You are on page 1of 6

Applied Catalysis B: Environmental 69 (2007) 207–212

www.elsevier.com/locate/apcatb

Green approach for the preparation of biodegradable lubricant


base stock from epoxidized vegetable oil
Piyush S. Lathi *, Bo Mattiasson
Department of Biotechnology, Center for Chemistry and Chemical Engineering, Lund University,
P.O. Box 124, SE 221 00 Lund, Sweden
Received 10 April 2006; received in revised form 16 June 2006; accepted 20 June 2006
Available online 4 August 2006

Abstract
A novel process for the production of biodegradable lubricant-based stocks from epoxidized vegetable oil with a lower pour point via cationic
ion-exchange resins as catalysts was developed. This involves two steps, first, ring-opening reactions by alcoholysis followed by esterification of
the resultant hydroxy group in the first step.
The ring-opening reaction of epoxidized soybean oil with different alcohols such as n-butanol, iso-amyl alcohol and 2-ethylhexanol was carried
out in presence of Amberlyst 15 (Dry) as a catalyst; identity of products was confirmed by IR and NMR. Pour points of the products were observed
in the range of 5 to 15 8C. The hydroxy group of ring-opening product of n-butanol was further reacted with acetic anhydride in presence of
catalyst Amberlyst 15 (Dry), which was previously used to carry out ring-opening reaction by alcoholysis and identity of the resulting product was
confirmed by IR. Pour point of the resulting product was observed to be 5 8C.
# 2006 Elsevier B.V. All rights reserved.

Keywords: Bio-lubricants; Epoxidized soybean oil; Cationic ion-exchange resins

1. Introduction and nitrogen compounds with traces of a number of metals.


Due to their inherent toxicity and non-biodegradable nature
Renewable raw materials are going to play a very noteworthy they pose a constant threat to ecology and vast ground water
role in the development of sustainable green chemistry. They reserves [4]. Synthetic oils include others polyalphaolefins,
offer a large number of possibilities for applications which can be synthetic esters and polyalkylene glycols. Polyalphaolefins are
rarely met by petrochemistry [1]. Oils and fats of vegetable and petrochemical derived synthetic oils that mostly resemble
animal origin share the greatest proportion of the current mineral oils. Synthetic esters form a large group of products,
consumption of renewable raw materials in chemical industry which can be either from petrochemical or oleochemical
[2]. origin. In the preparation of re-refined oil, oil undergoes an
A steady increase in the use of eco-friendly consumer extensive re-refining process to remove contaminants to
products like lubricants has occurred as a result of strict produce fresh base oil.
government regulation and increased public awareness for a These concerns have resulted in an increasing interest in
pollution free environment [3]. There are wide ranges of vegetable oils with high content of oleic acid, which are
lubricant base oils in current use which includes mineral oils, considered to be potential substitutes to conventional mineral
synthetic oils, re-refined oils, and vegetable oils. Among these, oil-based products [5,6]. Vegetable oil lubricants are preferred
mineral oils are the most commonly used. They consist not only because they are renewable raw materials but also
predominantly of hydrocarbons but also contain some sulfur because they are biodegradable and non-toxic [7]. They also
acquire most of the properties required for lubricants such as
high index viscosity, low volatility and good lubricity and are
* Corresponding author. Tel.: +46 46 222 73 63; fax: +46 46 222 47 13.
also good solvents for fluid additives. However, vegetable oils
E-mail addresses: Piyush.Lathi@biotek.lu.se, lathips@yahoo.com, have poor oxidative and thermal stability, which is due to the
lathips@gmail.com (P.S. Lathi). presence of unsaturation [8]. This unsaturation restricts their
0926-3373/$ – see front matter # 2006 Elsevier B.V. All rights reserved.
doi:10.1016/j.apcatb.2006.06.016
208 P.S. Lathi, B. Mattiasson / Applied Catalysis B: Environmental 69 (2007) 207–212

use as a good lubricant. Several attempts have been made to the second step, i.e. esterification of the resulting hydroxyl
improve their oxidative stability such as transesterification of group with acid anhydride was done via pyridine [4,26,27].
trimethylopropane and rapeseed oil methyl ester [9]; selective Though these products come with the tag of green and
hydrogenation of polyunsaturated C C bonds of fatty acid biodegradability, the reported ways are not eco-friendly and
chains [10] and conversion of C C bonds to oxirane ring via green. Maybe therefore, these processes did not gain industrial
epoxidation [11,12]. Among these, epoxidation received importance.
special attention because it opened up a wide range of feasible There were few reports in which cationic ion-exchange
reactions that can be carried out under moderate reaction resins such as Dowex 50W-X8, Duolite C 26, and Lewatit S 100
conditions due to the high reactivity of the oxirane ring [13]. were used as an acylative ring cleavage catalyst for epoxidized
For instance, the epoxide can react with different nucleophiles soybean oil [28]. These organic catalysts showed poor cleavage
to produce mono-alcohols, diols, alkoxyalcohols, hydroxye- efficiencies as compared to sulfuric acid. It may be due to the
sters, N-hydroxyalkylamides, mercaptoalcohols, aminoalco- microporous gel type nature of Dowex 50W-X8 and Lewatit S
hols, hydroxynitriles, etc. [1,14]. 100. Since they have no discrete pores, solute ions have to
Standard industrial production for epoxidation of vegetable diffuse through the particle to interact with the exchange site
oil is based on in situ epoxidation, in which peracid is generated which may cause the diffusional limitation and reduce the
by reacting acetic or formic acid with hydrogen peroxide in the overall efficiency.
presence of strong mineral acids such as H2SO4 and H3PO4 The use of homogenous catalysts is very common in the
[15]. The strong mineral acid leads to many side reactions, such chemical and refinery industries, and those technologies
as oxiran-ring opening to diols, hydroxyesters, estolides and employ highly corrosive, hazardous and polluting liquid acids.
other dimers formation. Furthermore, they cause equipment The present work deals with the preparation of lubricant-based
corrosion and must be neutralized and also removed from the stock by using cationic ion-exchange resins, particularly, the
end product. macroporous variety. It is very well documented that resins are
Warwel and Klaas reported an alternative process that is very versatile catalysts and are known for carrying out number
milder and more selective, wherein lipase is used to catalyze of reactions such as esterification, etherification, transalkyla-
the peracid formation from a fatty acid and hydrogen peroxide tions, hydration and alkylation [29]. They also offer distinct
[16]. This is a new and a very promising technique for advantages over homogenous catalysts with respect to
epoxidation of double bonds. It has several advantages over corrosion, product recovery, selectivity, etc., both from the
the chemical catalysts such as: (i) mild reaction conditions, standpoint of catalysis as well as engineering of reactions for
(ii) formation of stable hydroperoxides directly from fatty commercial purposes [30]. Hence, it is worthwhile to develop a
acid, i.e. no need for acetic or formic acid addition, (iii) high green and environmental friendly process based on macro-
region and stereoselectivity, (iv) significant suppression of porous resins with an emphasis on biodegradability of the
side reactions and (v) high conversion. In most of the recent lubricant. As previously mentioned, the preparation of
publications, volatile organic solvents were used as the lubricant-based stock from epoxidized vegetable oil involves
reaction media [17,18]. However recently Orellana-Coca two steps, ring-opening reaction via alcoholysis followed by
et al. from our department has developed the solvent free esterification of the resulting hydroxy group through alcoho-
chemo enzymatic process for the epoxidation of fatty acid lysis. Novelty of the current work is that the same catalysts
[19]. without any treatment can be used for both the reactions, i.e.
It was reported earlier that epoxidized unsaturated fatty ring-opening reaction via alcoholysis and also for esterification
acids can be used as metal working fluids and lubricating of the resulting hydroxyl group.
additives to eliminate corrosion from chlorine containing
compounds [20,21]. Adhvaryu and Erhan reported epoxidized 2. Materials and methods
soybean oil as a potential source for high temperature
lubricants applications [12]. Various procedures have been 2.1. Materials
described for the preparation of alkoxyalcohols and hydro-
xylated fatty acids. Gast et al. reported that the product The epoxidized soybean oil was a gift sample from Akzo
obtained by ring-opening reaction of epoxidized fatty acid Nobel Industrial Coatings Sweden, and was used as received.
esters followed by esterification of the resulting hydroxyl Unless otherwise specified, all other chemicals including
group shows good performance for lower temperature alcohols, Amberlyst 15 (Dry) (% water content 1.5 m3/m3 of
lubricant applications [22]. resin, ionic form H+) and acetic anhydride were purchased from
In most of these processes, sulfuric acid was used as a Sigma and used without further purification.
catalyst for alcoholysis, whereas the other homogenous
catalysts such as p-toluenesulfonic acid, boron trifluoride, 2.2. Analytical methods
and sodium methoxide were used for comparative studies. The
resulting hydroxy group via alcoholysis is further reacted with The oxirane value was determined titrimetrically with
acid anhydride in presences of pyridine [23–25]. In some perchloric acid [31]. The infrared (IR) spectra were recorded
cases, the first step of the ring opening was performed by on a Fourier transform IR 8300 from Shimadzu. All the
addition of acetic acid, perchloric acid or formic acid, and in spectra’s were recorded by spreading the samples between
P.S. Lathi, B. Mattiasson / Applied Catalysis B: Environmental 69 (2007) 207–212 209

NaCl plates. The spectra were recorded in transmittance mode.


The 1H and 13C NMR spectra were recorded qualitatively using
a Bruker ARX-400 spectrometer (Bruker, Rheinstetten,
Germany) at a frequency of 400 and 100 MHz, respectively,
using a 5 mm dual probe. For 1H and 13C experiments, sample
solutions were prepared in deuterated chloroform (CDCl3
99.8% D, Aldrich Chemical Company). Proton NMR spectra
were obtained from 16 co-added FIDs with a delay time of 1 s
and 13C experiment was carried out with NS = 5000, AQ
0.62 s, DW 019 ms, D1 = 0.1 s. Pour point measurement was
done using a Freezer Hetofrig CB 8 (ISO 3106). Kinematic
viscosity measurement was done at 40 8C by using a Schott-
Geräte CT 42 (ISO 3104) and dynamic viscosity measurement
was done at 28 and 40 8C by using a Bohlin Visco 88
Viscometer.

3. Experimental set-up

3.1. Ring-opening reaction

The experimental set-up consisted of a 3 cm i.d. fully baffled


mechanically agitated reactor of 150 cm3 capacity, with a four-
bladed pitched-turbine impeller. The entire reactor assembly
was immersed in a thermostatic oil bath maintained at a desired
temperature with an accuracy of 1 8C.
In a typical experiment, the reaction mixture consisted of
epoxidized soybean oil 1 (60 g, 0.285 mol of epoxy group) and
an alcohol. The reaction mixture was agitated at 70 8C for
15 min at a speed of 1000 rpm and 1.5% of (w/w of epoxidized
soybean oil) catalyst Amberlyst 15 (Dry) was added to initiate
the reaction to obtain the products 2a, 2b and 2c. Table 1
summarizes the reaction conditions. Clear samples were
withdrawn periodically and titrated to measure the concentra-
tion of epoxide.

3.2. Esterification of resulting hydroxyl group in the ring 4. Results and discussion
opened product
4.1. Effect of alcohols on pour point and viscosity
The same experimental set-up mentioned in Section 3.1 was
used. The reaction mixture consisted of 50 g of product For investigation of alkyl group effect on the low-temperature
obtained by ring-opening reaction of epoxidized soybean oil properties, three different alcohols, namely n-butanol, iso-amyl
with n-butanol (2a) and 50 mL of acetic anhydride. The alcohol and 2-ethylhexanol were chosen [32,33]. The main
reaction mixture was agitated at 90 8C for 15 min at a speed of purpose for choosing the above three alcohols is that n-butanol is
1000 rpm and the catalyst which was used in the previous a short straight chain alcohol, iso-amyl alcohol is a short branch
reaction was added without any further treatment to initiate the chain alcohol and 2-ethylhexanol is a medium branch chain
reaction to obtain product 3. alcohol. Pour point values for the ring-opening product of the

Table 1
Reaction conditions
Product entry Alcohol Temperature (8C) Mole ratio Catalysts loading (%) Reaction time (h)
2ai n-Butyl alcohol 100 1:2 2 15
2aii n-Butyl alcohol 100 1:3 2 15
2b iso-Amyl alcohol 110 1:2 2 17
2c 2-Ethyl hexanol 120 1:2 2 24
3 Esterification of resulting hydroxyl 90 1:2 2 15
group in the ring opened product 2ai
210 P.S. Lathi, B. Mattiasson / Applied Catalysis B: Environmental 69 (2007) 207–212

respective alcohols were in the range 0 to 5, 5 to 10 and 10


to 15 8C. Thus, it was observed that as the chain length of
branched alcohol increased, it gives lower pour point. 2-
ethylhexanol, a branched medium chain alcohol, gives low pour
point values as compared to other alcohols.
The ring-opening reaction was monitored by epoxy titration
and stopped when the ring-opening reaction was 100%. The
structure of the product was confirmed from infrared spectra by
the disappearance of epoxy group at 822 cm 1 and appearance
of the hydroxy peak at 3450 cm 1. The reaction completion
was further confirmed by 1H NMR spectra. As summarized in
Table 1, different alcohols required different conditions to
complete the reaction. Longer or bulkier alcohols required
longer reaction time and higher temperature.
The retention of the triacylglycerol backbone is important
for maintaining the biodegradability of the vegetable oil and
this was confirmed by 1H NMR spectra given in Fig. 1. 1H
NMR measurement on epoxidized vegetable 1 indicates that
the epoxy group is present in the d 3.0–3.2 ppm region. The
methine proton of –CH2–CH–CH2 backbone at d 5.1–
5.3 ppm, methylene proton of –CH2–CH–CH2– glycerol’s
backbone at d 4.0–4.4 ppm, CH2 proton adjacent to two epoxy
group at d 2.8–3.0 ppm, –CH– proton of the epoxy ring at d
3.0–3.2 ppm, a-CH2 to C O at d 2.2–2.4 ppm, a-CH2 to
epoxy group at 1.7–1.9 ppm, b-CH2 to C O at d 1.55–
1.7 ppm, b-CH2 to epoxy group at d 1.4–1.55 ppm, saturated
methylene group at d 1.1–1.4 ppm and terminal –CH3 groups
at d 0.8–1.0 ppm region.
Compounds 2a and 3 retain most of the characteristic peaks
Fig. 1. Comparison of the 1H NMR spectra of the epoxidized vegetable oil (1), of compound 1 except the one at d 2.8–3.2 and 1.4–1.55 ppm
product of the ring-opening reaction of epoxidized vegetable oil with n-butanol region, corresponding to hydrogen’s attached to the epoxy
(2ai) and product of the esterification reaction of resulting hydroxyl group in the groups and methylene groups adjacent to epoxy, respectively.
product 2ai with acetic anhydride (3). 13
C NMR measurement on epoxidized vegetable 1 indicates
that the epoxy group is present in the d 56.9–58.5 ppm region.

Fig. 2. Comparison of the 13C NMR spectra of the epoxidized vegetable oil (1), product of the ring-opening reaction of epoxidized vegetable oil with n-butanol (2ai)
and product of the esterification reaction of resulting hydroxyl group in the product 2ai with acetic anhydride (3).
P.S. Lathi, B. Mattiasson / Applied Catalysis B: Environmental 69 (2007) 207–212 211

Fig. 3. Comparison of the infrared spectra of the epoxidized vegetable oil (1),
product of ring-opening reaction epoxidized vegetable oil with n-butanol (2a)
and product of esterification reaction of resulting hydroxyl group in the product
2a with acetic anhydride (3).

Compounds 2a and 3 retain most of the characteristic peaks of


compound 1 except the one at d 56.9–58.5 ppm region,
corresponding to carbon’s attached to the epoxy groups (Fig. 2).
This gives the confirmation that no epoxide group remains in
the products.
The resulting hydroxyl group in 2a was further reacted with Fig. 4. Dynamic viscosity study performed at 28 8C at different shear rates to
acetic anhydride to obtain the corresponding product 3. The observe its effect on the viscosity: (a) shear rate vs. viscosity and (b) shear rate
product was confirmed by IR spectra by the disappearance of vs. shear stress.
the hydroxyl peak at 3450 cm 1. IR spectra of 1, 2a and 3 are
given in Fig. 3. The reaction completion was further confirmed
by 1H NMR spectra by the presence of an additional peak at d
2.0 ppm for product 3.
Pour point becomes very important in cold weather
applications and in cylinder and packing lubricant with very
cold suction temperature. The pour points of 2a–2c and 3 were
measured by using a standard method mentioned in materials
and method and the results are shown in Table 2.
Viscosity is of paramount importance to a lubricant. The
kinematic viscosity of 2a–2c and 3 were measured and the
results are shown in Table 2. The product 2a was synthesized
with two different mole ratios of epoxidized oil to n-butanol
and it was found that the viscosity of 2aii is higher than that of
2ai. Product 2aii was prepared by using higher molar
concentration of n-butanol.
Dynamic viscosity of 2ai was measured at 28 and 40 8C at
different shear rates. Figs. 4b and 5b show that viscosity does
not change with an increasing shear rate, whereas the plots of
shear stress versus shear rate (Figs. 4a and 5a) show a straight

Table 2
Kinematic viscosity and pour point values
Serial no. Product Kinematic Pour point
entry viscosity (mm2/s) (8C)
1 2ai 232.7 > 5, <0
2 2aii 312.9 > 5, <0
3 2b 172.4 > 5, <0
4 2c 73.2 > 15, < 10 Fig. 5. Dynamic viscosity study performed at 40 8C at different shear rates to
5 3 350.8 > 5, <0 observe its effect on the viscosity: (a) shear rate vs. viscosity and (b) shear rate
vs. shear stress.
212 P.S. Lathi, B. Mattiasson / Applied Catalysis B: Environmental 69 (2007) 207–212

line passing through the origin, thereby indicating Newtonian References


type fluid.
[1] H. Baumann, H. Buhler, H. Fochem, F. Hirsinger, H. Zobelein, J. Falde,
Angew. Chem. 27 (1988) 41–62.
4.2. Biodegradability study
[2] U. Biermann, W. Friedt, S. Lang, W. Luhs, G. Machmuller, J.O. Metzger,
M.R. Klaas, H.J. Schafer, M.P. Schneider, Angew. Chem. Int. Ed. 39
A biodegradability and toxicity study for the lubricants was (2000) 2206–2224.
performed. It was observed that lubricants are non-toxic and [3] I. Rhee, NLGI Spokesman 60 (1996) 28.
biodegradable. In toxicity studies following tests such as [4] A. Adhvaryu, Z. Liu, S.Z. Erhan, Indust. Crops Prod. 21 (2005) 113–119.
toxicity screening, phytotoxicity [34] and algal toxicity [35] [5] S.J. Randles, M. Wright, Synthetic Lubric. 9 (1992) 145–161.
[6] S. Asadauakas, J.M. Perez, J.L. Duda, Lubric. Eng. 52 (1996) 877–882.
were executed. In each case no-observed-effect-concentration [7] N.S. Battersby, S.E. Pack, R.J. Watkinson, Chemosphere 24 (1992) 1998–
(NOEC); the highest concentration without adverse effects was 2000.
2000, 1000 and 1000 mg/L, respectively. In biodegradability [8] R. Becker, A. Knorr, Lubric. Sci. 8 (1996) 95–117.
study, values for biochemical oxygen demand (BOD) and [9] E. Uosukainen, Y.Y. Linko, M. Lamasa, T. Tervakangas, P. Linko, J. Am.
chemical oxygen demand (COD) were measured. It has been Oil Chem. Soc. 75 (1998) 1557–1563.
[10] L.E. Johansson, S.T. Lundin, J. Am. Oil Chem. Soc. 56 (1979) 974–980.
documented that if the ratio of BOD5/COD is around 0.5 and [11] S. Sinadinovic-Fiser, M. Jankovic, Z.S. Petrovic, J. Am. Oil Chem. Soc. 78
higher it means compounds are easily biodegradable [36,37]. (2001) 725–731.
This value for the present prepared lubricant products is in the [12] A. Adhvaryu, S.Z. Erhan, Indust. Crops Prod. 15 (2002) 247–254.
range of 0.43–0.63, which shows that compounds are [13] A. Padwa, S.S. Murphree, ARKIVOC 3 (2006) 6–33.
[14] L.A. Rios, P.P. Weckes, H. Schuster, W.F. Hoelderich, Appl. Catal. A: Gen.
biodegradable.
284 (2005) 155–161.
[15] Z.S. Petrovic, A. Zlatanic, C.C. Lava, S. Sindinovic-Fiser, Eur. J. Lipid
4.3. Reusability of catalyst Sci. Technol. 104 (2002) 293–299.
[16] S. Warwel, M.R. Klaas, J. Mol. Catal. B: Enzyme 1 (1995) 29–35.
Reusability of Amberlyst 15 (Dry) was tested by [17] G.D. Yadav, I.V. Borkar, AIChE J. 52 (2006) 1235–1247.
conducting four runs. After the reaction, the catalyst was [18] G.D. Yadav, K. Manjula Devi, J. Am. Oil Chem. Soc. 78 (2001) 347–351.
[19] C. Orellana-Coca, U. Toernvall, D. Adlercreutz, B. Mattiasson, R. Hatti-
filtered and then refluxed with 50 mL of alcohol for 30 min to Kaul, Biocatal. Biotransform. 23 (2005) 431–437.
remove any adsorbed material from the catalyst surface and [20] S. Watanabe, T. Fujita, M. Sakamota, J. Am. Oil Chem. Soc. 65 (1988)
pores and dried at 110 8C after every use. It was observed that 1311–1312.
there is only a marginal decrease in conversion. Some catalyst [21] D. Tao, H.L. Zhu, Z.M. Hu, Lubric. Sci. 8 (1996) 397–407.
was lost during filtration. There was no makeup catalyst added [22] L.E. Gast, C.B. Croston, W.J. Schneider, H.M. Teeter, Indust. Eng. Chem.
46 (1954) 2205–2208.
during each reuse study. Thus, the catalyst was reusable up to [23] H.M. Teeter, L.E. Gast, E.W. Bell, J.C. Cowan, Indust. Eng. Chem. 45
four times. (1953) 1777–1779.
[24] H. Hwang, S.Z. Erhan, J. Am. Oil Chem. Soc. 78 (2001) 1179–1184.
5. Conclusion [25] H. Hwang, A. Adhvaryu, S.Z. Erhan, J. Am. Oil Chem. Soc. 80 (2003)
811–815.
[26] B. Dahlke, S. Helbardt, M. Paetow, W.H. Zech, J. Am. Oil Chem. Soc. 72
The current work has focused on the synthesis of bio- (1995) 349–353.
lubricant from epoxidized vegetable oil and alcohol using solid [27] S.Z. Erhan, A. Adhvaryu, Z. Liu, US Patent 6,583,302 (2003).
acid catalyst Amberlyst 15. This investigation showed that the [28] M.H. El-Mallah, F.A. Zaher, M.M. El-Hefnawy, Res. Indust. 34 (1989)
introducing branching on epoxidized soybean oils leads to 286–289.
dramatic improvement of low pour point value. The ring- [29] G.D. Yadav, P.H. Mehta, Indust. Eng. Chem. Res. 33 (1994) 2198–2208.
[30] M.M. Sharma, React. Funct. Polym. 26 (1995) 3–23.
opening reaction of epoxidized soybean oils with 2-ethylhex- [31] R. Dijkstra, E.A.M.F. Dahmen, Anal. Chim. Acta 31 (1964) 38–44.
anol gives the low pour point value. It was observed that the [32] A. Özgulsun, F. Karaosmanoglu, M. Tuter, J. Am. Oil Chem. Soc. 77
catalyst has excellent reusability and stability. Lubricants are (2000) 105–109.
non-toxic and biodegradable. [33] N. Dörmö, K. Belafi-Bako, L. Bartha, U. Ehrenstein, L. Gubicza, Bio-
chem. Eng. J. 21 (2004) 229–234.
[34] T. Essam, H. Zilouei, A. Amin Magdy, O. El Tayeb, B. Mattiasson, B.
Acknowledgements Guieysse, Chemosphere 63 (2006) 277–284.
[35] Organization for Economic Cooperation and Development (OECD), Algal
The work is a part of the GREENCHEM program sponsored Acute Toxicity Test. Test Guideline 203. OECD Guidelines for the Testing
by MISTRA (The Swedish Strategic Fund for Environmental of Chemicals, OECD, Paris, France, 1984.
Research) and the authors thank the same for the financial [36] Organization for Economic Cooperation and Development (OECD),
Guidelines for Testing of Chemicals. Test 302 B: Zahn–Wellens/EMPA
support. The authors are very grateful to Mrs. Birgit Vainio Test 302 C: Modified MITI Test (II), OECD, Paris, France, 1993.
(Binol BioSafe Oy Laboratory, Finland) for the measurement of [37] G. Tchobanoglous, F.L. Burton, H.D. Stensel, Wastewater Engineering:
the viscosity and the pour point tests. Treatment and Reuse, McGraw Hill, New York, 2003.

You might also like