement[edit]
The units of measurement of tablet hardness mostly follows standards used in materials testing
the International System of Units.

Hardness

Kilogram (kg) The kilogram is recognized by the SI system as the primary unit of mass.

Newton (N) The Newton is the SI unit of force; the standard for tablet hardness testing. 9.807
Newtons = 1 kilogram.

Pound (lb) Technically a unit of mass but can also be used for force and should be written as pound
force or lbf in this case. Sometimes used for tablet strength testing in North America, but it is not an SI
unit. 1 kilogram = 2.204 pounds.

Kilopond (kp) Not to be confused with a pound. A unit of force also called a kilogram of force. Still
used today in some applications, but not recognized by the SI system. 1 kilopond = 1 kgf.

Strong-Cobb (SC) An ad hoc unit of force which is a legacy of one of the first tablet hardness testing
machines.[4] Although the SC is arbitrary, it was recognized as the international standard from the 1950s
to the 1980s. 1 Strong-Cobb represented roughly 0.7 kilogram of force or about 7 newtons.[7] Although the
Strong-Cobb unit is arbitrarily based on the dial reading of a hardness tester, it became an international
standard for tablet hardness in the 1950s until it was superseded by testers using SI units in the 1980s.
[6]

The Strong-Cobb is a unit with a very unusual name for a unit of measurement since it is named after

the company, Strong-Cobb Inc. The inventor of the hardness tester was Robert Albrecht, [4] the plant
engineer for the Strong-Cobb Company. He sold the patent to the company for $1.00.
Sources[edit]
1.

^ Jump up to:a

b c

Joseph Price Remington (2006). Remington: The Science And Practice Of

Pharmacy. Lippincott Williams & Wilkins. ISBN 0781746736.


2.

^ Jump up to:a

3.

^ Jump up to:a

b c d e f

"Tablet hardness testing". Sotax. Retrieved 16 February 2013.


"Some Information on Tablet Hardness Testing". Engineering Systems.

Retrieved 16 February 2013.


4.

^ Jump up to:a

b c

Robert Albrecht (Jul 21, 1953). "Tablet Hardness Testing Machine". United

States Patent Office. Retrieved 16 February 2013. US 2645936


5.

^ Jump up to:a

b c d e

6.

^ Jump up to:a

"Quality control of solid dosage form". Scribd. Retrieved 16 February 2013.

McCallum, A., Buchter, J. and Albrecht, R. (1955). "Comparison and correlation

of the Strong Cobb and the Monsanto tablet hardness testers". Journal of the American
Pharmaceutical Association 44 (2): 8385. doi:10.1002/jps.3030440208. PMID 14353719.
7.

Jump up^ Russ Rowlett (September 1, 2004). "How Many? A Dictionary of Units of
Measurement". University of North Carolina. Retrieved 16 February 2013.

Hardness
Further reading[edit]

J. E. Rees and P. J. Rue (1978). "Work Required to Cause Failure of Tablets in Diametral
Compression". Drug Development & Industrial Pharmacy 4 (2): 131156. Retrieved16 February 2013.

American Society for the Testing of Materials (ASTM), Designation: E407, 'Standard Practices for Force

Verification of Testing Machines'.


Categories:

Hardness tests

Measuring instruments

Laboratory techniques

Pharmacology

Persoz pendulum
From Wikipedia, the free encyclopedia

Persoz pendulum
A Persoz pendulum is a device used for measuring hardness of materials. The instrument consists of
a pendulum which is free to swing on two balls resting on a coated test panel. The pendulum hardness test is
based on the principle that the amplitude of the pendulum's oscillation will decrease more quickly when
supported on a softer surface. The hardness of any given coating is given by the number of oscillations made
by the pendulum within the specified limits of amplitude determined by accurately positioned photo sensors.
An electronic counter records the number of swings made by the pendulum

Hardness
Construction[edit]

Persoz pendulum device


The pendulum consists of balls which rest on the coating under test and form the fulcrum. The Persoz
pendulum is very similar to the Konig pendulum. Both employ the same principle, that is the softer the
coating the more the pendulum oscillations are damped and the shorter the time needed for the amplitude of
oscillation to be reduced by a specified amount. The two pendulums differ in shape, mass and oscillation
time, and there is no general relationship between the results obtained using the two pieces of equipment. In
either case, the test simply involves noting the time in seconds for the amplitude of swing to decrease from
either 6 to 3 degrees (Konig pendulum) or 12 to 4 degrees (Persoz pendulum).[1]
References[edit]
1.
Categories:

Jump up^ PRA. Mechanical Properties. Accessed: May 6, 2015.

Hardness instruments

Materials science

Pendulums

Coatings

Você também pode gostar