Sei sulla pagina 1di 28

Accepted Manuscript

Technical report

Investigation on the Effect of Lubrication and Forming Parameters to the Green


Compact Generated from Iron Powder through Warm Forming Route

M.M. Rahman, S.S.M. Nor, H.Y. Rahman

PII: S0261-3069(10)00388-2
DOI: 10.1016/j.matdes.2010.06.013
Reference: JMAD 3182

To appear in: Materials and Design

Received Date: 2 February 2010


Accepted Date: 7 June 2010

Please cite this article as: Rahman, M.M., Nor, S.S.M., Rahman, H.Y., Investigation on the Effect of Lubrication
and Forming Parameters to the Green Compact Generated from Iron Powder through Warm Forming Route,
Materials and Design (2010), doi: 10.1016/j.matdes.2010.06.013

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT

Investigation on the Effect of Lubrication and Forming Parameters to the Green Compact

Generated from Iron Powder through Warm Forming Route

M. M. Rahman*, S. S. M. Nor & H. Y. Rahman

Department of Mechanical Engineering, Universiti Tenaga Nasional

Putrajaya Campus, Selangor Darul Ehsan, Malaysia

e-mail: mujibur@uniten.edu.my

Abstract

In order to generate green compacts of iron ASC 100.29 powder at above ambient temperature and

below its recrystallization temperature, a warm compaction rig is designed and fabricated which

can be operated at various temperature and load. The aim of this paper is to present the outcomes

of an investigation on the effect of lubrication and forming parameters, i.e, load and temperature

to the green compacts generated through warm compaction route. The feedstock was prepared by

mechanically mixing the main powder constituent, i.e, iron ASC 100.29 powder with different

weight percent of zinc stearate at different mixing time. Compaction load was varied from

105 kN to 125 kN using simultaneous compaction mechanism. The microstructures of the green

compacts were analyzed by Scanning Electron Microscopy (SEM), and the mechanical properties

are measured through density measurement, hardness test and electrical conductivity test. The

study found that increase in compaction load as well as forming temperature give improved

microstructure and mechanical properties. It is also found that effects of lubrication to the

mechanical properties of green compacts are strongly dependant on the lubricant content as well

as mixing time of iron powder with the lubricant.

Keywords: C. powder metallurgy; E. mechanical; G. scanning electron microscopy

*
Corresponding author. Tel: +6 03 89287269; Fax: +6 03 89212116

1
ACCEPTED MANUSCRIPT

1. Introduction

Nowadays, the number of engineering components and automotive parts manufactured from

shaped and sintered metal powders has been increasing. In the last three decades, a wide range of

structural components especially for the use as automotive parts has been developed for production

using this method [1]. Most of them are manufactured by a process called die compaction method

through warm forming route or conventional cold forming route. The production of solid

components from powders by compaction and sintering is known as powder metallurgy, which has

been existence since the early nineteen hundreds as a new generation of manufacturing process [2].

Powder compaction process is the production of any powder material by compacting in a

container to a desired shape. The compacted powder mass is called a green compact, which only

has sufficient strength to be handled for further treatment [3]. Powder compaction is an attractive

forming process since it offers an approach to net-shape or near-net-shape manufacturing. The

main objective in powder forming process is to obtain green compacts with relatively high density

and uniform density distribution in the component [4]. Metal powder may be compacted either at

room temperature, which is termed as conventional cold compaction, or at elevated temperature

which is warm compaction, which was introduced in the market in 1994 [5].

Compaction process involves the transfer of pressure to each particle contained in the

powder. Generally, the compaction mechanism relates directly to the density and the pressure [6].

When a powder is being compacted in a rigid cylindrical die, the axial pressure exerted upon the

powder by the compaction punch is only partly transformed to radial pressure upon the die wall.

This radial pressure can be quite substantial but it does not reach the level of the axial pressure

because the powder is not a liquid and does not have any hydraulic properties [7]. Aggregates of

metallic particles are considered as frictional, nearly non-cohesive and granular [8]. When the load

acts upon the powder in the die cavity, the powder deforms plastically. Each of the particles fills

the gap among them. This results in reduction of pores. This means, density increases inside the

2
ACCEPTED MANUSCRIPT

compacted powder. During compaction, powder exhibits strain or work hardening, the volume

decreases and hardness increases [9].

Previous study on the effect of temperature to the deformation of powder has been

conducted by Wechsler [10]. The study found that, the strength of the product is less dependent on

forming temperature and pressure, but is strongly influenced by density, particle size and shape.

However, an experimental work on microscopic aspect has been conducted by Gagne [11] in

classifying the behavior of metal powder under compaction at 23°C and 150°C using Scanning

Electron Microscope (SEM). This work showed that there is influence of temperature to the

particle arrangement in powder compact. It also found that there are interaction forces among

powder particles where this phenomenon lead to the enhancement of the strength of the green

compacts.

The influence of temperature to yield point of green compact has been presented by

Höganäs [12] and stated that at higher temperature, the green compact experiences lower value of

yield point. This principle of a temperature-dependant yield point emerges from the theoretical

hysteresis curves. The curves clarified that maximum radial pressure increases but the residual

pressure, after complete release of the axial pressure, decreases when the yield point is lowered at

elevated temperatures. Therefore, it can be stated that the deformation of metal can be looked upon

as thermally activated process [13]. Warm compaction experiments at forming temperature from

100°C to 150°C are conducted and reported in [14], also found to obtain a green densification level

of greater than 98%. Furthermore, Rahman [15] has presented the experimental work on a full

cycle of warm powder compaction process that shows a decreasing trend of friction with the

increase of forming temperature.

The optimum forming temperature in generating crack-free green compacts is important for

obtaining the most effective production system. Rutz et al. [16] has found that the optimum

temperature for ferrous powder during the warm compaction is 145°C. However, no further

3
ACCEPTED MANUSCRIPT

evidence from other researches is found to support this finding. In powder compaction process, a

considerable friction is generated which is the major cause of density variation throughout the

powder compact and also affects the component deformation [17]. Inter-particle and particle to die

wall friction also hinders pressure transmission and therefore produces density gradients inside the

compact [14]. The friction also generates some amount of heat that affects the compaction process

generally, and material properties specifically.

Previous researchers such as Rahman [15], Tabata et al. [18], Tran et al. [19], Ariffin [20],

Aidah [21], and Khoei & Lewis [22] worked on powder compaction including the study of friction.

The understanding of friction phenomenon in warm powder compaction enhances the capability in

solving the compaction problem numerically. They showed that friction is caused largely by

powder-die wall interaction and inter-particle friction among the powder particles. Therefore, there

is a need to use the lubricant as friction reduction agent. Two types of lubrication method that have

been studied by previous researchers, i.e., admixed and die wall lubrication method. However, this

study concentrates only on the admixed lubrication method, since die wall lubrication method

needs to manufacture high technology die that incurs extra manufacturing cost.

The use of an admixed lubricant may create problems in compaction, and it is therefore

important that the lubricant is well suited to the type of compaction carried out. In order to

generate defect-free green compacts, the lubricant should be forced out from the pore structure of

the powder composition during the compaction operation. Another reason to remove the lubricant

from the compact is that it would otherwise create pores in the compact after sintering. It is well-

known that large pores have an adverse effect on the dynamic strength properties of the product

[23]. Another problem in using the lubricant is that most of the lubricants used in cold compaction

cannot be used for the compaction at elevated temperature, since they seem to be effective within a

limited temperature range only.

4
ACCEPTED MANUSCRIPT

Interesting findings reported by Zhornyak & Oliker [24] on the optimization of zinc stearate

content in ensuring the greatest compressibility of a powder. The higher the density of a compact

and the smaller the pore size, the smaller is the amount of zinc stearate it can hold. The weakening

action of zinc stearate, which manifests itself in stress relief at inter-particle contacts during

pressing, substantially increases the density of compacts and at the same time decreases their

strength. This phenomenon is particularly pronounced with compacts from poor compactibility

powders. After pressing at any given pressure, the strength of a compact with zinc stearate is

higher than the strength of a compact without lubricant. This is due to the decrease of the contact

surface brought about by the elastic aftereffect in the zinc stearate which is counterbalanced by the

increase of this surface resulting from better compressibility of the powder. They found that the

addition of an optimum amount of zinc stearate to the iron powder is always more effective than

lubrication of the die walls and compaction without lubrication.

Previous researches have given significant knowledge on powder compaction involving its

parametric analysis. However, all reported research works have been conducted at the maximum

forming temperature of 150°C. Furthermore, the studies on admixed lubrication by previous

researchers considered just only involving the weight percent of lubricant without considering the

importance of mixing time. Therefore, this study concentrates in expanding the scope of warm

compaction parameters by considering different loading, lubricant content, mixing time as well as

forming temperature.

2. Experimental Works

A lab-scale warm compaction tool was designed and manufactured (Figure 1) which enabled the

generation of cylindrical shape green compacts. Iron powder ASC 100.29 was used during the

experiment because this type of powder is normally used in most of the powder compaction

industries [29, 30]. The as received powder has the particle size range of 20-180 µm. The powder

5
ACCEPTED MANUSCRIPT

manufactured by Höganäs AB Company has the composition of 1.5% Cu, 0.5% Mo, and 4% Ni

balanced with Fe.

A cold to elevated temperature compaction experiment is conducted using simultaneous

multi axial compaction mechanism. Four point heaters (50 Watt each) are attached to the

cylindrical shape die. The forming temperature is ranged from 30°C to 180°C while the

compaction load is ranged from 105 kN to 125 kN. However, the lubrication study is conducted at

constant compaction load of 80 kN in order to focus only on the effect of lubricant content and the

mixing time.

The generated green compacts are tested for their microstructure and mechanical

properties in order to evaluate the influences of the parameter applied. Scanning Electron

Microscopy (Quanta 400, Serial Number 4022 262 32071, Manufactured by Fei Company) is used

for observing only the iron powder particle arrangements. Mechanical properties are analyzed

through density measurement, Rockwell hardness (HRF) test, tensile test, and electrical resistivity

test. Fill density was obtained by dividing the mass of the powder required to fully fill the die

cavity by the volume of the die cavity. Green density was measured by dividing the mass of the

green compact by the volume of the green compact. The hardness of the compact was measured

by digital Rockwell hardness tester. Steel ball of diameter 1/16 inch was used as indenter with the

load of 60 kg. The compact was placed on the platform which was moved upward by the rotating

wheel. Upon contact with the ball indenter on the top of the sample, a reading appeared at the

screen indicating the hardness of the sample in HRF.

The bending strength of each compact was measured through the Instron Universal Testing

machine. The compact was hold tightly at both ends when a transverse force is applied at the

middle of the compact. All data such as elongation, maximum stress and maximum strength were

collected through the data acquisition system in the computer until the compact fractured or broke

down. Electrical resistivity is obtained by measuring the electrical resistance of the green compact

6
ACCEPTED MANUSCRIPT

by a multimeter. The red and black pointers of the multimeter were placed at both ends of the

green compact and the reading was recorded once it was stable.

3. Results and Discussions

Figure 2 presents the micrograph of green compacts formed at different load and temperature. It

shows that lower compaction pressure gives poor particle adhesion, and even some macropores are

found in the sample formed by 105 kN at 30°C forming temperature (Figure 2b, d). Calcination

caused fusion among particles, and it can be found in the sample formed by 125 kN at

170°C (Figure 2a) which showed almost perfect particle fusion by leaving no particle boundaries.

An increasing trend of relative density with respect to compaction pressure is obtained for

the compaction at every forming temperature (Figure 3). At elevated temperature, the green

compact requires less pressure to achieve higher relative density. This phenomenon shows that

temperature is an important factor in reducing applied pressure in obtaining higher relative density.

This finding strengthens the established theory which stated that the deformation of metals can be

looked upon as a thermally activated process where both the elastic and inelastic deformations are

affected [25].

Figure 4 shows that higher weight percent of zinc stearate in powder mass contributes to

lower load needed for the same deformation. Zinc stearate is fused during the compaction

pressure, penetrated between particles of the iron powder by capillary force, and is uniformly

dispersed in the particles so as to decrease the contact resistance between the particles, accelerating

the rearrangement of the particles, and accelerating the increase in density of the compact [26].

Density is the ratio between the mass of a compact and its volume whereas relative density is the

ratio between the instantaneous density and the true density of a compact produced from the same

material. Higher relative density means the lower volume. Therefore, higher weight percent of zinc

stearate requires lower load in obtaining the maximum displacement means lower volume hence

increased the density of the green compacts.

7
ACCEPTED MANUSCRIPT

It is also found that a longer mixing time improved the density of the green compact. The

mixing time of 60 minutes has been found to be more effective compared to 30 minutes (Figure 4).

The reason is that longer mixing time gives more space and chances to the lubricant to fully

assimilate with the powder. The more uniform zinc stearate dispersion inside powder mass, the

more iron particles have the chance to slide between each other during compaction, thus increases

the density of green compact. As the function of lubricant to the powder mass is to increase the

flowability of iron particles and promoting free spaces to be filled, it is important that the lubricant

to be forced out from the powder mass [23]. In order to achieve this purpose, longer mixing time

has been tested in this research referring to the studies reported in [31-35].

Figure 4 also illustrates the effect of mixing time to the densification of powder mass

compacted at different forming temperature. The increase of density as a consequence of different

mixing time during the preparation of powder mass is due to the tendency of flowing granular

material mixture to segregate [31]. Because of the difference in density and particle size,

segregation occurs [32] where lower density particles, in this case, zinc stearate concentrates at the

periphery of the powder mass [31]. During die filling, assuming all the powder particles at its final

mixing position, significant amount of zinc state is located at the inter phase between powder mass

and die wall. Therefore, this amount of zinc state contributes to the lower friction between die wall

and the powder mass. This shows that, study on mixing time has expanded the function of zinc

stearate from admixed method to die wall lubrication.

Figures 5 and 6 show the hardness variation of green compacts formed at different forming

temperatures and lubricant contents. It clearly shows that 0.6wt% zinc stearate produced harder

green compacts both at 30 and 60 mixing time. The graphs also show that green compacts

generated at higher temperature are harder than formed at lower temperature. The improvement of

hardness is due to the increase in density obtained by warm compaction. As the temperature

increases, the particles are brought into close proximity and are moved relative to each other, thus

8
ACCEPTED MANUSCRIPT

promoting adhesion between them. Hence, the bond strength of particles inside the green compact

increases [27]. However, at elevated forming temperature i.e., 180°C, the hardness decreases. This

is due to the function of lubricant that might exceed its effective limit [28].

The study also found that the best formulation is 0.6wt% lubricant and 30 minutes mixing

time in terms of hardness and relative density (Figures 5 and 6). Due to higher amount of zinc

stearate, less time needed to distribute the lubricant uniformly throughout the iron powder. It can

be observed in Figures 7 and 8 that improved hardness and tensile strength is obtained by suggested

0.6wt% lubricant content and 30 minutes mixing time compared to green compacts without

lubricant. Although the production time increases, the mechanical properties of produced green

compacts are greatly improved. Based on the above results (Figure 4 - Figure 6), improved

properties of green compacts, i.e., higher relative density and hardness can be obtained either by

mixing 0.6wt% zinc stearate for 30 minutes or 0.3wt% of zinc stearate for 60 minutes.

It is also evident from Figure 5 and 6 that shorter mixing time produced harder green

compacts. This can be explained referring to the study conducted by Sommer [33] on the variance

of ideal disorder of a multi-component mixture where the research suggested that mixing efficiency

is time dependant and there is a limit where the mixing process is completed. Further mixing time

does not improve the mixing quality. This may be a reason for no further increase in hardness as

longer mixing time applied. This argument is also in line with the results reported in [36] where 60

minutes mixing time is found to be better in generating T-shape green compacts from iron powder.

The variation of electrical resistivity with respect to forming temperature at different

lubricant contents are presented in Figures 9 and 10. The decreasing trend of resistance due to

increase in forming temperature is observed. This phenomenon implies that electrical conductivity

is increased across the green compact. As the mixing time increases, more uniform zinc stearate is

distributed, hence decreases the chance of zinc stearate to accumulate at certain section that can

reduce the electrical conductivity. The formulation of 0.3wt% lubricant content and 30 minutes

9
ACCEPTED MANUSCRIPT

mixing time is found to be good formulation compared to 0.6wt% lubricant content and 60 minutes

mixing time. Figure 11 shows that the addition of lubricant to the iron powder mass improves the

electrical conductivity of the green compacts. Although pure iron powder should conduct electrical

current better, it would not achieve the desired higher density at same amount of pressure compared

to powder mass with lubricant [28]. Increase in density means increase in particle packing order

and good particle rearrangement as discussed in Figures 2 and 3. This leads to better electrical

conductivity in the present of lubricant. However, higher lubricant content may give adverse effect

due to pore generation, mainly at elevated temperature during sintering at later stage.

4. Conclusions

The study concludes that compaction at higher pressure as well as elevated temperature can

improve microstructure and mechanical properties of green compacts. No defects are observed for

the sample formed by 125 kN compaction load at 170˚C. The suitable lubrication formulation also

improves the mechanical properties of green compact where mixing time can minimize the adverse

effect that usually found in higher weight percent of lubricant content. It also can be highlighted

that admixed lubrication method is essential in powder compaction process.

Acknowledgment

The authors want to thank Ministry of Science and Technology and Innovation (MOSTI) Malaysia

for funding this research project under 03-02-03-SF0146. Thanks to Universiti Tenaga Nasional for

the permission given to the authors to conduct the necessary testing.

References

[1] D. Whittaker, Powder metallurgy applications in the automotive industry, Proceeding

World Conference on Powder Metallurgy, vol. 90, pp. 109-116, 1990.

[2] A.K. Ariffin, A. Jumahat and M.M. Rahman, The simulation of die movement in designing

P/M parts, National Design Seminar 2001, Johor Bahru - Malaysia, 15th January, 2001.

10
ACCEPTED MANUSCRIPT

[3] A.K. Ariffin, M.M. Rahman and A. Jumahat, An Experimental investigation of warm

powder compaction process, BSME-ASME International Conference on Thermal

Engineering, 31 December 2001 - 2 January 2002, Dhaka, 2002.

[4] D.C. Zenger and Chai, Common causes of cracks in P/M compacts, Int. Journal of Powder

Metallurgy, vol. 34(4), pp. 35-52, 1998.

[5] P. Seung-Jun, N.H. Heung, H.O. Kyu and N.L. Dong, Model for compaction of metal

powder, International Journal of Mechanical Sciences, vol. 41, pp. 121-141, 1999.

[6] CyberCut, Pressure sintering, Internet: http://cybercut.berkeley. edu/mas2/html/processes/

sinter_pressure/more.html, 2005.

[7] A.K. Ariffin and M.M. Rahman, Warm metal powder compaction process, Advances in

Materials Processing, vol. 1, pp. 159-195, 2003.

[8] A. Bejarano, M.D. Riera and J.M. Prado, Simulation of the compaction process of a two-

level powder metallurgical part, Journal of Materials Processing Technology, vol. 143-

144, pp. 34-45, 2003.

[9] M.M. Rahman, Themomechanical modelling of powder compaction process, M.Sc. Thesis,

Universiti Kebangsaan Malaysia, 1998.

[10] A.E. Wechsler, Thermal and mechanical properties of evacuated powders, Powder

Technology, vol. 3(1), pp. 163-169, 1969.

[11] M. Gagne, Behaviour of powder mix constituents during cold and warm compactions,

Advances in Powder metallurgy and Particulate materials, vol. 1(3), pp. 19-33,1997.

[12] A.B. Höganäs, Iron and steel powders for sintered components, USA: North American

Höganäs, 1998.

[13] B. Johanse, Experience with warm compaction of densmix powder in the production of

complex parts, Proceedings of 2000 Powder Metallurgy World Congress, 536-539,1998.

11
ACCEPTED MANUSCRIPT

[14] G.F. Bocchini, G. Cricri and R. Esposito, Friction effects in metal powder compacting part

two: experimental result, Advances in Powder and Particulate Materials, vol. 1(2),

pp. 141-151,1996.

[15] M.M. Rahman, Finite element modeling of warm powder compaction process. Ph.D.

Thesis, Universiti Kebangsaan Malaysia, 2001.

[16] H.G. Rutz, A.J. Rawlings and T.M. Cimino, Advanced properties of high density ferrous

powder metallurgy materials, Proceedings International Conference on Powder

Metallurgy and Particulate Materials, pp. 1-22, 1995.

[17] M. Hehenberger and J.E. Crawford, A predictor method for finite element analysis of

sliding friction, Scandinavia Journal of Metallurgy, vol. 12, pp. 285-288, 1983.

[18] T. Tabata, S. Masaki and K. Kamata, Determination of the coefficient of friction between

metal powder and die-wall friction in compaction, Journal of Plasticity, vol. 21(236),

pp. 773-779, 1980.

[19] D.V. Tran, R.W. D.T. Gethin and A.K. Ariffin, Numerical modeling of powder

compaction processes: displacement based finite element method, Powder Metallurgy,

vol. 36(4), pp. 257-266, 1993.

[20] A.K. Ariffin, Powder compaction, finite element modeling and experimental validation,

Ph.D. Thesis, University of Wales Swansea, 1995.

[21] A. Jumahat, Analisis kelakuan termo-mekanik bagi proses padatan hangat, M.sc. Thesis,

Universiti Kebangsaan Malaysia, 2001.

[22] A.R. Khoei and R.W. Lewis, Finite element simulation for dinamyc large elastoplastic

deformation in metal powder forming. 1999. Internet: http://www.swan.ac.uk/mecheng/

staff/arkhoei/powder. html, 1999.

[23] S. Deepak and A. Diran, Control strategy for de-lubrication of P/M compacts,

International Journal of Powder Metallurgy, vol. 38(3), pp. 71-79, 2002.

12
ACCEPTED MANUSCRIPT

[24] A.F. Zhornyak and V.E. Oliker, Pressing behavior of atomized iron powders III: Effect of

zinc stearate on compressibility and particle cohesion, Powder Metallurgy and Metal

Ceramics, vol. 20(5), pp. 323-328, 1981

[25] A.B. Höganäs, Handbook for warm compaction, Sweden: Höganäs AB, 1998.

[26] H. Vidarsson, H. Lubricant for metallurgical powder composition, Internet:

http://www.freepatentsonline.com/6605251.html, 1997.

[27] G.T. Murray, Introduction to engineering material; behaviour, properties and selection,

New York: Marcel Dekker Inc., 1993.

[28] S.S.M. Nor, M.M. Rahman, F. Tarlochan, B. Shahida and A.K. Ariffin, The effect of

lubrication in reducing net friction in warm powder compaction process, Journal of

Materials Processing Technology, vol. 207(1-3), pp. 118-124, 2008.

[29] E. Pavier & P. Doremus, Triaxial characterisation of iron powder behaviour, Powder

Metallurgy, 42(4): 345-352, 1999.

[30] O. Coube, A.C.F. Cocks & C.-Y. Wu, Experimental and numerical study of die filling,

powder transfer and die compaction, Powder Metallurgy, 48: 68-76, 2005.

[31] D.V. Khakhar, A.V. Orpe and S.K. Hajra, Segregation of granular materials in rotating

cylinders, Physica A, vol. 318, pp. 129-136, 2003.

[32] B. Daumann and H. Nirschl, Assessment of the mixing efficiency of solid mixtures by

means of image analysis, Powder Technology, vol. 182, pp. 415-423, 2008.

[33] K. Sommer, Sampling of Powders and Bulk Materials, New Yorko: Springer Verlag,1986.

[34] J.J. McCarthy, D.V. Khakhar and J.M. Ottino, Computational studies of granular mixing

Powder Technology, vol. 109, pp. 72-82, 2000.

[35] S.B. Savage and C.K.K. Lun, Particle size segregation in inclined chute flow of dry

cohensionless granular solids, Journal of Fluid Mechanics, vol. 189, pp.311-335, 1988.

13
ACCEPTED MANUSCRIPT

[36] M.M. Rahman and S.S.M. Nor, An experimental investigation of metal powder

compaction at elevated temperature, Mechanics of Materials, vol. 41, pp. 553-560, 2009.

14
ACCEPTED MANUSCRIPT

Figure Captions

Figure 1. T-15 Compaction Machine with Heating System

Figure 2. SEM micrographs of green compacts of iron powder formed at different compaction

load and temperature (a) 125 kN at 170°C, (b) 105 kN at 170°C, (c) 125 kN at 30°C,

(d) 105 kN at 30°C

Figure 3. Relative density variation at different forming temperature

Figure 4. Relative density variation at different forming temperature and lubricant contents

Figure 5. Hardness of green compacts for 30 minutes mixing time

Figure 6. Hardness of green compact for 60 minutes mixing time

Figure 7. The effect of lubricant on hardness of green compact

Figure 8. The effect of lubricant formulation to the tensile strength of green compact

Figure 9. Electrical resistivity of green compact (30 minutes mixing time)

Figure 10. Electrical resistivity of green compact (60 minutes mixing time)

Figure 11. The effect of lubricant to the electrical resistivity of green compact

15
Figure 1
ACCEPTED MANUSCRIPT
Figure 2
ACCEPTED MANUSCRIPT

(a)

(b)

(c)
ACCEPTED MANUSCRIPT

(d)
Figure 3
ACCEPTED MANUSCRIPT

0.85

0.84

0.83
Relative Density, rr

0.82

0.81
30ºC
0.80 170ºC

0.79

0.78
100 105 110 115 120 125 130
Load (kN)
Figure 4
ACCEPTED MANUSCRIPT

0.77
0.3 wt% - 30 minutes
0.76
0.3 wt% - 60 minutes
0.75
0.6 wt% - 30 minutes
0.74 0.6 wt% - 60 minutes
Relative Density, rr

0.73

0.72

0.71

0.70

0.69

0.68
0 50 100 150 200
Temperature (ºC)
Figure 5
ACCEPTED MANUSCRIPT

106

104
Rockwell Hardness (HRF)

102

100

98

96
0.3 wt%
94 0.6 wt%

92

90
0 50 100 150 200
Temperature (ºC)
Figure 6
ACCEPTED MANUSCRIPT

105

100
Rockwell Hardness (HRF)

95

90
0.3 wt%
0.6 wt%
85

80
0 50 100 150 200
Temperature (ºC)
Figure 7
ACCEPTED MANUSCRIPT

120

100
Rockwell Hardness (HRF)

80

60

40 0.6 wt% - 30 minute


without lubricant

20

0
0 50 100 150 200
Temperature (ºC)
Figure 8
ACCEPTED MANUSCRIPT

1.20

1.00
Tensile strength (GPa)

0.80

0.60

0.40 0.6 wt% - 30 minute


without lubrcant
0.20

0.00
0 50 100 150 200
Temperature (ºC)
Figure 9
ACCEPTED MANUSCRIPT

0.14

0.12

0.1
Resistivity (Ω.m)

0.08

0.06 30 minutes mixing

0.3 wt%
0.04 0.6 wt%

0.02

0
0 50 100 150 200

Temperature (ºC)
Figure 10
ACCEPTED MANUSCRIPT

0.4

0.35

0.3
Resistivity (Ω.m)

0.25

0.2
60 minutes mixing time
0.15 0.3 wt%
0.6 wt%
0.1

0.05

0
0 50 100 150 200

Temperature (ºC)
Figure 11
ACCEPTED MANUSCRIPT

0.045

0.040

0.035

0.030
Resistivity (Ω.m)

0.025

0.020

0.015

0.010

0.005 0.3 wt% - 30 minute without lubricant

0.000
0 50 100 150 200

Temperature (ºC)

Potrebbero piacerti anche