Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Powder Technology 155 (2005) 101 – 107

www.elsevier.com/locate/powtec

Friction and adhesion of single spray-dried granules containing a


hygroscopic polymeric binder
Karin M. Andersson, Lennart Bergström *
Department of Physical, Inorganic and Structural Chemistry, Arrhenius Laboratory, Stockholm University, SE-106 91 Stockholm, Sweden

Received 24 February 2004; received in revised form 17 February 2005; accepted 18 May 2005
Available online 1 July 2005

Abstract

The atomic force microscope has been used to study the friction and adhesion of single spray-dried granules containing a mixture of fine
tungsten carbide and cobalt powders and various amounts of a polymeric binder, polyethylene glycol (PEG). The pull-off and friction forces
between two single granules (representing intergranular friction) and between a granule and a hard metal substrate (representing die – wall
friction) have been determined as a function of relative humidity. It was found that the granule – wall friction increased with binder content
and relative humidity. The small friction force at the lowest addition of PEG was related to a small contact area due to the high surface
roughness of the granules. The substantial increase in the friction coefficient at PEG-addition >1 wt.% was related to the plasticity of the
binder-rich granule surface where an increase in binder content or relative humidity increases the deformability. The granule – granule friction
and adhesion was independent of the relative humidity and substantially lower than the granule – wall friction at all PEG contents, which has
important implications for the handling of granular matter.
D 2005 Elsevier B.V. All rights reserved.

Keywords: Friction; Granule; PEG; Hard metal; Pressing; AFM

1. Introduction fracture and deformation will have a negative effect on


flowability. The deformation behaviour of the granules is
Dry pressing and cold isostatic pressing are probably strongly affected by the physical properties of the
the most important forming techniques for industrial polymeric binder. Several studies have shown that
production of ceramic and hard metal materials [1,2]. densification of granulated powders is enhanced when
Green bodies are formed by pressing free flowing the glass transition temperature, T g, of the binder is
granules in a die. The granules are formed from a lowered below the pressing temperature [8,11,13]. The
suspension of a fine powder and a polymeric binder binder deforms plastically at temperatures above the T g
using a granulation technique, e.g. spray-drying or freeze and the intergranular pores are thus more easily elimi-
granulation [3,4]. It is well known that the quality of the nated. Binders commonly used for pressing of fine
pressed body strongly depends on the properties of the ceramic and hard metal powders are poly(vinyl alcohol)
granules [5 – 13]. If the granules are not completely (PVA) and poly(ethylene glycol) (PEG). The T g for pure
broken-down during pressing, the remnant structure may PVA is normally above room temperature. However, PVA
induce large defects during sintering. Hence, the granules can be plasticised, i.e. T g can be reduced, by the addition
should not be too hard. However, soft granules may cause of low molecular weight polymers e.g. poly(ethylene
problems with handling and mould filling since granule glycol) (PEG). PVA is also plasticised by the absorption
of water, which means that the pressing performance can
vary drastically with relative humidity [9]. In contrast,
* Corresponding author. Tel.: +46 8 16 23 681; fax: +46 8 15 21 87. PEG has a T g below 0 -C and thus deforms viscoelas-
E-mail address: lennartb@inorg.su.se (L. Bergström). tically at room temperature. Hence, sufficient granule
0032-5910/$ - see front matter D 2005 Elsevier B.V. All rights reserved.
doi:10.1016/j.powtec.2005.05.055
102 K.M. Andersson, L. Bergström / Powder Technology 155 (2005) 101 – 107

deformation may be achieved without the addition of a 2. Experimental


plasticiser if PEG is used as a binder [12,13]. However,
the hygroscopic nature of PEG and the reduction in 2.1. Materials
hardness with water uptake also makes this binder
sensitive to changes in relative humidity [13]. The The granulated powder, provided by CERMeP (Greno-
water-soluble polymeric binder can migrate to the granule ble, France), was produced from an aqueous powder
surface during spray-drying, which enhances the impor- suspension of 12 vol.% solids loading using a laboratory
tance of the binder properties controlling the friction spray-drier (Anhydro, Denmark) with a capacity of evapo-
between granules [14,15]. There have been some attempts rating 7.5 kg of water per hour. The hard metal powder is a
to relate the frictional behaviour of granular assemblies to mixture of 89.5% fine grain WC (grain size 0.5 – 1 Am after
the filling, rearrangement and pressing properties [5– 10]. sintering), 10% Co, and 0.5% Cr3C2 by weight. The binder
Briscoe and Rough found that a poor rearrangement of the was polyethylene glycol, PEG, with a molecular weight
granules at the early stages of pressing, which was related MW = 3400 (BASF, Ludwigshafen, Germany). Four differ-
to a large die – wall friction, induces stresses in the ent concentrations of PEG: 0.5, 1, 2, and 3 wt.% (weight %
compacted part that may lead to cracking upon ejection of the solids content of the suspension) were added to the
[6]. Recent work by Balasubramanian et al. showed that a powder suspension prior to spray-drying.
reduction in the intergranular porosity could be achieved We have used a polished hard metal pressing tool punch
by the addition of an internal lubricant that results in an provided by AB Sandvik Coromant (Stockholm, Sweden) as
enhanced granule rearrangement [5]. the substrate for the granule – wall measurements. The
Despite the importance of friction and adhesion in material consisted of 94.3 wt.% WC (grain size of 1 Am),
controlling the rearrangement of the undeformed granules 5 wt.% Co, and 0.7 wt.% Cr3C2. The RMS surface
at low applied pressures, there is a lack of fundamental roughness of the polished surface is 3.4 nm as evaluated
studies on the level of the single granule. Microscopic from a 10  10 Am AFM topographic image. Substrates for
friction and adhesion measurements were greatly facili- the granule –granule measurements were prepared by gluing
tated by the introduction of Friction Force Microscopy WC –Co granules onto a flat metal disc. An example of a
(FFM) by Mate et al. [16]. They used the atomic force substrate with granules of a typical size around 150 Am is
microscope (AFM) for friction force measurements shown in Fig. 1.
between a tungsten tip and a graphite surface. Since then The surface roughness of the spray-dried granules
the progress in detector and cantilever calibration techni- containing various amounts of PEG binder was measured
ques has enabled quantitative friction measurements on a using a non-contact profilometer (ZygoLOT New View
nanoscopic length scale. Implementing the colloidal probe 5000, ZygoLOT GmbH, Germany). The interference pattern
technique [17], where a spherical particle of a size ranging of two light beams reflected from the granule and a
from a few micrometers up to the order of 50 Am of an reference surface, respectively, is used to determine the
arbitrary material is attached to the cantilever, has topography and surface roughness parameters [26]. The
facilitated the investigation of frictional properties of a surface roughness is determined from the average of at least
wide range of materials [18 – 25]. Jones showed in a 35 granules.
recent study that FFM measurements can be related to
bulk flow and cohesion measurements, at low consolida-
tion stresses, for a range of model particles and cohesive
granular materials [22]. Meurk et al. showed in one of the
few previous FFM studies on granules that the relative
humidity has a strong effect on the adhesion and the
friction coefficient [24].
In this study, the FFM technique was used to
investigate the adhesional and frictional response of
spray-dried WC –Co granules containing various amounts
of PEG as a binder. Both the granule– wall and granule–
granule friction have been measured. The single granule
FFM measurements clearly illustrate the importance of
humidity and binder concentration on the granule– wall
friction and adhesion. It was shown that the intergranular
friction coefficient was lower than the external friction
coefficient (between a granule and a flat surface). The
implications of the single granule data on the filling,
rearrangement and pressing properties of granule assem- Fig. 1. SEM image of a granule substrate used for granule – granule friction
blies were discussed. measurements. The length of the size bar is 100 Am.
K.M. Andersson, L. Bergström / Powder Technology 155 (2005) 101 – 107 103

relatively small; in the range of 20– 30 Am. Fig. 2 shows a


granule attached to a tipless cantilever. The granules were
dried in a vacuum at 30 -C for at least 12 h and then stored
in a dry atmosphere prior to the measurements. The
measurements were carried out at three different relative
humidities (RH): 5%, 42%, and 70%, usually starting at the
lowest value and increasing the relative humidity for the
consecutive experiments. The relative humidity of the
ambient air during the friction and adhesion measurements
was controlled by using an atmospheric hood made in-
house, covering the AFM. The humidity was varied by
placing open containers of a drying agent (silica gel for 5%
RH) or saturated solutions of different salts (K2CO3 and
NaCl for 42% and 70% RH, respectively) inside the hood.
The system was left to equilibrate for 45 min for each RH
values prior to the measurements. All measurements were
carried out at 20 -C.
Fig. 2. SEM image of a spray-dried granule mounted on an AFM cantilever.
The length of the size bar is 10 Am.
3. Results and discussion
2.2. Methods
Fig. 3 shows the friction force as a function of load
An atomic force microscope (MultiMode SPM, Nano- between a spray-dried WC – Co granule containing 0.5 wt.%
scope IIIA, Digital Instruments, USA) was used for friction PEG and a hard metal surface. The friction force measure-
force measurements. Quantitative measurements of the ments were carried out by sliding the granule attached to a
friction force are attainable by recording the torsional tipless AFM cantilever over a polished hard metal surface.
bending of the cantilever having a spray-dried WC– Co The sliding distance was 5 Am, which was sufficiently long
granule attached to it. The friction force at a particular to ensure that the frictional response represent an average of
applied load is determined from the friction loop, which is the granule surface; several WC and Co-rich regions were
the lateral signal as a function of sliding distance when the probed in a single measurement. The pull-off force could be
cantilever slides sidewards in both directions. The load is obtained from the minimum in a retraction normal force –
increased and then decreased in steps to obtain a set of distance curve [29] (Fig. 4). The estimated adhesion force,
loading– unloading curves [16]. L a, is indicated in the friction – load curves in Fig. 3 as the
The normal surface forces between the granule and the value of the applied load at zero friction force.
flat surface were evaluated by compressing and retracting
the surfaces. The normal deflection of the cantilever with
the attached granule was registered as a function of the
movement of the piezoelectric scanner in the Z direction.
The raw data was transformed into true force curves as a
function of surface separation. The maximum compressive
load of 1500 – 2250 nN is in the same order as the maximum
applied load during the friction measurements.
The cantilevers used in this study were tipless, rectan-
gular aluminium-coated silicon beams (Silicon MDT,
Moscow, Russia). The normal spring constants of the
cantilevers were determined by attaching a series of
tungsten spheres of varying mass and measuring the
corresponding shift in frequency [27]. The spring constants
of the cantilevers used here were typically 10– 15 N/m. The
detector response and torsional bending of the cantilevers
were determined using the method by Feiler et al. [28].
Single granules of different PEG-content were glued to
Fig. 3. Loading – unloading friction measurements between a spray-dried
the apex of the cantilevers with a two-component epoxy
WC – Co granule containing 0.5 wt.% PEG and a flat hard metal surface at
resin (Araldit, Casco Akzo Nobel, Sweden) using a micro- relative humidity of 5% (g), 42% (>), and 70% (q). Filled symbols
manipulator 5171 (Eppendorf, Germany). The size of the represent the first loading – unloading cycle at a specified relative humidity.
granules used for the direct friction force measurements is Solid lines are fits to Eq. (1).
104 K.M. Andersson, L. Bergström / Powder Technology 155 (2005) 101 – 107

to the friction factor, k, only when n = 1 [34]. When n < 1, l


was evaluated from the derivative of the friction force, F f,
with respect to load, L, according to
dFf
l¼ ¼ nk ð L þ La Þn1 ð2Þ
dL
at the maximum load for the friction experiments in this
study, which was 2000 nN. The values of the constants may
differ between loading and unloading in the case of
hysteresis.
The influence of both binder content and relative
humidity on the friction behaviour is shown in Fig. 5. We
find that the friction coefficient increases with both binder
content and RH for granules containing between 1 and 3
wt.% PEG, which is consistent with a friction force
Fig. 4. Force – distance curves between a spray-dried WC – Co granule behaviour being controlled by the properties of the binder-
containing 0.5 wt.% PEG and a flat hard metal surface at relative humidity rich granule surface. The increase in the friction coefficient
˝ &
of 5% (g, ), 42% (>, ), and 70% (q,r). Empty (filled) symbols
represent the withdrawal (approach) curves.
with binder content (e.g. from 0.34 for granules containing 1
wt.% PEG (Fig. 5a) up to a value of 0.39 for granules
containing 3 wt.% PEG (Fig. 5c)), suggests that the ductility
The friction between rough surfaces in sliding contact is and deformability of the surface increases with increasing
generally well described by Amonton’s law with a direct binder content. The contact area increases more with applied
proportionality of the friction force to the applied load [30]. load at high binder additions compared to low binder
However, the presence of a soft material e.g. a polymer on additions because a binder-rich granule surface is less stiff
the surfaces may infer a boundary-type lubrication where and deforms easier than a granule surface containing a low
the interfacial rheology of the polymer will control the amount of polymeric binder.
friction. It is well known that spray-drying often results in a The increase in the friction coefficients with increasing
binder-rich granule surface due to an evaporation-driven RH for the granules containing at least 1 wt.% PEG can be
migration of soluble polymers [14,15]. The friction behav- attributed to the softening of the binder with water uptake
iour of polymeric surfaces can generally be described by the [13], which should result in a more deformable granule
adhesion model of friction [30], which relates the frictional surface. It is interesting to note that the friction coefficient
force to the interfacial shear strength and the real area of for the granules containing 0.5 wt.% PEG (Fig. 3) is quite
contact. The shear strength is a function of the contact low (0.20) and do not vary with RH. The granule surface
pressure, and thus the friction force can be expressed as a is not saturated with polymeric binder at this low binder
function of both applied load and contact area. Combining content, which is expected to result in a relatively hard
this with the load dependence of the contact area of a surface where the polymer binder has a minor influence on
spherical particle from contact mechanics theories [31] one e.g. the plasticity and deformability. Hence, the friction
arrives at a simple expression of the form [32,33] force is mainly governed by the surface morphology, i.e.,
surface roughness. A rough granule surface yields a
Ff ¼ k ð L þ La Þn ð1Þ
smaller contact area than a smooth surface and results in
to describe the frictional properties of the granules, where k lower friction forces. Hence, the significantly lower
is the friction factor, L is the applied load, L a is the adhesion friction coefficient for the granules containing only 0.5
force and n the load index. The load index can vary between wt.% PEG compared to granules with higher binder
2/3 and 1; the latter case is equivalent to plastic deformation content can be attributed to the harder and less deformable
or a multiple asperity contact in accordance with Amonton’s granule surface together with the significantly larger
law, whereas n = 2/3 represents an elastic single asperity surface roughness of the granules containing 0.5 wt.%
contact. Previous work has shown that this simple model PEG is higher compared to the granules containing a large
can describe the interfacial sliding friction for polymer amount of binder (Table 1).
saturated surfaces without including contributions from a It is interesting to note that the pull-off force does not
ploughing component where large volumes of deformation vary much with PEG content (Fig. 5), which suggests that
are involved [34]. the adhesion energy is relatively independent of the binder
Fitting the data at 5% RH in Fig. 3 to Eq. (1) results in a content of the granule. This is another indication that the
nearly direct proportionality between friction force and composition of the granule surfaces is dominated by PEG
applied load; hence the value of the load index n is close to due to the migration during spray-drying. However, the
1. This indicates a multiple asperity contact in agreement thickness of the binder rich surface layer may differ and may
with Amonton’s law. The friction coefficient, l, is identical thus result in various degrees of deformation at high loads.
K.M. Andersson, L. Bergström / Powder Technology 155 (2005) 101 – 107 105

Table 1
Surface roughness of WC – Co granules with various amounts of PEG
binder
PEG (wt.%) RMS (nm)
0.5 530
1 433
2 354
3 337

The observed increase of the friction and adhesion of


single granules to a hard metal substrate with increasing
binder content and relative humidity can explain previous
findings on a decrease in uniformity of parts that have been
compacted at high humidity [7]. It is well known that the
die –wall friction is the main cause of density gradients in a
pressed body.
It is also possible to relate the deviation between the
loading and unloading curves, the adhesion hysteresis [35],
to the granule properties. It was found that the granules
containing 1 wt.% PEG display adhesion hysteresis at 70%
RH but not at 5 and 42% RH (Fig. 5a). Adhesion hysteresis
appears at all relative humidities for the granules containing
2 and 3 wt.% PEG when the loading– unloading friction
cycle is run the first time (Fig. 5b –c). Fundamentally, the
adhesion hysteresis can be related to a non-elastic deforma-
tion behaviour of surfaces in contact. One example is the
first loading curve of the 3 wt.% PEG granule at 42% RH
(Fig. 5c) where the friction force increases abruptly at an
applied load å 800 nN. When the applied load is decreased,
the contact area remains high due to the plastic deformation
of the granule, which results in a large friction force during
unloading. The load index, n, is always smaller for the
unloading curve compared to the loading curves. After the
first loading – unloading measurement, the hysteresis typi-
cally diminishes and the curves overlap (see Fig. 5c).
However, it was found that the hysteresis reappeared in the
loading– unloading curves when the relative humidity was
increased and a sufficient amount of time had passed. We
speculate that the relaxation time of the polymer chains was
reduced by the absorption of water, enabling the binder-rich
surface to deform plastically again. This time-dependent
viscoelastic behaviour of the binder may be related to the
spring-back problems that often hamper the pressing
performance [9].
The hysteresis observed in the first set of loading–
unloading curves between the WC– Co granule containing
0.5 wt.% PEG and the hard metal surface (Fig. 3) has a
different origin. The scatter in the friction – load curve
suggests that the contact area fluctuates. This can be related
to the brittle nature of the surface of granules containing a
small amount of polymeric binder. An addition of only 0.5
wt.% of PEG is simply not sufficient to bind all the primary
Fig. 5. Loading – unloading friction measurements between a spray-dried
WC and Co particles together to form smooth spherical
WC – Co granule containing (a) 1 wt.% PEG, (b) 2 wt.% PEG, (c) 3 wt.%
PEG and a flat hard metal surface at a relative humidity of 5% (g), 42% granules. This is corroborated by the larger surface rough-
(>), and 70% (q). Filled symbols represent the first loading – unloading ness (Table 1) of the granules with low PEG content. The
cycle at a certain relative humidity. Solid lines are fits to Eq. (1). poorly bound particles at the granule surface are easily
106 K.M. Andersson, L. Bergström / Powder Technology 155 (2005) 101 – 107

eroded away when sliding. The scatter in the friction – load


curves was significantly reduced when these particles had
been removed, see e.g. the curves at 5% RH (Fig. 3).
The load index, n, of the unloading curves for the two
granules containing the highest amounts of PEG, 2 and 3
wt.% (Fig. 5b and c) was reduced from a value close to 1 to
2/3 with increasing RH and the number of consecutive
measurements. For the first runs at 5% RH to the last run at
70% RH there was a steady decrease of n from 0.98 to 0.72
and from 0.98 to 0.67 for the granules containing 2 and 3
wt.% PEG, respectively. This indicates that the asperities at
the granule surfaces deformed plastically at the first contact
rendering a smoother, more elastic contact for the consec-
utive friction measurements. It is interesting to note that the
humidity dependence of the friction behaviour is to some
extent reversible (not shown). When the RH was decreased
from 70% to 42%, the friction coefficient and static friction
decreased so that the friction curve overlapped with the
previous curve achieved at 42% RH. The same trend could
be seen in the normal force – distance measurements where
the adhesion forces tended to decrease reversibly when the
RH was decreased. These results indicate that the deforma-
tion of the granule surface is not permanent, which is
probably related to the viscoelastic nature of PEG.
It was found that the friction and adhesion was decreased
further when the RH was lowered to 5%; however, the low
friction coefficient and adhesion force that initially was
measured at 5% RH could not be reached again when the
granules had been subjected to a high RH. This poor
repeatability is probably a result of absorbed moisture that
cannot be removed from the hygroscopic binder at room
temperature. The water absorption was estimated by measur-
ing the change of the resonant frequency of the cantilever
with the attached granule [27]. The mass of a 35 Am granule Fig. 6. Loading – unloading friction measurements between two spray-
dried WC – Co granules containing (a) 0.5 wt.% PEG and (b) 3 wt.% PEG
increased 0.08 ng when the RH was cycled from 5% to 70%
at a relative humidity of 5% (g), 42% (>), and 70% (q). Solid lines are
and 0.03 ng remained when the RH was reduced back to 5%. fits to Eq. (1).
Hence, almost 40% of the water that was taken up at 70% RH
remained in the granule when the RH was reduced back to
5%. This amount of irreversibly absorbed water relates to granules, which makes the granule –granule friction rela-
approximately 1% of the amount of PEG in the granule. tively insensitive to changes in PEG content and RH. It was
Fig. 6a and b show the friction – load curves for a granule found that the load indices were close to 1 irrespective of the
sliding on another granule. The granule –granule system binder content, which suggests that the contact between the
displayed significantly lower values of L a and the friction at granules is dominated by the asperities on the surface. These
zero load compared to the granules sliding on a hard metal results have important implications for die filling and the
substrate (Figs. 3 and 5c). Our finding that the intergranular initial steps of dry pressing. Previous work has shown that
adhesion is small compared to the corresponding granule– the packing of granules during die filling and rearrangement
wall adhesion, has also been noticed by other authors at low applied loads depend on the friction between the
[22,24]. The weak adhesion between the granules suggests granules [5]. These results suggest that the internal pore
that the contact area was small, which can probably be structure is not expected to be critically dependent on
explained by the relatively large surface roughness. The variations in RH and PEG binder content.
small contact area can also explain the very small variation in
adhesion with relative humidity. Moreover, there was only a
small difference in the static friction and L a between granules 4. Conclusions
containing 0.5% and 3% PEG, and the friction coefficient is
almost independent of RH. Hence, the adhesion is of minor The atomic force microscope (AFM) has been used for
importance for the friction response of the two individual direct measurement of friction and adhesion of single spray-
K.M. Andersson, L. Bergström / Powder Technology 155 (2005) 101 – 107 107

dried WC –Co granules containing a hygroscopic polymeric [5] S. Balasubramanian, D.J. Shanefield, D.E. Niesz, J. Am. Ceram. Soc.
binder, PEG. The AFM friction force measurements provide 85 (2002) 134 – 138.
[6] B.J. Briscoe, S.L. Rough, Powder Technol. 99 (1998) 228 – 233.
an accurate and versatile method to evaluate the humidity [7] T.A. Deis, J.J. Lannutti, J. Am. Ceram. Soc. 81 (1998) 1237 – 1247.
dependence of the intergranular and external friction of [8] R.A. DiMilia, J.S. Reed, Ceram. Bull. 62 (1983) 484 – 488.
relevance in powder pressing. [9] J. Nam, W. Li, J.J. Lannutti, Powder Technol. 133 (2003) 23 – 32.
We found that the granule– wall friction was controlled [10] F. Negre, E. Sánchez, in: V.E. Henkes, G.Y. Onoda, W.M. Carty
(Eds.), Science of Whitewares, The American Ceramic Society,
by the binder content and the relative humidity (RH). The
Westerville, OH, 1996, pp. 169 – 181.
friction coefficient increased with increasing RH for [11] C.W. Nies, G.L. Messing, J. Am. Ceram. Soc. 67 (1984) 301 – 304.
granules with a high PEG content. This behaviour was [12] W.J. Walker, J.S. Reed, Ceram. Eng. Sci. Proc. 14 (1993) 58 – 79.
ascribed to an increase in the plasticity of the granule [13] D.W. Whitman, D.I. Cumbers, X.K. Wu, Am. Ceram. Soc. Bull. 74
surface, which was caused by the binder becoming softer by (1995) 76 – 79.
the absorption of water. This behaviour may explain the [14] S. Baklouti, T. Chartier, J.F. Baumard, J. Eur. Ceram. Soc. 18 (1998)
2117 – 2121.
increased tendency of density gradients in powder compacts [15] Y. Zhang, X. Tang, N. Uchida, K. Uematsu, J. Mater. Res. 13 (1998)
with increasing RH. The granule – granule friction and 1881 – 1887.
adhesion was lower than the granule– wall friction at all [16] C.M. Mate, G.M. McClelland, R. Erlandsson, S. Chiang, Phys. Rev.
PEG contents. This was related to the large surface Lett. 59 (1987) 1942 – 1945.
roughness of the granules, which suggested that the contact [17] W.A. Ducker, T.J. Senden, R.M. Pashley, Nature 353 (1991) 239 – 241.
[18] G. Bogdanovic, F. Tiberg, M.W. Rutland, Langmuir 17 (2001)
area was smaller between two granules compared to a 5911 – 5916.
granule and a flat substrate. Hence, the external friction is [19] R.G. Cain, N.W. Page, S. Biggs, Phys. Rev., E 62 (2000) 8369 – 8379.
always significantly larger than the intergranular friction. [20] A. Feiler, I. Larson, P. Jenkins, P. Attard, Langmuir 16 (2000)
Thus, the packing of granules, which is to a large extent 10269 – 10277.
controlled by intergranular friction, is expected to be [21] I.C. Hahn Berg, L. Lindh, T. Arnebrant, Biofouling 20 (2004) 65 – 70.
[22] R. Jones, Granul. Matter 4 (2003) 191 – 204.
relatively insensitive to binder content and RH. The density [23] A. Meurk, I. Larson, L. Bergström, in: N.R. Moody, W.W.
gradients in pressed bodies are related to the external Gerberich, S.P. Baker, N. Burnham (Eds.), Fundamentals of Nano-
(granule – die wall) friction, which suggests that the RH and indentation and Nanotribology, Materials Research Society, Pitts-
binder content needs to be controlled and optimised for burg, 1998, pp. 427 – 432.
reliable production using dry pressing. [24] A. Meurk, J. Yanez, L. Bergström, Powder Technol. 119 (2001)
241 – 249.
[25] S. Zauscher, D.J. Klingenberg, Nord. Pulp Pap. Res. J. 15 (2000)
459 – 468.
Acknowledgements [26] T.R. Thomas, Rough Surfaces, Imperial College Press, London, 1999,
p. 278.
[27] J.P. Cleveland, S. Manne, D. Bocek, P.K. Hansma, Rev. Sci. Instrum.
This work has been performed within the Brinell Centre
64 (1993) 403 – 405.
Inorganic Interfacial Engineering (BRIIE), supported by the [28] A. Feiler, P. Attard, I. Larson, Rev. Sci. Instrum. 71 (2000) 2746 – 2750.
Swedish Agency for Innovation Systems (VINNOVA) and [29] M. Enachescu, R.J.A. van den Oetelaar, R.W. Carpick, D.F. Ogletree,
the following industrial partners: Erasteel Kloster AB, C.F.J. Flipse, M. Salmeron, Phys. Rev. Lett. 81 (1998) 1877 – 1880.
Höganäs AB, AB Sandvik Coromant, Seco Tools AB, and [30] F.P. Bowden, D. Tabor, The Friction and Lubrication of Solids, Oxford
Atlas Copco Secoroc AB. Univ. Press, London, 1950.
[31] E. Meyer, R.M. Overney, K. Dransfeld, T. Gyalog, Nanoscience:
Friction and Rheology on the Nanometer Scale, World Scientific
Publishing Co., Singapore, 1998.
References [32] M.J. Adams, B.J. Briscoe, J.Y.C. Law, P.F. Luckham, D.R. Williams,
Langmuir 17 (2001) 6953 – 6960.
[1] S.J. Lukasiewicz, J.S. Reed, Ceram. Bull. 57 (1978) 798 – 801. [33] B.J. Briscoe, F.R.S. Tabor, in: D.T. Clark, W.J. Feast (Eds.), Polymer
[2] J.S. Reed, Introduction to the Principles of Ceramic Processing, John Surfaces, John Wiley & Sons, New York, 1978, pp. 1 – 24.
Wiley & Sons, New York, 1988. [34] M.J. Adams, B.J. Briscoe, L. Pope, in: B.J. Briscoe, M.J. Adams
[3] E. Carlström, in: R.J. Pugh, L. Bergström (Eds.), Surface and Colloid (Eds.), Tribology in Particulate Technology, Adam Hilger, Bristol,
Chemistry in Advanced Ceramics Processing, Marcel Dekker Inc., 1987, pp. 8 – 22.
New York, 1994, pp. 245 – 278. [35] J.N. Israelachvili, Y.L. Chen, H. Yoshizawa, J. Adhes. Sci. Technol. 8
[4] W.J. Walker, J.S. Reed, J. Am. Ceram. Soc. 82 (1999) 1711 – 1719. (1994) 1231 – 1249.

You might also like