Sie sind auf Seite 1von 7

+ MODEL

ARTICLE IN PRESS

Advances in Colloid and Interface Science xx (2006) xxx – xxx


www.elsevier.com/locate/cis

Nanotechnology in action: Overbased nanodetergents


as lubricant oil additives
L.K. Hudson a , J. Eastoe a,⁎, P.J. Dowding b
a
School of Chemistry, University of Bristol, Bristol, BS8 1TS, United Kingdom
b
Infineum UK Ltd, Milton Hill Business and Technology Centre, Abingdon, Oxon, OX13 6BB, United Kingdom

Abstract

The synthesis and study of oil-soluble metal carbonate colloids are of interest in the area of lubricant additives. These surfactant-stabilised
nanoparticles are important components in marine and automotive engine oils. Recently introduced, environmentally driven legislation has focused on
lowering of gaseous emissions by placing limits on the levels of phosphorous sulphur and ash allowed in engine oil systems. These chemical limits,
coupled with improved engine performance and extended oil drainage intervals, have lead to renewed interest in the production of stable, efficient
nanodetergent systems. To date, this has resulted in modification of existing surfactant structures and development of new generations of surfactants.
This review covers the current state of research in the area of nanodetergents.
© 2006 Elsevier B.V. All rights reserved.

Keywords: Lubricant additives; Lubricants; Overbased; Detergents; Nanoparticles; Non-aqueous colloids

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
2. Nanoparticle preparation from microemulsions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
3. The detergent core . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
4. Surfactants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
5. Solvent . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
6. Characterisation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
6.1. SANS and SAXS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
6.2. Langmuir trough measurement and molecular dynamics simulations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
6.3. Acid neutralisation mechanism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
7. Conclusions and future outlook. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 0

1. Introduction additives in lubricant oils since the 1940's. A major challenge


currently facing potential industrial applications of new nano-
This review describes recent advances in a commercial ap- technologies is the ability to generate well-defined, functionalised
plication of nanoparticles, which have been routinely used as particles on a mega-tonne scale. Lubricant detergents represent an
important case study for commercialisation of nanotechnology.
⁎ Corresponding author. Tel.: +44 117 9289180; fax: +44 117 9250612. Ever since the invention of the internal combustion engine
E-mail address: julian.eastoe@bristol.ac.uk (J. Eastoe). lubricant oils have been an essential component. Over the last
0001-8686/$ - see front matter © 2006 Elsevier B.V. All rights reserved.
doi:10.1016/j.cis.2006.05.003

CIS-00807; No of Pages 7
ARTICLE IN PRESS
2 L.K. Hudson et al. / Advances in Colloid and Interface Science xx (2006) xxx–xxx

genous impurities in the fuel, and oxidative degradation of the


lubricant to form organo-acids. If allowed to build up, these acids
would cause severe corrosion, especially under engine start-up
conditions. These nanoparticles are relatively insensitive to tem-
perature, explaining their effectiveness at high temperatures as
slow-release acid neutralisers [3]. The principal functions per-
formed by detergents in engine oil formulations are: (1) acid
neutralisation, (2) high temperature detergency, (3) oxidation inhi-
bition, and (4) rust prevention. These functions promote engine
cleanliness, fuel efficiency and extended trouble-free operation.
Various surfactants are used to stabilise the CaCO3 particles,
typically long chain carboxylic acids, glycols, alcohols, phenates,
sulphonates, salicylates or phosphonates. Recently calixarene/
stearate surfactants have been developed [4]; these novel systems
address environmental concerns about existing detergents (by
containing no sulphur or phosphorous). The term “overbased”
refers to the fact that the quantity of base incorporated in the
particle cores is greater than that needed to neutralise the acid
surfactant. The neutralising strength of an overbased detergent
Fig. 1. Schematic picture of an overbased sulphonate detergent particle. is measured by its ‘total base number’ (TBN). Where TBN is
Reproduced by permission of Elsevier. defined as the quantity of acid, expressed in terms of the
equivalent number of milligrams of potassium hydroxide that
is required to neutralise all basic constituents present in 1 g of
century, an increasing range of additives have been incorporated overbased detergent.
into lubricating oils to confer chemical stability, improved per-
formance and beneficial physiochemical properties [1]. An 2. Nanoparticle preparation from microemulsions
important class are the so-called overbased additives (detergents),
which are colloidal nanoparticles of calcium carbonate (and cal- The commercial synthesis of colloidally stable metal carbo-
cium hydroxide) stabilised by a surfactant layer. These nanode- nates is known as the “oxide/hydroxide” process; the mechanism
tergents essentially consist of an inorganic core (15–40 mass %) of formation is thought to proceed through a microemulsion route
stabilised by oil-soluble surfactants (20–45 mass %) incorporated [5]. The reaction system consists of an excess of metal hydroxide,
into a lubricating base oil, as depicted schematically in Fig. 1 [2]. appropriate surfactant and a mixture of hydrocarbon and polar
Fuel combustion yields many by-products, including inorganic solvents. The polar solvents are usually methanol and water:
and organic acids formed by oxidation of sulphurous and nitro- water, surfactant and methanol form swollen reverse micelles,

Fig. 2. XANES spectra obtained on overbased sulphonate detergent particles (in grey) and on pure calcite powder (in black). Reproduced by permission of Elsevier.
ARTICLE IN PRESS
L.K. Hudson et al. / Advances in Colloid and Interface Science xx (2006) xxx–xxx 3

Fig. 3. Positive ion TOF-SIMS spectrum of overbased sulphonate detergent particles showing the presence of calcium and calcium hydroxide ions. Reproduced by
permission of Elsevier.

which contain dispersed Ca(OH)2. The surfactants are made ‘in aragonite, and also as a variety of unusual forms produced prin-
situ’ by the reaction of the corresponding organic acid with the cipally by living organisms, [9]. Calcite has a rhombohedral
metal base. Diffusion of gaseous CO2, which is bubbled through crystal structure, whereas aragonite an orthorhombic crystal struc-
the system, generates CaCO3 in the reverse micellar cores, which ture. Heyes and co-workers [10] have modelled CaCO3 nano-
then nucleate at supersaturation [5]. The micellar cores grow particles, and their simulations suggest an essentially amorphous
through coalescence. Literature suggests that a stable system microstructure. They suggest that there is no thermodynamic bene-
requires a certain amount of residual calcium hydroxide, which fit from forming local crystalline order in the cores due to the large
ensures that the inorganic core is amorphous, [5]. If carbonation is surface to volume ratio. Using complimentary techniques such as,
driven to completion, the cores transform into crystalline calcite, X-ray photoelectron spectroscopy (XPS), X-ray adsorption near
which agglomerates to form larger particles and these eventually edge structure (XANES) and time-of-flight secondary ion mass
sediment during the reaction [6]. The products are then filtered spectroscopy (ToF-SIMS) with examples shown in Figs. 2 and 3,
under high pressure and temperature to remove unreacted the authors [11] have been able to prove the existence of Ca(OH)2
inorganic material and agglomerated particles. The initial water- in the reversed micelle core and suggest that it is preferentially
in-oil microemulsion reaction system results in the formation of located as an outer shell in the micelle core. This can be explained
nanometre-sized particles, with a narrow particle size distribution. by residual calcium hydroxide that is not fully carbonated, but still
Although the mechanisms involved in these reactions are not yet is present in the core thus creating an amorphous structure. Further,
fully understood, [5,7,8], Jacquet and co-workers [6] have studied simulations [2,10] have shown that once nanoparticles of CaCO3
the oxide/hydroxide reaction and have drawn some general are formed, the stabilising surfactant molecules are essentially
conclusions on the behaviour of promoters and determination of “locked” in place on the surface, perhaps by strong coulombic
factors controlling the overbasing reaction. forces originating from the inorganic core material.
In 1996 new inorganic colloidal dispersions of calcium, so-
3. The detergent core dium and potassium thiophosphate were produced and sub-
sequently characterised [12,13]. These systems have beneficial
The detergent core is typically 1–10 nm in diameter as de- anti-wear properties.
monstrated by small-angle neutron scattering (SANS) experi-
ments. Many metals have been incorporated into detergents but 4. Surfactants
currently, based on cost/performance, the three most commonly
used are calcium, magnesium and sodium. The detergent CaCO3 As mentioned above, a range of surfactants can be used to
core naturally occurs in many forms, in the bulk as calcite and stabilise CaCO3 nanoparticles. The chain lengths of these

Table 1
Comparison of detergent properties
Property Phenates Sulphonates Salicylates Phosphonates
Total base number 0–300 0–500 0–300 0–80
(TBN) range
Hydrolytic stability Good Moderate Good Moderate
Oxidation stability Very good Poor Very good Moderate
Thermal stability Excellent Excellent Excellent Moderate
Detergency Good Good Excellent Good
Fig. 4. Molecular structures of surfactants commonly employed in overbased
Rust inhibition Low Good Low Good
detergent systems a) alkyl suphonate, b) sulphurised alkyl phenol (SAP) and c)
Antioxidant effect Very good None Very good Good
alkyl salicylate where R1 is a long chain alkyl tail group.
ARTICLE IN PRESS
4 L.K. Hudson et al. / Advances in Colloid and Interface Science xx (2006) xxx–xxx

properties [14]. There are four major classes: sulphonates,


phenates, salicylates and phosphonates shown schematically
in Fig. 4 [10]. Table 1 summarises the detergent properties of
the different surfactants commonly employed in stabilising
CaCO3 nanodetergents, [15]. Sulphonates are the most widely
used detergent additives followed by phenates, salicylates and
phosphonates.
Environmental concerns, primarily the problems of sulphur
emissions and sulphur/phosphorous poisoning of automobile
catalytic converters, have led to the demand for sulphur-free
overbased detergents such as salicylates and alkyl bridged
phenates (of which calixarenes are an example, generic
structures of these are shown in Fig. 5) and combination of
such structures (e.g. salixarenes) [16].

5. Solvent

The solvent system employed in preparing these overbased


detergents is very important. Small changes in the composition of
the solvent mixtures result in very different final products. Jacquet
and co-workers [17] have methodically explored the effects of
varying the solvent level in the preparation of overbased calcium
Fig. 5. Schematics of generic calixarene structures.
sulphonate. It was shown that: (1) a minimum amount of water is
required, (2) a minimum amount of methanol is required to give a
surfactants are generally between C9 and C60. On neutralising, significant TBN; however, an excess of water has a detrimental
the surfactant prevents both the formation of “varnish” on pistons, effect on the rheology, and may even result in an organogel, and
bears polishing of the cylinder and also confers antifriction (3) increasing the amount of xylene increases the rate of

Fig. 6. Measured small-angle neutron scattering curves and analysed core shell form factors for surfactant-stabilised nanodetergent particles in mixtures of d8-toluene
and h8-toluene. a) and b) are for 6.2 nm particles with an 1.8 nm shell, a) in 100% h8-toluene, b) in 23% d8-toluene, 77% h8-toluene, c) and d) are for 10.5 nm particles
with a 2.6 nm shell, c) in 60% d8-toluene, d) in d8-toluene. Reproduced by permission of Springer.
ARTICLE IN PRESS
L.K. Hudson et al. / Advances in Colloid and Interface Science xx (2006) xxx–xxx 5

6.2. Langmuir trough measurement and molecular dynamics


simulations

Various molecular simulations have been performed on de-


tergent systems. Heyes and co-workers [4,23] have shown, by
combined Langmuir trough measurements and molecular dynam-
ics simulations, that different geometries in the detergent particle
are observed when the surfactant chemistry is varied. It was shown
that sulphonate and salicylate stabilised particles were approxi-
mately spherical in shape (Fig. 7 shows an example of modelled
overbased sulphonate detergent); however, phenates and calixa-
rates generated oblates and prolates, respectively (Fig. 8 depicts a
modelled overbased phenate detergent). The addition of a stearate
cosurfactant gave more spherical particles by breaking the natural
regular packing tendency of the surfactants.
The simulations have shown that the internal structure and
arrangement of the ions in the core are sensitive to the surfactant
type, and produce quite specific core shape and internal ionic
Fig. 7. Space filling (calcium ions) and stick (all other molecules) representa- structures. Strong coulombic forces between the ions provide
tions of overbased sulphonate detergent nanoparticles using repulsive alkyl tails.
the driving mechanism for the Ca2+, CO32−, and surfactant
Reproduced by permission of the Royal Society of Chemistry.
moieties to arrange into a reverse micellar structure, with the

carbonation without changing the final TBN. It is known that the


methanol also acts as a promoter at 70 °C for the preparation of
CaCO3 nanoparticles [18].

6. Characterisation

Detergents have been characterised by many different methods


over the years including acid neutralisation, SANS and Trans-
mission electron microscopy [11,19–22,25,26].

6.1. SANS and SAXS

Both small-angle neutron scattering (SANS) and small-angle


X-ray scattering (SAXS) are ideal methods to study nanodeter-
gents [19–22]. Markovic and Ottewill [19,20,21] prepared cal-
cium carbonate particles, stabilised by alkylbenzene sulphonate
surfactants in toluene and dodecane. They used SANS to char-
acterise the system and the Guinier approximation and full model
fitting to analyse the data. Using these approaches, they could
independently determine the metal carbonate core diameter and
the thickness of the surfactant layer. The data and analysis in
Fig. 6 provide a firm evidence for the core shell structure proposed
by others. More detail can be found in relevant papers [20,21].
Further work by Ottewill and co-workers [22] investigated mag-
nesium carbonate systems. During the synthesis of these systems,
conversion of magnesium oxide stopped naturally when the
basic salt 3MgCO3·Mg(OH)2·3H2O was formed. The stabilising
surfactant used was alkylbenzene sulphonate, and the solvents
were iso-octane and toluene. When using dilute systems, it was
found that a single shell model could be used for iso-octane, but a
double shell model had to be employed to accurately account for
the scattering data with toluene solvent systems. This was at-
Fig. 8. a) Front and b) side projections for space filling (calcium ions) and stick
tributed to a partial solvation of the steric stabilising layer by the (all other molecules) representations of overbased phenate detergent nanopar-
toluene. These detailed analyses were broadly consistent with a ticles using repulsive alkyl tails. Reproduced by permission of the Royal Society
core dimension of 1–10 nm and a steric stabilising layer of 1–5 nm. of Chemistry.
ARTICLE IN PRESS
6 L.K. Hudson et al. / Advances in Colloid and Interface Science xx (2006) xxx–xxx

calcium carbonate in the core and surfactant anions forming an more rapidly. The greater neutralising ability of magnesium
external stabilising shell [3]. The level of penetration of the core detergents was attributed to the fact that neutralisation of inorganic
is greatest for phenate systems and least for sulphonates [23]. acid could occur within the aqueous phase, and that magnesium
Simulations performed by Tobias and Klein [24] modelled carbonate had greater aqueous solubility than calcium carbonate.
sulphonate stabilised nanodetergents, suggesting that the For organic acids (propanoic, cyclohexanoic, heptanoic, nona-
structure was roughly spherical. They also hypothesized that noic and decanoic acids), it was found that neutralisation occurred
due to large dipole moments, it was conceivable that long-range over a longer timescale (up to minutes) than observed for inor-
electrostatic effects could drive strong attractions between free ganic acid neutralisation, and never appeared to reach 100%
polar acid molecules dispersed on the bulk non-polar solvent completion. This was explained by the possibility that neutralised
and the polar micelle core. This suggests a possible mechanism metal carboxylate salt could remain attached to the particle surface
for the enhanced stabilisation in the oil medium. rather than penetrate the inner core. Over time, formation of such a
layer on the periphery of the nanodetergent core would be ex-
6.3. Acid neutralisation mechanism pected to hinder any further reaction between the basic core and
acid molecules. The relative rates of neutralisation between mag-
The main purpose of nanodetergent oil additives is to neutralise nesium and calcium carbonate nanodetergents were reversed for
acids generated by combustion of petrochemical fuels. Various organic acids as compared to inorganic acids, and this is thought to
studies have been made to attempt to elucidate the mechanism of be linked to the relative basicity of the inorganic species, i.e.
neutralisation [7,25]. Robinson et al. [26] have studied the reaction CaCO3 > MgCO3 (and Ca(OH)2 > Mg(OH)2). No significant
of organic and inorganic acids with overbased detergent additives, effects of structure or molecular weight of the sulphonate sur-
in organic media, using Fourier transform infra-red spectroscopy factant were reported with regard to rate of acid neutralisation:
(FTIR) and stopped flow UV/Vis spectroscopy; an example of a again, higher surfactant levels gave rise to faster rates of acid
typical profile generated by a stopped flow experiment is shown in neutralisation. Studies of neutralisation kinetics as a function of
Fig. 9. The systems they studied were amorphous calcium car- acid molecular weight showed that the rate of neutralisation de-
bonate and magnesium carbonate cores stabilised alkylbenzene creased with increased acid molecular weight. This was explained
sulphonate surfactant with TBN of 300. by the bulkier acids being more effective at screening the ino-
For hydrochloric acid, it was noted the acid was neutralised on rganic core, by forming a thicker surface layer, and hence hin-
a timescale of milliseconds to seconds, with the reaction being dering neutralisation. In addition, the relative solubilities of the
both temperature and solvent dependent. Results showed that for a carboxylic acids in the solvent (heptane) will also affect neutral-
‘matched set’ of calcium and magnesium overbased sulphonate isation rate: those with relatively low solubilities (such as pro-
detergents magnesium detergents neutralised inorganic acid more panoic acid) effectively ‘driving’ the acid towards neutralisation.
rapidly than their calcium analogues. The molecular weight and Organic acid neutralisation has also been investigated by Papke
structure of the sulphonate surfactant appeared to have a less [27] using FTIR and overbased calcium sulphonate nanodeter-
marked influence on neutralisation rate. However, detergents gents. In addition, Robinson and co-workers [26] measured –
containing higher levels of surfactant were seen to neutralise acid COOH absorption at 1705 cm− 1 to follow the percentage of acid

Fig. 9. Typical stopped flow trace for the reaction between overbased detergent particles and acid containing water-in-oil microemulsion droplets using methyl range as
the pH indicator. The time (t1/2) taken for the absorbance to decrease by half of the total change (ΔA / 2) is used as a measure of the neutralisation rate. Reproduced by
permission of Elsevier.
ARTICLE IN PRESS
L.K. Hudson et al. / Advances in Colloid and Interface Science xx (2006) xxx–xxx 7

neutralised as a function of time. Papke used the sulphonate band reduce the levels or replace petrochemical solvents with green
at 1065 cm− 1 as a reference, due to the fact that no sulphonate was alternatives.
lost in the precipitate (carboxylate salt). This enabled a direct
determination of the relative amount of solubilized carbonate Acknowledgments
(865 cm− 1) by comparison to the sulphonate level. The measure-
ments were performed with hexanoic acid; a plateau region was LH thanks Impact Faraday (EPSRC) and Infineum UK Ltd for
observed initially during neutralisation, where the amount of the provision of a Ph.D. Scholarship.
carbonate did not change, followed by a linear region corre-
sponding to reaction of carbonate with acid. The author concluded References
that a non-carbonate base was selectively neutralised, which is
residual hydroxide in the core of the nanodetergent. Furthermore, [1] Ullmann's encyclopaedia of industrial chemistry, vol 20, sixth edition, chapter
it must also be accessible and, therefore, is probably located near 6, p 118 Jürgen Braun and Jürgen Omeis.
[2] Bearchell CA, Heyes DM, Moreton DM, Taylor SE. Phys Chem Chem Phys
the surface of the inorganic core, consistent with the simulations of 2001;3:4774.
Heyes and co-workers [2,10] suggesting an outer hydroxide shell. [3] Griffiths JA, Heyes DM. Langmuir 1996;12:2418.
[4] Cunningham ID, Courtois J-P, Danks TN, Heyes DM, Moreton DJ, Taylor
7. Conclusions and future outlook SE. Colloids Surf A Physicochem Eng Asp 2003;229:137.
[5] Bandyopadhyaya R, Kumar R, Gandhi KS. Langmuir 2001;17:1015.
[6] Roman JP, Hoornaert P, Faure D, Biver C, Jacquet F, Martin J-M. J Colloid
Overbased detergents have been an integral element of lu- Interface Sci 1991;144:324.
bricant additive packages for over 50 years. Indeed detergents are [7] Galsworthy J, Hammond S, Hone D. Curr Opin Colloid Sci 2000;5:274.
examples of nanoparticles, which can be reproducibly produced [8] Kandori K, Kon-No K, Kitahara K. J Colloid Interface Sci 1991;144:324.
on an industrial scale. Production of uniform nanoparticles with [9] Bearchell CA, Heyes DM. Mol Simul 2002;28:517.
tailored properties on a large scale is a focus of current research [10] Bearchell CA, Danks TM, Heyes DM, Moreton DJ, Taylor SE. Phys Chem
Chem Phys 2000;2:5197.
interest. The surfactants used to stabilise detergents are generally [11] Cizaire L, Martin JM, Le Mogne Th, Gresser E. Colloids Surf 2004;238:
aromatic in nature, often based on structures commercially avail- 151.
able in petrochemical plants. The need for optimised nanoparticle [12] Delfort B, Chivé A, Barré L. J Coll Colloid Interface Sci 1997;186:300.
detergents is increasing, driven by: [13] Delfort B, Normand L, Dascotte P, Barré L. J Colloid Interface Sci 1998;207:
218.
[14] Mansot JL, Hallouis M, Martin JM. Colloids Surf 1993;75:25.
• Improvements in engine design for better fuel economy and [15] Connor SPD, Crawford J, Cane C. Lubrication Sci 1994;6:297.
longer oil drain intervals [16] Moreton DJ. United States patent, 6200936.
• Environmental legislation aimed at providing reduced gaseous [17] Gallo R, Jacquet F, Hoornaert P, Roman J-P. Rev Inst Fr Pét 1991;46:251.
emissions (by imposing chemical limits on the fuel and [18] Abou el Naga HH, Abd El-Azim WM, Bendary SA, Awad NG. Indian Eng
Chem Res 1993;32:3170.
lubricant)
[19] Markovic I, Ottewill RH, Cebula DJ, Field I, Marsh JF. Colloid Polym Sci
• With the use of more refined base oils (which are more aliphatic 1984;262:648.
in nature), the necessity for development of stable detergents [20] Markovic I, Ottewill RH. Colloid Polym Sci 1986;264:65.
with optimised performance is increasing. [21] Markovic I, Ottewill RH. Colloid Polym Sci 1986;264:454.
[22] Markovic I, Ottewill RH, Singra E, Marsh JF, Heenan RK. Colloid Polym
This has manifested in the development of novel surfactant Sci 1992;270:602.
[23] Bearchell CA, Heyes DM, Moreton DJ, Hanks TN. Phys Chem Chem Phys
structures. However, a significant amount of research is still re- 2000;2:5197.
quired to characterise and optimise these new systems before they [24] Tobias DJ, Klein ML. J Phys Chem 1996;100:6637.
become commercially viable. [25] Wu RC, Papadopolous KD, Campbell CB. J AIChE 1999;45:2011.
In addition, other issues should be addressed with regard to [26] Hone DC, Robinson BH, Steytler DC. Langmuir 2000;16:340.
environmental acceptability, in particular the synthesis routes. [27] Papke BC. Tribol Trans 1988;31:420.
Current synthesis methods rely on using mega-tonne quantities
of solvents (xylene, methanol, etc.), here work could seek to

Das könnte Ihnen auch gefallen