Anda di halaman 1dari 8

http://www.materialsaustralia.com.au/lib/pdf/Mats.%20Forum%20page%20118_125.

pdf

MATERIALS FORUM VOLUME 30 - 2006 Edired by R. Wuhrer and M. Cortie Institute of Materials Engineering Australasia Ltd

TESTING MATERIALS HARDNESS SCIENCE, TECHNOLOGY & APPLICATION


Zbigniew H. Stachurski Department of Engineering, College of Engineering and Computer Science The Australian National University, Canberra, ACT 0200 ABSTRACT In everyday life solid material objects come in contact with each other, exerting and transferring forces. What happens to the materials at the contact depends on the relative hardness of the two and the magnitude of the stresses induced relative to their yield or fracture strengths. To be able to predict what can happen, the hardness of materials is tested and measured. The science of mechanics of materials aims to relate hardness to the more fundamental properties, such a modulus of elasticity, fracture strength, onset of plastic deformation, plastic flow and strain hardening. Depending on the ratio of applied stress to material strength, the resulting deformation can be elastic (Hertzian), or a permanent deformation which can be brittle (Griffiths fracture) or ductile (von Mises yielding). A review of the most common methods of hardness testing is presented, followed by a study of the interface between glass and polymer matrix using nano-scratch technique.

1. BACKGROUND 1.1 The Ages of Civilisation (i) (ii) (iii) the Stone Age the Bronze Age the Iron Age (to approx. 5000 BC) (to approx. 2000 BC) (from approx. 2000 BC)

With a more penetrating insight, the Iron Age can be subdivided into several more specific periods. In particular, the period from about 1850 to 1950 should be named the Century of Steel. The Industrial Revolution had begun with the riches of resources brought in from the Colonies and spurred on by the new and mechanised manufacturing methods. It had unleashed great demands for strong, hard and durable materials, of which steel was the best. The Industrial Revolution is known to us not only through the misery, poverty and human exploitation described by Charles Dickens and Karl Marks, but also through the eyes of science fiction and adventures of Jules Verne, From the Earth to the Moon (1864) and 20,000 Leagues Under the Sea (1867), and through great engineering accomplishments. These were inspired by the ability to make new machines of incredible strength and unimaginable reach. 1.2 Engineering The Century of Steel represented an immense increase in engineering works and endeavours in the industrialised countries. Isambard K. Brunel in Great Britain was a major force in building railways, ships (Great Western, Great Eastern and others) and bridges (suspension bridge in Bristol), in Europe railways were built: Moscow to Vienna, Berlin to Paris, London to Manchester, etc. In USA two separate railways lines were built simultaneously from east to west coast, as well as north 118

to south lines. The first steel bridge in the USA was built in 1874 over the Mississippi river at St. Louis, followed by the John Roebling Brooklyn bridge in 1883. The Golden Gate bridge built in 1930 required 120,000 km of steel wire for its suspension cables. Ship-building of massive battle cruisers and aircraft carries were undertaken in Great Britain, USA, Japan. Motorcar industry grew around the turn of the century (Daimler and Benz in Germany 1890, Renault in France 1905, Ford in USA 1908). Farming and agricultural machinery has become motorised and grew in size, particularly in the large scale production methods in the USA. Electric power generation required large steel works, and bridge building consumed large amounts of high strength steel. 1.3 Steel Production Steel (a modification of cast iron) was known for a long time through a laborious process of reheating and reworking of meteorites or bloom iron. Around 1750 Huntsman of Sheffield devised a method heating scraps of steel and blister iron in sealed fireclay crucibles with addition of oxygen which both burnt out most of the carbon dissolved in the metal and also kept the temperature high. In 1856 Henry Bessemer obtained a patent for making steel by the tons using a fireclay lined converter which blaster air under high pressure into the molten metal. At the same time Siemens brothers (William and Friedrich) and Martin in France devised the open hearth furnace which proved to be even nore flexible and economical. Thus the age of steel has arrived in Britain, Germany and France. Within a decade the open hearth and converter were being used by every major steel and iron manufacturer. The next advance in making high quality steel was the electric furnace which was

invented again by Siemens and adopted particularly in France were electricity was in sufficient supply. Soon, steel works opened up in the USA (Pennsylvania being the main centre with ready supply of iron ore and coal). Since then the growth of steel production has been phenomenal (see Table 1). Table 1. The production of steel world-wide and in USA from 1850 to 2004. Year 1850 1867 1880 1910 2004 World [1000s of tonnes] ~ 0.1 ~ 500 ~ 1,750 ~ 35,000 ~ 1,000,000 USA [1000s of tonnes] ~0 ~ 10 ~ 1,000 ~ 24,000 ~ 150,000

The building and use of large steel structures and machines involved accidents. Accidents were followed by liability, and in turn, by increased costs of insurance. Investors required guarantees from engineers, and engineers imposed testing and standards on manufacturers. This was one of the main reasons for the need and for the introduction of testing materials hardness.

1.4 Why Testing Materials Hardness is Important The relative hardness of materials has been known in ancient times. Since stone and rock were the main building materials, it was important to know which stone (mineral) can be used to cut and break other materials. During the Yuan dynasty in China (13th century) large ship building was undertaken, and the hardness of different kinds of wood from its vast empire were assessed and recorded. George Agricola, in his book De re Metallica in 1556 described how mining can be accomplished with the use of hard materials, and how blocks of suitable stone can be cut from the rock (see Figure 1a). In 1822 Mohs published a scale of hardness as we know it today (Figure 1b). Although it refers to hardness, the test is essentially that of resistance to scratching and cutting of one brittle material by another. H 1 2 3 4 5 6 7 8 9 10 Mineral Talk Gypsum Calcite Fluorite Apatite Feldspar Quartz Topaz Corundum Diamond

Figure 2. Raiwlay disasters followed by liability claims and insurance costs.

2. THE METHODS 2.1 The Brinell Method of Hardness Testing The first modern hardness test on steels was introduced in 1900 by a Swedish engineer, D.A. Brinell of the Fagersta Iron Works company, and it is known as Brinell hardness. The method consists of indenting the metal with a 10 mm diameter steel ball subjected to a load of 3000 kg (~30kN) for 30 seconds (later a lower load was introduced, 500 kg (~5kN), for soft metals). Then the Brinell hardness number (BHN) was calculated as follows [1]:

BHN =

2L D D D 2 d 2

(1)

Figure 1. (a) A rock cutter with diamond tipped hand drill, and (b) Mohs scale of mineral hardness. The confusion between testing brittle materials (rock and minerals) and ductile materials (metals) prompted L.B. Tuckerman of the National Bureau of Standard in the USA to describe the state of hardness testing at the beginning of the last century as a hazily conceived conglomeration or aggregate of properties of a material more or less related to each other. But to a construction engineer, the term hardness means resistance to indentation (plastic deformation), and, therefore, refers to a ductile material.

where: L is load in kilograms, D is diameter of the ball in millimetres, and d is the diameter of indentation in millimetres. Note that the denumerator represents the surface of the indentation (see Figure 3). As found through experience, the method suffered a major problem of inconsistency (see Figure 4). This is due to the fact that hardness of the material increases with depth of penetration. After yield (limit of elastic behaviour) the resistance of the material to plastic deformation increases with strain. The plastic part of the stress-strain curve is usually described by a parabolic law: = kn, where n (~0.5) is called the strain hardening exponent.

119

2.3 The Rockwell Method

Figure 3. Indentation of a hard sphere into a ductile metal. 2.2 The Meyer Analysis

Eugene Meyer of the Materials Testing Laboratory at the Imperial School of Technology (Charlottenburg, Germany) made intensive study of Brinell method and published the results in 1908 [2]. Meyers work showed that resistance to penetration by a ball varies with the degree of penetration, and follows the relation:
L = ad m

The difficulties of Brinell method were overcome by the introduction of the Rockwell test in which the depth of penetration (rather than the diameter) is measured. In 1919 Stanley Rockwell, a metallurgist in a ball bearing company, devised a machine with the indenter as 1/16 inch steel ball, pressed into the surface with an initial load of 10 kg (~100N), followed later by 100 kg (~1kN) applied for 5 seconds [1]. To begin with, an initial load is applied to settle in the tested piece, then the major load is applied. The difference in depth between the two positions is taken as the hardness number. Later, two more weights were included (60 kg and 150 kg), and different diameter balls were introduced (1/8, 1/4, 1/2 inch, and diamond Brale indeter as well). Most importantly, the reading of the hardness was taken at the time when the major load was realeased back to the minor load, so that the compliance of the machine was eliminated from the measurement (see Figure 5).

(2)

where: L is the load, d is diameter of indentation, a is resistance of the material to initial penetration, and m is measure of the effect of the deformation on the hardness of the metal, and to a good approximation equal to the reciprocal of the strain hardening exponent as defined above in section 2.1. Meyer showed that a better measure of hardness is the mean pressure under the indenter, Pm, known as the Meyer hardness number:

Figure 5. The principle of Rockwell hardness measurement (max - min) depth of indentation.

Pm =

4L

d 2

(3)

This relationship is confirmed by the diagram shown in Figure 4.

The units of hardness are so chosen that one point of hardness is equal to a depth of only 0.002 mm. Many changes were introduced to the Rockwell tester. The scale was reversed so that low numbers corresponded to low hardness and high numbers corresponded to high hardness, respectively. Different scales were assigned to indentors and weights. The consistency of measurements and versatility of the technique has made Rockwell Method the most comonly used in engineering, both in metallurgical laboratories and engineering manufacturers.
2.4 The Vickers Method

This method was introduced in England in 1925 by R. Smith and G. Sandland [3]. It is similar to the Brinell method, in that the hardness number is calculated as:

DPH =
Figure 4. left - Curve 1 shows how the diameter of the impression increases as the load on the ball is increased, Curve 2 shows the hardness number as determined by the Brinell formula (1), Curve 3 shows the mean pressure values as determined by formula (3), Figure 1. right - determination of Meyers constants a and m (see text).

2L sin( / 2) d2

(4)

where: d is the diagonal of the impression, and is the angle between opposite faces of the diamond = 136.
2.5 The Microhardness Methods

As the relationship between the microstructure and properties of materials increased, a need appeared for a 120

finer indentation methods [4]. One of the first was the Knoop method in which a specially shaped diamond indentor (see Figure 6 top) is pressed into the surface with a force of between 0.1 kg (~1N) to 1 kg (~10N), depending on the hardness of the tested area [1]. As in Brinell and Vickers methods, the dimension of the indentation is measured and related to corresponding hardness scale. Other makers of the instruments use pyramid-like diamond indentor, as seen in Figure 6 bottom. This allows for measuring hardness of coatings, case-hardening and differences in hardness between grains.

System). It is capable to apply 50 mN force with movement resolution of 0.05 nm. There are some 30 of the machines around Australia, and over 55 around the world. It was originally licensed to Australian Scientific Instruments (ASI) Pty. Ltd. Now FischerCripps Laboratories Pty Ltd has taken over all UMIS activities. UMIS was one of the first nanoindenters on the market, exhibited in San Diego in 1989. One of these machines is in the Electronic Materials E at ANU (see the October 2003 issue of Materials Monthly for details).

3.2 Other Nano-indentation Methods

Recently, the ANU has just received one of the worlds most sophisticated nanotech indenters and it promises to revolutionise the way we interact with materials at the nanoscale. Its called the Hysitron TriboIndenter. Using a diamond-tipped probe, it can put nanoscale indentations into a range of materials with astounding precision and sensitivity writes David Salt of CSEM [5,6]. The nano-indentation test was originally designed for investigation of materials properties of thin films and surfaces [7]. The atomic force microscope (AFM) is an instrument designed to measuring forces between atoms and molecules. It is the ultimate instrument for measuring hardness of materials at the smallest scale. The indentor, usually specially prepared single crystal (see Figure 7) is attached to the end of a very fine, high compliance cantilever beam approximately 0.1 mm long (see Figure 7). The relative movement of the indentor and sample is achieved by piezoelectric transducers. Two modes are possible: (i) direct indentation, and (ii) scratch or friction. The principle of operation is sketched also in Figure 8.

Figure 6. Top left the geometry of Knoop indentor, top right indentations on case hardened block of steel, showing variations towards the edge, bottom pyramid-type indentor on a different tester, impression in silicon wafer showing cracks running from the corners of the indentation.

2.6 Testing Hardness of Soft Materials

With the growth and development of the rubber and plastics industries specialised testers were developed for these new materials: (i) Shore Rubber Tester, and (ii) Barcole Tester for plastics.

3. THE AGE OF SCIENCE 3.1 The UMIS method

The field of nanoindentation creating indentations on the scale of nanometres is more of a recent innovation, and really only took off in the 1980s. It began in an effort to measure the hardness of ion-implanted surfaces to measure improved wear resistance. With the implanted region being only 100-200 nm deep, there was a need for indentation techniques that could operate at this scale. Australia has played a major role in development of this science with Ron Hutchings of ANSTO and John Field of CSIRO being co-authors of the original patent for nanoindentation, and Mike Swain managing one of the first instruments in the field while working with CSIRO. That instrument is the UMIS (Ultra Micro Indentation 121

Figure 7. on the left a view of an atomic force microscope (AFM), on the right normal mode above and scratching mode below.

with the new crack surfaces. The criterion for brittle fracture is given by:
E P f = c = Ai cc

(6)

where: f is the fracture stress equal to the ratio of the applied critical force, Pc, over the contact area Ai, E is the modulus of elasticity, is the surface energy and cc is the critical surface crack (defect) size pre-existing in the brittle solid, and is a geometrical factor [9]. The crack extension force for a cone fracture, as shown in Figure 9, is given by:
G= k ( ) 1 2 d P = P2 = 2 2 dc Ec 3

(7)

where: is the compliance of the tested system, c is the crack length, and k(v) is a material related constant.

Figure 8. top highly pointed indentor shown upsidedown, bottom the feedback system of an AFM.

4. THE SCIENCE OF DEFORMATION 4.1 The elastic contact of solid surfaces

Consider a ball, of radius, R, with elastic modulus, E, and Poissons ratio, v, pressed in contact with a hard elastic surface. If the ball is pressed into the surface with a force less than that required to cause permanent deformation (i.e., within the regime of Hookes law), then the resulting elastic deformation, known as Hertzian solution, is given by [8]:
5 1 2 3 2 E R 2 a 3 = , a = PR (5) , U = 5 2 R 1 E 4
2
1

Figure 9. A cone fracture produced in glass block with steel sphere indentation. 4.3 The Huber/von Mises Theory of Plasticity

where: is the distance of approach of two undeformed points between the centre of the sphere and a point in the bulk of the surface, a is the radius of the circle of contact between the ball and the surface, P is the applied force pressing on the ball, and U is the elastic stored strain energy. Notice that the properties of the flat surface, onto which the ball is pressed, do not appear in this relationship.
4.2 The Griffiths Theory of Brittle Fracture

In 1908 Huber, working in the Lvov (Lemberg) University of Technology, published a generalised theory of plastic deformation of ductile materials based on the distortion energy criterion. A graduate of the university, Richard von Mises, who later moved to England, was to become famous for the so-called von Mises yield criterion, based on Hubers theory.
2 ( 1 2 ) 2 + ( 2 3 ) 2 + ( 3 1) 2 = 2 y

(8)

In 1921 A.A. Griffith published a paper in which he proposed a new theory of fracture of brittle materials containing a crack. He based his theory on the balance of energy between the work done by the applied stress to extend the crack versus the surface energy associated 122

where: the stresses with numeral subscripts represent principal stresses, and y is the yield strength of the material in tension. Substitution of the Hertzian (principal) stresses in the above equation leads to the simple (although approximate) relationship between Brinell hardness expressed in MN/m2, and yield strength expressed in the same units:

BH 3 y [MN/m2 ]

(9)

5. THE APPLICATION 5.1 Hardness of fibre-matrix interface

The cross-section of a glass fibre reinforced composite material was studied by nano-indentation and nanoscratch technique in order to ascertain if an interface between the two phase could be detected [10]. The existence of the interface is thought to arise as a result of intermixing and entanglement of the matrix resin with the excess sizing applied to the surface of the glass fibre during manufacturing. Glass fibres are typically of 20 m diameter, imbedded in a crosslinked polymer matrix. A cross-section of a sample studied is shown in Figure 10.
Figure 11. A view of surface topology with indentations as revealed by AFM scan.

To find the existence of an interface we have applied the scratch mode of testing, in which the indenting tip is dragged along the surface, crossing over the fibres and the matrix regions. The resulting force of friction is measured, and it is plotted on a graph as shown in Figure 12. The nano-scratch test was used in order to investigate the width of the fibre-matrix interphase region. This novel technique involves moving a sample while being in contact with diamond tip. The coefficient of friction is determined from the fraction of the lateral and the normal force. Therefore, the coefficient of friction indicates the resistance of the material to the tip penetration in the tangential direction. In this work, the normal force was kept constant during the experiment. The tip was moving from the matrix to the fibre gradually decreasing its penetration (profile) depth after contacting the harder interphase region. A detailed information about the test is available elsewhere [10]. The scratch length was about 60 m starting from the matrix and crossing two fibres in this range that were found on the surface of each sample. Two matrix/fibre interphase regions were investigated in one run. (The part of the scratch path from the fibre to the matrix could not be analysed due to the loss of balance in the system when the tip suddenly drops to softer material). The experiments were carried out with two values of the normal force, 0.4 mN and 1 mN, in order to investigate the influence of the penetration depth to the final measurement of the interphase width. In the figure below are two experimental traces, one of the position (profile) of the indentor with respect to surface level, and the second is the friction coefficient calculated as the ratio of the tangential force over the normal force applied to the indentor. Also sketched on the diagram are construction lines indicating the leading edge of the indentor, surface level, and lines indicating transition.

Figure 10. A cross-section through a composite material comprising glass fibres imbedded in a polyester resin.

The apparatus used in this work was Nano Indenter II made by Nano Instruments, Inc. A detailed description of the instrument is available elsewhere [11]. A sample prepared for testing is mounted in the sample holder and positioned by the motorised precision table, visually controlled through the optical microscope. The Berkovich indenter, mounted on the indenter head, is a three-sided triangular-based pyramidal diamond. Depths of indents were programmed to have a constant value of 30 nm. Displacements of indentation depths were consistent to within 5 nm. From the shape of the Berkovich indenter it can be inferred that the indents were 210 nm wide. Each successive indent was displaced by 260 nm in order to avoid overlapping of plastic deformation zone onto neighbouring indents. The indents were made along a path of approximately 7 m in the matrix and 7 m in the fibre. Figure 11 shows a series of traversing indentations in the matrix, as recorded by the topographic mode of the AFM. The depth of the indentations is approximately 30 nm, and the force required to induce these indentations was of the order of 20N. For the same indentation depth, the force increased to approximately 150 N for the interface, suggesting increased hardness of about 7.5 times, but less than that of glass.

123

0.7

100

Coef.Fric. Profile
0.6

~ surface level (2)

0.5

The leading edge of the tip, moving in the scratch direction along the scratch profile.

-100

0.4 -200 0.3 -300 0.2

(1)
0.1

profile slope the leading edge of the tip


8 9 10 11 12 13

-400

width of the interphase region


-500 14 15 16

Scratch length ( m) scratch direction

Figure 12. Test results of nano-scratch over the surface of a composite material. The variation of profile depth and friction coefficient indicate that the polymer phase is on the right-hand side and the glass phase is on the left-hand side.

The results of a nano-scratch test are shown in Figure 8. In the diagram, the profile is high on the left, and low on the right, and friction coefficient is high on the right and low on the left. This indicates that the softer phase (the polymer) is on the right hand side of the diagram and the harder phase (glass fibre) on the left side. The leading edge of the indentor is shown in the diagram as an inclined line, advancing towards the glass fibre on the left. In position (1) it is over the polymer matrix, with the lowest profile depth and highest coefficient of friction (approx. 0.35). In position (2) it has come in contact with the cylindrical edge of the glass fibre, and it is beginning to rise rapidly to a higher profile, whilst at the same time friction coefficient is decreasing to a lower value (approx. 0.2). We deduce that the distance between positions (1) and (2) is a transition from pure matrix to pure glass phases. It is the interface region. To confirm our observations, the nanoindenter was positioned over the phases: the glass fibre, the interface, and the polymer matrix. The results obtained in normal mode testing are shown in Figure 13. The instrument provides sufficient data for the loading and unloading points to appear as continuous curves. The upward turning curve during the loading part indicates, as expected, an increasing resistance to deformation with increasing depth in indentation. On unloading the curve does not return to zero indicating that permanent deformation has taken place (brittle fracture in the case of the glass fibre, and ductile deformation in the case of the matrix and interface).

Figure 13. Load versus indentation depth for three areas on the composite surface, the glass fibre, the interface, and the polymer matrix.

Hardness of material calculated from an indent produced by Berkovich tip is calculated from the equation below:
H= P
2 24.5 hc

(10)

where: P is the load and hc is the contact depth of the indent. Hardness values for polyester/glass, systems range from approximately 10 GPa for the glass to less than 1 GPa in the matrix. The transition region observed between the matrix and the fibre shows material properties distinct from those of the matrix and the fibre. The width of the transition zone is approximately 1 m.

124

6. SUMMARY 6.1 Static hardness testing involves instruments capable of the following range of indentation forces:

atomic force indentation (AFM) nanohardness (TriboIndenter) ultramicrohardness (UMIS) microhardness (Knoop, Shimadzu) bulk hardness (Rockwell, Brinell)
6.2 Dynamic hardness testing

(1 pN to1 N) (1 N to 1 mN) (1 mN to 1 N) (100 to 103 N) (103 to 104 N)

microscale macroscale gigantic scale

grinding wheel on steel shaft abrasive disk on ceramic tiles cutting tool on machined piece impact by meteorites (or man made objects)

6.3 Materials properties which can be derived from hardness testing:

Engineering strength of metals Composition of metals and ceramics Wear of materials Anisotropy of materials surface Cutting ability of materials Abrasive strength of materials Fracture strength Surface coating quality, depth and structure Hardness of different components of alloys Degree of crystallinity in thermoplastic polymers Degree of crosslinking (curing) in thermosetting polymers Glass transition temperature Time-dependent viscoelastic properties

W.C. Oliver, R. Hutchings, and J.B. Pethica, Measurement of hardness at indentation depths as low as 20 nanometres. Microindentation Techniques in Materials Science and Engineering. ASTM STP 889, P. J. Blau and B. R. Lawn, Eds., American Society for Testing and Materials, Philadelphia, 1986, pp. 90-108. 8. L.D. Landau and E.M. Lifshitz, Theory of Elasticity, Pergamon Press, Oxford, 1956. 9. B.R. Lawn and T.R. Wilshaw, Fracture of Brittle Solids, Cambridge Solid State Science Series, Cambridge, 1975. 10. A. Hodzic, Z.H. Stachurski and J.K. Kim, The nano-scratch technique as a novel method for measurement of an interphase width, J. Mater. Sci. Letters, Vol.19 (2000) 1665-1667, also A. Hodzic, Z.H. Stachurski and J.K.Kim, Nanoindentation and Nano-scratch of Polymer/Glass Interface. Part I: Experimental and Mechanical Analysis, Polymer, Vol.41 (2000) 6895-6905, also A. Hodzic, J.K. Kim and Z.H. Stachurski, Nano-indentation and Nano-scratch of Polymer/Glass Interface. Part II: Model of Interfaces in Water Aged Composites, Polymer, Vol.42 (2001) 5701-5710. 11. B. Bharat, Handbook of Micro/Nanotribology. CRC Press, 1995. 7.

Bibliography

1.

References

1. 2.

3.

4. 5.

6.

Vincent E. Lysaght, Indentation Hardness Testing, Reinhold Publishing Corp., New York, 1949. E. Meyer, Untersuchingen ber Hrtepfrng und Hrte Brinell Methoden, Z. Ver. Deut. Ing., 52, 1908. Some Notes on the Use of a Diamond Pyramid for Hardness Testing, J. Iron Steel Inst. (London), III, 1925. B.W. Mott, Micro-indentation Hardness Testing, Butterworths Scientific Publications, London, 1956. David Salt (Ed.) Materials Monthly, The ANU Centre for Science and Engineering of Materials, www.anu.edu.au/CSEM/. http://www.hysitron.com/Products/ProductPages/pro ducts_triboindenter.htm.

The Macquarie History of Ideas, Publ.: Macquarie Library Pty. Ltd, Melbourne, Australia 1983. 2. A.F. Burstall, A History of Mechanical Engineering, Faber, 1963. 3. E. de Bono, An Illustrated History of Inventions from the Wheel to the Computer. 4. R.A.S. Hennessey, Railways, Batsford Press, 1973. 5. Judith Jackson, Man and the Automobile, McGraw-Hill, 1979. 6. H.L. Peterson, Encyclopaedia of Firearms, Connoiseur Press, 1970. 7. L. Rogin, The Introduction of Farm Machinery, University of California Press, 1931. 8. Time Life International, The Great Ages of Man series, 1966. 9. Arthur Street and William Alexander, Metals in the Service of Man, Penguin Books, London, 1989. 10. L.T.C. Rolt, Isambard Kingdom Brunel, Pelican Books, London, 1970. 11. J.E. Gordon, The New Science of Strong Materials, or why dont you fall through the floor, Pelican Book, London,1968 12. E. Foner, and J.A. Garraty (Eds.), The Reader and Companion to American History, Houghton Mifflin Co., N.Y., 1991.

125

Anda mungkin juga menyukai