Sie sind auf Seite 1von 17

J.

of Supercritical Fluids 63 (2012) 133149


Contents lists available at SciVerse ScienceDirect
The Journal of Supercritical Fluids
j our nal homepage: www. el sevi er . com/ l ocat e/ supf l u
Review
Fluid properties needed in supercritical transesterication of triglyceride
feedstocks to biodiesel fuels for efcient and clean combustion A review

George Anitescu

, Thomas J. Bruno

Thermophysical Properties Division, National Institute of Standards and Technology, Boulder, CO 80305, United States
a r t i c l e i n f o
Article history:
Received 23 September 2011
Received in revised form
22 November 2011
Accepted 22 November 2011
Keywords:
Biodiesel fuel
Lipid feedstock
Supercritical processing
Fluid properties
Supercritical combustion
a b s t r a c t
This review focuses on the potential synergy between uid properties and supercritical (SC) processing/
combustion of biodiesel fuels. These fuels are the extenders/expanders of choice for petroleum-derived
diesel fuels (PDDF) due to overall performance in the environment, safety, feedstock, and fuel quality. A
typical biodiesel fuel meets commercial specications of the American Society for Testing and Materials
(ASTM D6751) or European Union (EN 14214). Biodiesel fuels, mainly mixtures of fatty acid methyl or
ethyl esters (FAMEs or FAEEs), are currently produced by base/acid catalytic transesterication (BAC-
TE) of triglyceride feedstocks with methanol or ethanol. These methods require rened oil feedstocks
and complex product separation/purication that leads to noncompetitive prices compared with PDDFs.
Alternatively, a noncatalytic technology based on SC-TE processing of various lipid feedstocks has been
reported to mitigate these drawbacks. One version of this technology, the one-step SC-TE method, poten-
tially has major advantages over the BAC-TE, mainly due to shorter reaction times (59min versus 16h)
and the reduction of glycerol to acceptable ASTM levels in fuels of superior quality. The latter advantage
originates from glycerol and polyunsaturated FAME/FAEE thermal conversion to lighter fuel products.
Based on technical and economic analyses, the manufacturing cost of biodiesel fuels from a one-step
SC-TE process could be one half of the BAC-TE current cost. To optimize biodiesel fuel production and
quality, leading to a more efcient and clean combustion, a close connection between uid properties
and fuel processing/combustion must be considered. Insights from recent case studies and real-world
examples of applications of the principles of sustainability in the development and implementation of
biodiesel fuel projects are given. The reviewincludes sustainability metrics, resource efciency, and sus-
tainable process integration. These themes are woven together into a perspective on how sustainability
and green-chemistry principles are being implemented for cost-effective biodiesel fuel production and
advanced combustion.
Published by Elsevier B.V.
Contents
1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134
2. Triglyceride feedstock methanol/ethanol properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136
2.1. Overview. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136
2.2. Triglyceride feedstock composition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136
2.3. Heat capacity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136
2.4. Viscosity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
2.5. Density and thermal conductivity. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
2.6. Miscibility of triglycerides with alkyl alcohols . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
2.7. Thermodynamic parameters of triglyceride feedstocks and mixtures with alcohols . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138
2.8. Phase transitions in triglyceridealcohol mixtures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139

This article is not subject to U.S. copyright. Certain commercial equipment, materials or supplies are identied in this paper to adequately specify the experimental
procedure or description. Such identication does not imply recommendation or endorsement by the National Institute of Standards and Technology, nor does it imply
that the equipment, materials or supplies are the best available for the purpose.

Corresponding author. Present address: Biomedical and Chemical Engineering Department, Syracuse University, Syracuse, NY 13244, United States.
Tel.: +1 315 443 1917; fax: +1 315 443 9175.

Corresponding author. Tel.: +1 303 497 5158; fax: +1 303 497 5044.
E-mail addresses: ganitesc@syr.edu, George.Anitescu@colorado.edu (G. Anitescu), bruno@boulder.nist.gov (T.J. Bruno).
0896-8446/$ see front matter. Published by Elsevier B.V.
doi:10.1016/j.supu.2011.11.020
134 G. Anitescu, T.J. Bruno / J. of Supercritical Fluids 63 (2012) 133149
3. Properties involved in supercritical transesterication . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139
3.1. Overview. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139
3.2. Energies of activation from the kinetics of TE reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139
3.3. Reversibility/irreversibility of SC-TE reactions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140
3.4. Heat of the overall TE reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140
3.5. SC-TE process design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
4. Properties of biodiesel fuels produced by SC-TE processes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
4.1. Overview. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
4.2. Viscosity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
4.3. Volatility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142
4.4. Thermal conductivity. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143
4.5. Density and speed of sound . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143
4.6. Fit-for-purpose properties. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143
4.6.1. Higher heating value (HHV) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143
4.6.2. Cetane number . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143
4.7. Properties from computational techniques . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144
4.8. Blended PDDF-biodiesel fuels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144
5. Combustion of biodiesel fuels produced by SC-TE methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144
5.1. Overview. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144
5.2. Spray/jet characteristics and fuel properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144
5.3. Biodiesel fuel properties and combustion quality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145
5.4. In-engine biodiesel fuel generation and SC combustion. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146
6. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146
Appendix A. Supplementary data. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147
References. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147
1. Introduction
Biodiesel fuels, mainly mixtures of fatty acid methyl or ethyl
esters (FAMEs or FAEEs), have been the focus of signicant media
attention, industrial interest, and scientic research. These fuels
are potential blending additives for petroleum-derived diesel fuels
(PDDFs) due totheir overall performance onenvironmental impact,
safe handling, feedstock availability, and fuel quality [1,2]. A mer-
chantable biodiesel fuel meets commercial specications of the
American Society for Testing and Materials (ASTM) or European
Union norms (EN), and is the only fuel produced at a commercial
scale that qualies as an advanced biodiesel under the US EPA
Renewable Fuels Standard [35]. Biodiesel fuels exhibit increased
lubricity compared to ultra-low-sulfur PDDF, have no signicant
health risk, are biodegradable, and decrease combustion emissions
of carbon monoxide, unburned hydrocarbons (UHC), and partic-
ulate matter (PM) [1]. To be a viable PDDF extender/expander,
biodiesel fuels shouldbe protable without signicant government
subsidies, should not compete with food supplies, and should pro-
vide a net energy gain through the entire production-use cycle.
Thesefuels couldprovide93%moreusableenergythantheenergy
needed for their production and could reduce net greenhouse
gas emission by 41% compared with PDDFs [6]. Disadvantages
that must be addressed include oxidative instability, mois-
ture absorption, higher viscosity, cold-ow properties, and NO
x
emissions [1].
Triglyceride feedstocks and methanol/ethanol availability and
renewability have favored production of commercial biodiesel
fuels by base/acid catalytic transesterication (BAC-TE) methods.
Nevertheless, various technical and economic aspects require fur-
ther improvement of both processing and combustion of these
fuels [611]. Sustainable production pathways must be established
to implement cost-effective processing technologies [1216]. To
optimize TE processes leading to competitive and high-quality
fuels with more efcient and clean combustion, the connection
between uid properties and fuel processing/combustion must
be addressed. Fluid properties represent the integrating factor
to keep the fuel production and application steps synchronized.
These steps include feedstock selectionandpretreatment, bioren-
ing, blending, transport, storage, and combustion. Fuel properties
contribute to the selection of the optimum upstream processing
pathways anda downstreamcombustiontechnology. While for n-
ished fuels there are sets of specications (ASTM D6751 and EN
14214) [3,4], the most wide-ranging and demanding needs are for
properties required for chemical process design in biodiesel fuel
production.
Biodiesel fuels considered in this review are largely limited
to those produced by emerging supercritical (SC) TE methods
(triglyceridealcohol mixtures in SC states). Various aspects of
these methods are describedinrecent reviewarticles [1720]. Rep-
resentative references on this topic for laboratory-scale research
are shown in Table 1 [2151]. Reaction conditions for the feed-
stocks studied are included along with reactor type and maximum
conversion attained. As potential alternatives to the current BAC-
TE commercial technologies, the SC-TE methods can give nearly
complete conversion in a short processing time (minutes versus
hours). In addition to the high conversion and reaction rates, a
SC-TE process can tolerate feedstocks with a high concentration of
free fatty acids (FFAs) and water, up to 36% and 30% (mass/mass),
respectively [1719]. The development of the early SC-TEprocesses
was, however, hinderedby concerns that temperatures higher than
350

C would be detrimental to fuel quality, while large alcohol-to-


oil molar ratios (42:1) would be required to achieve high levels of
conversion at lower temperatures [1720].
To overcome the main disadvantages of the emerging SC-TE
methods (e.g., large excess of methanol and glycerol separation),
an advanced, cost-effective biodiesel fuel technology based on a
one-stepSC-TE process at 400

Cwas reported[7,21]. This contin-


uous processingmethodclaims major advantages over bothBAC-TE
and early SC-TE methods, due to shorter processing time (<10min),
lower molar ratio of alcohol to triglycerides (<12:1), and elimina-
tionof theglycerol toacceptableASTMstandardlevels inthebiofuel
[7,21,5053]. The latter advantage is due to thermal decomposi-
tion of the resulting glycerol to fuel products of lower molecular
weight, or the formation of acceptable oxygenated fuel products
due to etherication. Consequently, the quantity (on a mass basis)
of fuel produced by this method is approximately the same as the
mass of the reactants, and its quality appears to be superior to that
of biodiesel fuels produced by lower-temperature methods: lower
viscosity, better cold-owproperties, and improved volatility [52].
G. Anitescu, T.J. Bruno / J. of Supercritical Fluids 63 (2012) 133149 135
Table 1
Reaction conditions reported for noncatalytic supercritical transesterication (SC-TE).
Oil (fat)/cosolvent T (

C) P (MPa) ROH:TG
a
(molar) Residence time (min) Reactor Maximum conversion (%) Ref.
Soybean 350425 1025 36 26 Continuous 100 [21]
Soybean 250350 1243 43 1590 Batch 100 [22]
Coconut; Palm 270350 1019 642 133 Continuous 95; 96 [23]
Canola 420450 40 1145 4 Continuous 100 [24]
Soybean 280 6.27.4 1520 20120 Batch 96 [25]
Chicken fat; Tall oil 275325 16.5 1040 20 Batch 91 [26]
Soybean 310 35 40 25 Continuous 96 [27]
Castor; Linseed 200350 20 1070 10300 Batch 98 [28]
Soybean 350 20 40 N/A Continuous 78 [29]
Soybean/CO
2
280 14.3 24 10 Batch 98 [30]
Soybean/C
3
H
8
280 12.8 24 10 Batch 98 [31]
Sunower 200400 20 40 40 Batch 98 [32]
Soybean/C
3
H
8
270315 8.821.1 64 10 Batch 99 [33]
Palm 270350 2040 40 525 Continuous 100 [34]
Soybean/C
6
H
14
/CO
2
260350 N/A 42 10 Batch 98 [35]
Soybean/acid 270350 0.115 2060 821 Continuous 91 [36]
Palm; Jatropha 200400 20 50 40 Batch 100 [37]
Palm kernel/C
6
H
14
290350 1522 1242 10 Batch 87 [38]
Palm 200400 40 380 0.520 Batch 95 [39]
Jatropha 239340 5.78.6 1043 4 Batch 100 [40]
Soybean 200375 720 10100 117 Continuous 85 [41]
Rapeseed 200487 7105 42 4 Batch 95 [42]
Camelina/C
6
H
14
290 11.4 45 40 Batch 90 [43]
Sesame; Mustard 275350 1920 3080 4570 Batch 100 [44]
Jatropha/C
6
H
14
/CO
2
200300 24 25 4580 Batch
b
100 [45]
Radish 295325 914 3252 1529 Batch 95 [46]
Castor 300 20 2040 1550 Continuous 75 [47]
Soybean/CO
2
250325 1020 2040 1550 Continuous 80 [48]
Camelina/CO
2
400 19.6 42 15 Batch N/A [49]
Chicken fat 350400 1030 312 310 Continuous 100 [50]
Triolein 355400 15 9 210 Continuous 98 [51]
a
ROH is either methanol or ethanol.
b
Extraction+in situ TE.
Furthermore, the manufacturing cost of the one-stepSC-TEmethod
is assessed to be about half that of BAC-TE [7,21].
To assess the technological feasibility and obtain material and
energy balances for economic analyses, process simulations must
be performed. Despite some differences between a process simu-
lation and a real-world operation, the former is commonly used to
provide reliable information on process operation. This capability
is based on component libraries, comprehensive thermophysical
property packages, and advanced computational methods. The rst
simulation steps consist in selecting the chemical components for
the process along with a thermodynamic model. Additionally, unit
operations and operating conditions, plant capacity and input con-
ditions must all be selected and specied [711]. To meet the
property requirements for biodiesel fuel production and appli-
cations, comprehensive databases are being created or extended
[15,54]. For biodiesel fuel-processing steps, process simulation
packages such as ChemCad and Aspen Plus or HYSYS (ASPEN Tech)
are available [711,5557]. By using uid properties contained
in the databases of these programs as well as properties from
other available databases, one can design a process for optimized
biodiesel fuel production[711]. Anexample of suchanapplication
is available as supporting information (Appendix).
Specication of a process chemical component requires input
of a number of properties, such as the normal boiling point, densi-
ties, molecular weight, as well as the critical properties of the pure
substance. Complex mixtures, including blends of biodiesel fuels
with PDDFs, are typically modeled as surrogate blends of a limited
number of components that capture the essential characteristics
of the uid mixtures. Property measurements for pure compounds
in a surrogate include but are not limited to densities, heat capac-
ities, vapor pressures, enthalpies, thermal conductivities, critical
properties, speeds of sound, and viscosities. Thermodynamic con-
cepts such as structure-reactivity relationships, group contribution
methods, and extended corresponding-states models are currently
applied to generate acceptable surrogate blends [15,58,59].
Various specications that a biodiesel fuel must meet are con-
tained in relevant standards, such as ASTM D6751 and EN 14214
[3,4]. Among the fundamental and t-for-purpose properties in
these standards are cetane number, kinematic viscosity, oxidative
stability, and cold-ow properties. Other important properties to
consider that are not contained in these standards are exhaust
emissions under specied combustion parameters, lubricity, and
heats of combustion. Property ranges of biodiesel fuels obtained
fromvarious feedstocks by different TE methods are wide, because
any property is dependent on fuel composition. Moreover, proper-
ties reportedwithout compositional data cannot be usedtodevelop
property models. To assess the completeness of the TE reactions
according to the total glycerol level specied (maximum 0.24%
(mass/mass)), analytical methods such as gas chromatographic
(GC) analysis are required.
In this review we focus on uid properties relevant to biodiesel
fuel production by SC-TE methods and to engine performance. This
approach will be useful to a broad range of scientists and engineers
working on a scientic foundation for biodiesel fuels and advanced
combustion in diesel engines. Despite recent progress, there is
presently no single comprehensive source of reliable property
data for biofuels that can be called upon by industry. An example
is representative for the status of this information: During the
design and test phases associated with the development of the
portable biodiesel equipment, public domain information on the
material properties of supercritical methanol with any lipid was
found to be incomplete. Hence, the portable equipment was con-
servatively designed to account for these uncertainties [60]. The
new analyses and insights to the eld can stimulate researchers in
this area for new R&D projects. Contributions that dealt explicitly
with the ecological landscape and regulatory issues that will likely
136 G. Anitescu, T.J. Bruno / J. of Supercritical Fluids 63 (2012) 133149
Table 2
Relevant properties of SC-TE reactants and reaction products compared to those of n-hexadecane [54,57].
Property (SI unit) Methanol Ethanol Triolein Glycerol Methyl oleate n-Hexadecane
Molar mass (kg/kmol) 32 46 885 92 296 226
Density (kg/m
3
)/25

C 789 785 909 1,261 872 769


Normal boiling point (

C) 65 78 629 288 344 287


Melting point (

C) 98 114 5 18 20 18
Critical temperature (

C) 239 241 705 577 491 449


Critical pressure (MPa) 8.1 6.1 0.33 7.5 1.28 1.43
Critical density (kg/m
3
) 274 276 286 348 279 218
Dipole moment (D) 1.70 1.69 3.06 2.68 1.44 0.00
Acentric factor, 0.566 0.644 2.095 0.513 1.049 0.717
^cH

(MJ/kg) 23.875 29.696 37.077 16.054 37.500 47.341


C
p,G
(J/(mol K))/25

C 44 65 1301 115 445 369


Cp,L (J/(mol K))/25

C 81 112 1796 222 596 500


Viscosity (mPa s)/25

C 0.54 1.08 69 866 5.41 3.06


accompany the continuous growth of the biodiesel fuel industry
are also in high demand, but these issues are not discussed herein.
2. Triglyceride feedstock methanol/ethanol properties
2.1. Overview
Triglyceride feedstocks are discussed as reagents in TE reactions
for biodiesel fuel production, neglecting their direct use as fuels or
in blends with PDDFs. The needed reactant properties for the TE
reactions are those of both pure reactants (a brief selection is given
in Table 2) and glyceridealcohol mixtures. Relevant thermophys-
ical properties include thermodynamic (feedstock composition,
relativetriglyceridealcohol miscibility, density, heat capacity, crit-
ical parameters) and transport (viscosity, thermal conductivity,
diffusivity, and Reynolds number) properties. The most important
property is the composition, which, in turn, controls other proper-
ties. Theunrenedoils andfats usuallycontainFFAs, phospholipids,
sterols, water, and other impurities [61]. Water and FFA contents
are important quality parameters of the rawmaterial and vary sig-
nicantly with feedstock origin.
Reliable information is needed for reactor design to optimize
feedstock use and processing to produce biodiesel fuels of high
quality. Accordingly, thermophysical and chemical properties of
various reactive mixtures, as well as those of TE reaction interme-
diates and nal products, are required over appropriate ranges of
conditions. Duetothelackof propertyinformationfor triglycerides,
triolein is often selected as a surrogate lipid feedstock in designs
that require data for liquid and vapor equilibria. For comparison,
approximate values for some of the triolein properties are given in
Table 2, along with those of methanol, ethanol, and major compo-
nents of biodiesel (methyl oleate), and diesel fuels (n-hexadecane
or cetane) [54,57,62]. While the volatility and viscosity, for exam-
ple, of methyl oleate are dramatically lower than those of triolein,
these properties are still signicantly higher than those of cetane.
Glycerol stands apart by its unusually high density and viscosity,
which make it difcult to handle and process.
2.2. Triglyceride feedstock composition
The most important input factors in the production and
economics of biodiesel fuels are the quality and cost of triglyc-
eride feedstocks. The oil content varies, reportedly up to 70%
(mass/mass) for some algal strains [63]. The feedstocks for current
and potential use for biodiesel fuel production include soybeans,
rapeseed/canola, cottonseed, sunower seeds, groundnut, copra,
sesame, linseed, castor seed, corn/maize, palm, coconuts, jatropha,
cuphea, camelina, and pennycress. Soybean is a main feedstock
in the United States, whereas rapeseed dominates in Europe and
Canada and palm/coconut in Southeast Asia [12,17].
The major supplies of animal fats and greases for biodiesel
fuel production are tallow, poultry fat, and lard (white grease).
Yellow grease is manufactured from spent cooking oil. The FFA
content of these feedstocks is usually high (1030% (mass/mass)),
making them rather unattractive for conventional catalytic TE,
which requires less than 0.5% (mass/mass) FFA content [61].
This composition is, however, advantageous for SC-TE methods
because the easiness to esterify FFAs [1719]. Although insuf-
cient in scale to signicantly affect PDDF consumption, these
feedstocks are low-cost resources for biodiesel fuel production via
SC-TE.
Biodiesel fuel fromalgae has been seen as a potential substitute
for diesel fuels [63,64]. Microalgae are attractive because of their
high triglycerides or hydrocarbon content, and they can be grown
by use of water and land resources that do not interfere with food
crop production [63,64]. Despite the optimistic evaluations, some
aspects in life-cycle analyses are difcult to quantify [64,65]. The
main concerns are with water resources and economic feasibility.
Productionof liquidbiofuel fromalgae is currentlyat the R&Dphase
[66,67].
The differences in quality characteristics between various lipid
feedstocks produce variations in the biodiesel fuel properties. Sat-
urated triglycerides tend to produce biodiesel fuel with slightly
inferior cold-ow properties but with better storage stability than
that of unsaturated oils. Some tests have shown animal fats to pro-
ducebiodiesel fuels withhigher cetanenumber anda slightlybetter
engine emission prole [1,68].
2.3. Heat capacity
Heat capacity is an important uid property to be considered
when designing heat exchangers and calculating heats of reaction.
Heat exchangers play a major role in the energy balance of SC-TE
processes by improving heat recovery from the reaction products
which lowers the process cost. For pure uids, the divergence
of the heat capacity in the vicinity of the critical point is a well
known feature, whereas for uid mixtures this property is a con-
tinuous function of temperature. Our literature search yielded no
heat-capacity data for triglyceridealcohol mixtures at high tem-
peratures, including the SC region. To illustrate the heat-capacity
behavior for complex mixtures, a plot of heat capacity versus tem-
perature is given in Fig. 1 for a ternary mixture of a diesel-fuel
surrogate, n-hexadecane, with the main components of exhaust
gas recycle (EGR), namely N
2
, CO
2
, and H
2
O [53]. In a TE process,
N
2
is often used as an inert gas to displace air, CO
2
may be used as
cosolvent, and water is present as impurity.
Heat capacities at constant pressure (C
p
) of the reactants are
needed over wide PT ranges and can be determined by use of a
differential scanning calorimeter (DSC) capable of acquiring C
p
data
with uncertainty of less than 1% up to 35MPa and 300

C [69]. The
G. Anitescu, T.J. Bruno / J. of Supercritical Fluids 63 (2012) 133149 137
150
200
250
300
350
400
0 100 200 300 400 500
T (
o
C)
C
p

(
J
/
m
o
l
.
K
)

Bubble-point curve
Dew-point curve
Fig. 1. Molar heat capacity of n-hexadecaneCO
2
N
2
H
2
O mixture as function of
temperature calculated by PengRobinson equation of state (adapted from [53]).
The molar composition of the isopleth is 0.4859 (C
16
), 0.3804 (N
2
), and 0.0668 (CO
2
and H
2
O). The critical value is reached at 427

C.
basic equation for heat capacity determination by this calorimetric
method is:
^
sr
=
s

r
= C
ps
d1
s
dt
C
pr
d1
r
dt
= (C
ps
C
pr
) (1)
where is the heat ow rate, is the average heat ow rate, and
subscripts s and r indicate the sample and the reference, respec-
tively. The specic heat capacity of any sample is given by [69]:
C
s
=

s

0

rcj

0

m
rcj
m
s
C
rcj
(2)
In this equation,
0
is the heat ow rate of the baseline,
ref
is the
heat ow rate of a known amount (m
ref
) of calibration substance
with known specic heat capacity (c
ref
), and
s
is the heat owrate
of any known amount (m
s
) of the sample (reactant or product).
2.4. Viscosity
Viscosity is one of the most important transport properties
for uids involved in the continuous-ow SC-TE methods. Simi-
lar to the triolein behavior illustrated in Fig. 2 [54], most of the
lipid feedstocks for biodiesel fuel production are very viscous at
ambient temperature and pressure. The viscosity decreases rapidly
with heating of the feedstock. The thermal variation of this prop-
erty at ambient pressure can be considered linear within limited
ranges of both low and high temperatures (T <60

C and T >230

C,
respectively). There are also available viscosity data for methanol
[62,70].
2.5. Density and thermal conductivity
Density and thermal conductivity are some of the other proper-
ties needed inthe process designof SC-TE, including the preheating
step of the reactants. The reported data for these properties are
usually limited to low temperatures. However, available data
at moderately higher temperatures show a quasi-linear thermal
behavior along pressure isobars that permits some data extrap-
olation. As illustrated in Fig. 3 for palm oil, the density, thermal
Fig. 2. Triolein viscosity at atmospheric pressure as function of temperature. With
data from [54].
conductivity, and heat capacity exhibit such a behavior up to
300

C [71]. This characteristic is important for modeling feed-


stock properties in the liquid phase. This linear behavior changes
rapidly around and beyond the critical point of the uids. Reli-
able data for the SC region of oilalcohol solutions are urgently
needed.
2.6. Miscibility of triglycerides with alkyl alcohols
Information on mutual solubility of the reactants and reaction
products, often an overlooked issue, is essential for biodiesel fuel
production design and process operation. Different SC-TE reaction
conditions reported for near complete triglyceride conversion
originate from triglyceridealcohol properties with miscibility
0.0
0.5
1.0
1.5
2.0
2.5
3.0
50 100 150 200 250 300 350
Temperature (
o
C)
P
a
l
m

O
i
l

P
r
o
p
e
r
t
i
e
s
Heat capacity (kJ/kg
.
K)
Density (g/cm
3
)
Thermal conductivity (W/m
.
K)
Fig. 3. Heat capacity, thermal conductivity, and density of palm oil versus temper-
ature at atmospheric pressure [71].
138 G. Anitescu, T.J. Bruno / J. of Supercritical Fluids 63 (2012) 133149
Fig. 4. Miscibility of a soybean oilethanol mixture (1:16 molar ratio) being rapidly heated from room temperature to 400

C (the volume of the view cell is 1mL). With


ACS permission from [21].
playing a major role [21,33,7280]. At similar PT conditions, this
property depends strongly on the triglyceride composition and
the alcohol-to-triglyceride ratio. The former parameter affects
the solubility by different properties of the FA components such
as polarity and volatility, while the latter acts by lowering the
(pseudo)critical points of the oilalcohol mixtures. Stoichiomet-
ric amounts of liquid triglycerides and methanol are not very
miscible in the liquid phase, so that reactions occur mainly at
the interface between the two liquid phases [1]. If the reaction
mixture is a homogeneous phase, higher conversion rates can
be achieved than if the reaction occurs under heterogeneous
conditions. Reactants and the TE products are partially soluble
under various PTx conditions. With an increase in the mass
fraction of FAMEs, the alcohol solubility in the triglycerideFAMEs
phase increases [77]. Glycerol has a high afnity for alcohol,
making a signicant portion of the reactant unavailable for TE
reactions.
The miscibility of soybean oil and ethanol (1:16 molar ratio)
is illustrated in Fig. 4 for the case of rapid heating from 26

C to
400

C in a high pressure, high temperature view cell at constant


volume/density[21]. Withtwoliquidphases at ambient conditions,
the system shows increasing homogeneity with increasing tem-
perature and pressure until a single SC phase is reached. At these
conditions, the conversion of triglyceride to FAMEs is very rapid
compared to that at subcritical conditions [17,21,52].
2.7. Thermodynamic parameters of triglyceride feedstocks and
mixtures with alcohols
High-pressure and high-temperature PVT data are lacking for
many natural lipids that are processed with SC uids. PVT data
were measured with a static-type bellows apparatus for various
fats and oils at 1150MPa and 3080

C and compared to those of


pure components trilaurin, triolein, and tridecane [81]. The PVT
behavior for the fats and oils above their melting points was similar
and was explained in terms of molecular weight, iodine value, and
fatty acid composition. Tait, Flory, and simplied perturbed hard
chain theory (SPHCT) equations provided satisfactory correlation.
With the correlated parameters, the PVT behavior of the oils was
described to within 4.9% average deviation in pressure.
While the critical properties of triglycerides are difcult to mea-
sure due to thermal decomposition, the critical points of various
lipidalcohol mixtures are typically determined by using a view
cell/quartz tube and a suitable optical system [21,33,53,76,82,83].
In addition to Fig. 4, Fig. 5 shows a phase transition from an L-V
two-phase mixture of diesel fuel #2 and CO
2
to a SC phase [83].
In practice, however, the reactivity of the triglycerides and alcohol
could generate esters, which would affect the critical conditions by
changing the systemcomposition. This effect will likely be negligi-
ble at lowester concentrations, withhighheating ux that will lead
to short residence time until the desired temperature is reached.
G. Anitescu, T.J. Bruno / J. of Supercritical Fluids 63 (2012) 133149 139
Fig. 5. Vanishing L-V meniscus in the diesel fuelCO
2
mixture transitioning to a SC state at 420

C and 10MPa [83].


Alternatively, the critical parameters can be calculated approxi-
mately by using a group-contribution technique or other methods
[51,52,76,80].
2.8. Phase transitions in triglyceridealcohol mixtures
By observing the PTx conditions of uid phase transitions like
those in Figs. 4 and 5, one can build various binary phase dia-
grams [53,83]. With a relatively simple methodology, it is possible
to experimentally observe and construct PTx phase diagrams for
selected binary systems of triglyceridealcohol solutions. For the
purpose of lipidalcohol phase transition studies, special attention
is required to minimize the interference of the TE conversion by
a rapid heating of these solutions. By mapping these phase tran-
sitions to SC states, a set of conditions to carry out an optimized
SC-TE process can then be selected.
Fig. 6 shows the critical temperatures for two pseudo
binary triglyceridemethanol mixtures as functions of the alco-
hol:triglyceride molar ratio (data from [23]). For a given
composition, at T <T
c
, the system is usually in a heterogeneous
region, while at T >T
c
the system is in a single homogeneous SC
phase. A large excess of alcohol is required to bring the system to
a SC state at relatively low temperatures. For example, at 300

C,
the mixture is in SC states for a methanol:triglyceride molar ratio
Tc
(
o
C)
MeOH:TG (molar)
0
50
100
150
200
250
300
350
400
450
500
550
600
0 5 10 15 20 25 30 35 40 45
Tc/PKO
Tc/CCO
Fig. 6. Critical temperatures for pseudo binary mixtures of methanol with palm
and coconut oils (PKO and CCO, respectively) versus methanol to triglycerides (TG)
molar ratio. Data from [23].
beyond 3540. As shown in Table 1, most of the reported results
in the literature are for molar ratios of 4042, claiming the shift of
the TE equilibriumto the right under excess alcohol. Instead, these
ratios position the oilalcohol mixtures in a SC state at 300

C, with
high and rapid conversion. For this reason, at 350

C and 400

C, for
example, SC phases can be reached at ratios of 15 and 8, respec-
tively (Fig. 6), signicantly cutting the cost of methanol recycling.
This critical temperature versus feedstock composition behavior is
similar for all feedstocks rather than just the two pseudo binary
mixtures mentioned in Fig. 6.
3. Properties involved in supercritical transesterication
3.1. Overview
This section briey describes the SC-TE processing of lipid feed-
stocks to biodiesel fuels and how these specications determine
fuel quality. A comparison with BAC-TE is also included. While the
availability of widely diversied feedstocks is a signicant advan-
tage for biodiesel fuel production, the compositional variability of
these feedstocks renders the selection of the processing conditions
rather difcult. For example, the available feedstocks of triglyc-
erides for biodiesel fuel production contain mainly ve fatty acids:
palmitic, stearic, oleic, linoleic, and linolenic; There are, however,
35triglycerides containingtheseacids withdifferent structures and
corresponding properties.
Despite signicant progress, the science behind biodiesel fuel
production is not yet well developed. For example, the heat of the
TE reactions, TE reversibility, and catalytic activity at high tem-
peratures are a few issues that have not yet been convincingly
addressed. Often, the role of uid mixing in increasing TE con-
version conferred by a solid mesh/bed in a TE reactor could be
mistakenly assigned to catalytic activity. Also, as we have already
discussed, the effect of excess alcohol used in SC-TE reactions is, in
some cases, erroneously interpreted as a shifting factor of reaction
equilibrium.
3.2. Energies of activation from the kinetics of TE reactions
It has been reported that the conversion of various vegetable
oils intothe correspondingFAMEs was enhancedconsiderablyinSC
methanol (Table 1). The apparent activationenergies of the TE reac-
tions were found to be different for the subcritical and the SC states
of methanol (e.g., 11.2 and 56.0kJ/mol, respectively, at 28MPa and
a molar ratio of methanol to oil of 42) [84].
The global TE process between a triglyceride (TG) of three iden-
tical fatty acids (e.g., triolein) and methanol can be written as
TG + 3CH
3
OH 3RCOOCH
3
+C
3
H
5
(OH)
3
(3)
140 G. Anitescu, T.J. Bruno / J. of Supercritical Fluids 63 (2012) 133149
0.0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1.0
0 1 2 4 3 5 6 7 8 9 10
FAMEs
TG
DG
MG
P
r
o
d
u
c
t

y
i
e
l
d

(
w
t
%
)
Residence me (min)
Fig. 7. Typical temporal proles of monoglyceride (MG), diglyceride (DG), triglyc-
eride (TG), and FAME in the SC-TE of chicken fat with methanol at 400

C and 20MPa
[50].
This overall reaction is a sequence of other competitive or con-
secutive reactions, with TG going to diglycerides (DG), DG to
monoglycerides (MG), and MG to glycerol; FAMEs (RCOOCH
3
) are
produced in each step of these TE reactions:
TG + CH
3
OH RCOOCH
3
+DG (4)
DG + CH
3
OH RCOOCH
3
+MG (5)
MG + CH
3
OH RCOOCH
3
+C
3
H
5
(OH)
3
(6)
At SC conditions, glycerol can also react with methanol to generate
ethers and thermally reacts to glycerol products (GPs):
C
3
H
5
(OH)
3
+3CH
3
OH Ethers + GPs (7)
The chemistry described above forms the basis of the one-step SC-
TE method for biodiesel fuel production [7,21]. Reaction (7) is very
important for the SC-TE processes at high temperature by produc-
ing smaller molecular fuel components to improve the overall fuel
volatility, viscosity, and cold owproperties, as well as by consum-
ing the undesired byproduct, glycerol [21,5052].
The complexity of reactants and reaction products masks the
intrinsic kinetics and reaction mechanisms/pathways that con-
trol SC-TE conversion processes. This motivates the use of simple
model compounds, which mimic the important reactive moieties
but also allow reliable developments of reaction pathways, kinet-
ics, and mechanisms. To study the kinetics of this complex set of
reactions, individual triglycerides (e.g., triolein) were subjected to
SC-TE conditions of 300400

C, 20MPa, and up to 9:1 methanol to


triglyceridemolar ratiofor 210minresidencetime[51]. Acompre-
hensive chromatographic analysis was performed to determine the
effect of the reaction conditions on triglyceride conversion, inter-
mediatespecies temporal prole, andnal-product yield. Rendered
chicken fat was also studied at SC-TE conditions [50]. The tempo-
ral proles of the main reaction species are shown in Fig. 7 [50].
Beyond 6min, the polyunsaturated FAMEs start to decompose to
lower molecular-weight products, whichmainly include esters and
straight-chainhydrocarbons. Toobtainthe energy of activation(E
a
)
for the main reactions involved in the SC-TE process, the kinetic
experimental data can be modeled with suitable software.
Fig. 8. GC-FID chromatographs of methyl oleate before (bottom) and after reaction
(top) with glycerol at 400

C and 10MPa for 10min. Peak identication: 1 glycerol,


2 internal standard (butanetriol), 3 monoolein, 4 internal standard (tricaprin),
and 5 diolein [51].
3.3. Reversibility/irreversibility of SC-TE reactions
Theoretically, TE is a reversible reaction, particularly for small
ester-alcohol molecules. In the production of FAMEs, however, the
reverse reaction is negligible, especially at high temperature. To
prove that the SC-TE is not an overall signicantly reversible reac-
tion, a mixture of glycerol and methyl oleate was heated at the
reactionconditions for 10min. Noformationof methanol or triolein
was observed [51,52]. Only small amounts of monoglycerides and
traces of diglycerides were generated, as shown in Fig. 8. Accord-
ingly, the claim of excess methanol as a driving force of the TE
equilibrium toward the right side is not persuasive under practical
conditions. The large excess is actually needed to facilitate mix-
ing and to bring reactants to one homogeneous phase, as discussed
earlier [21,52,7280].
3.4. Heat of the overall TE reactions
We have not found information regarding the endo- or exo-
thermic characteristics of the overall TE reaction (3). A calculation
based on the values of the heat of formation of the TE reactants and
products [54] shows a thermal effect of 13kJ/mol for the reaction
(3). This endothermic effect is in agreement with the increasing
triglyceride conversion with increasing TE temperature. DSC stud-
ies of the TE reactions for selected triglyceridemethanol systems
can produce reliable data. The thermal behavior of selected sys-
tems at different molar ratios should be examined for individual
triglyceride and oil/fat feedstocks with methanol.
The heat of the overall TE reaction, ^
r
H(T), can be obtained in
two ways. First, as soon as the C
p,i
(T) functions of reactants and
products are determined, ^
r
H(T) can be calculated by the Kirchhoff
equation [69]:

C
p,

1,p
=

^
r
H
1

p,
=

i
C
p,i
(8)
Inthis relationship, C
p,
is the heat capacity of the systemat con-
stant pressure and extent of reaction (),
i
are the stoichiometric
G. Anitescu, T.J. Bruno / J. of Supercritical Fluids 63 (2012) 133149 141
numbers, and C
p,i
are the partial, molar heat capacities of the reac-
tants and products. Second, the reaction enthalpy can be obtained
from the measured heat ow rates in a DSC by the following equa-
tion:

dQ
dt

p
= C
p,
(1)
d1
dt
+ ^
r
H
d
dt
(9)
Reactions can be conducted quasi isothermally. The reactants are
heatedas rapidlyas possiblefromroomtemperaturetothereaction
temperature. The reaction equilibrium is reached when a constant
heat owrate,
end
, is achieved. Since there is no contributionfrom
heat capacityintheisothermal mode, theaveragereactionenthalpy
is given by
^
r
H =

t
0
(
m

end
)dt

0
d
(10)
3.5. SC-TE process design
Alternative methods to produce biodiesel fuels through the
noncatalytic TE of triglyceride feedstocks in SC alcohols, mainly
methanol and ethanol, have been described (Table 1). Methods
based onSC alcohols were reported as simplifying the TE processes.
However, thetotal manufacturingcost of thesemethods is still high,
due to the large excess of alcohol used (molar ratio of 42:1 versus
stoichiometric 3:1) and glycerolFAME separation issues.
A continuous, one-step, SC uid technology coupled with
power cogeneration reportedly produces biodiesel fuels with-
out the highly complex conventional separation/purication steps
[7,21,5053]. The core of the integrated system consists of the SC-
TE of various triglyceride sources with methanol/ethanol (Fig. 9a).
Some of the reaction products can be combusted in a diesel engine
integrated in the system, which in turn provides the power needed
to pressurize the system and the heat for the SC-TE reactions from
the exhaust gases. The process has the advantage of near-complete
glyceride conversion in less than 10min to fuels with negligi-
ble glycerol content due to its secondary reactions to viable fuel
components (e.g., ethers). Further, the molar ratio of methanol to
triglyceride content in the feedstock is much closer to the stoi-
chiometric amount of 3:1, due to the complete miscibility of lipids
and methanol at SC conditions. A catalytic TE process owsheet is
shown for comparison in Fig. 9b.
Biodiesel fuel productionbythe one-stepSC-TEmethodis a sim-
ple process because of the direct triglyceride andFFAreactions with
alkyl alcohols in a SC state to produce alkyl esters [7,21,5053]. A
high reaction temperature in the proposed method (350400

C)
decomposes some of the unsaturated esters and the byproduct
glycerol further react to valuable fuel components, consequently
improvingthe fuel quality(e.g., lower viscosityandcloudpoint, and
higher volatility). Also, a lowmolar ratio of methanol to triglyceride
has the important effect of eliminating the step of excess alco-
hol recycling, bringing simplicity and cost savings to the method.
The high alkyl ester yield of this SC-TE method is not signicantly
affected by the impurity content (e.g., water and FFAs) of the unre-
ned lipid feedstocks.
When the process schematic diagram of the SC-TE method
shown in Fig. 9a is compared to a current base catalytic industrial
process as shown in Fig. 9b, the simplicity of the former method is
apparent. The simplicity of the SC-TE process is due to low excess
alcohol and glycerol decomposition to valuable fuel components
that eliminates the need for complex separations. In the SC-TE pro-
cess nocatalysts needtoberecovered, andcontinuous owreactors
canbe usedinsteadof batchreactors. All of these processing advan-
tages result fromthe convenient properties of the SCuids involved
in the SC-TE reactions.
Based on the process diagram in Fig. 9a, a SC-TE commercial
plant for biodiesel fuel production was designed with a CHEMCAD
simulation (Fig. 10) [7]. A list of uid properties for each stream in
the process is generated by the modeling software. For a reliable
design, careful attention has to be devoted to the uncertainty of
these properties as well as to the equations used in the modeling.
Of particular importancearetheuiddensities inthereactor, which
will permit the calculationof the residence time, z, of the reactants:
z =
v
q
v0

0
(11)
In Eq. (11), V is the volume of the reactor, q
v0
is the total vol-
umetric ow rate at the reactor inlet, and and
0
are the
triglyceridealcohol mixturedensities at steady-stateinthereactor
and at the reactor inlet, respectively. A list of the CHEMCAD prop-
erties of the uid streams in a pilot plant for one-step biodiesel
production is provided as supporting information (Appendix).
4. Properties of biodiesel fuels produced by SC-TE processes
4.1. Overview
Chemical differences betweenthermallystressedbiodiesel fuels
produced by a high-temperature SC-TE process and those from a
conventional BAC-TE method lead to signicant differences in their
physicochemical properties. These properties ultimately affect fuel
quality vis--vis the engine performance, efciency, combustion
characteristics, and emissions. The fuel quality of FAME mixtures
producedbyaSC-TEprocess is inuencedbyseveral factors, includ-
ing the quality of the feedstock, the FA prole of the parent lipid
feedstock, the processing conditions, especially temperature, other
materials used in the process, and post-production parameters
[1]. Different processing temperatures mainly affect specic fuel
properties, suchas volatility, viscosity, andcold-owbehavior illus-
trated by cloud and pour points.
Books and review articles analyzing the research data of the
most representative studies on biodiesel fuel properties have been
published [1,58,85,86]. While the information on biodiesel fuels
from conventional BAC-TE processing is well documented, the
available properties of the fuels from SC-TE are limited. These
properties are affected to different extents by chain length and
branching, the molecular degree of unsaturation, and the position
andgeometric conguration(cis or trans) of the double bonds. Most
commercial biodiesel fuels are composed mainly of C
16
C
18
chains
of FAMEs, for which the main structural difference is the number
of double bonds. As mentioned earlier, biodiesel fuels produced by
a SC-TE process at temperatures higher than 350

C exhibit sig-
nicant compositional variance, which can signicantly affect their
properties. In particular, the highly unsaturated FAMEs thermally
decompose to smaller C
7
C
15
FAMEs, both saturated and mono-
unsaturated [50,87]. A strong correlation between the degree of
unsaturation and the properties of alkyl esters has been reported
[8890]. With increasing unsaturation, some of the t-for-purpose
fuel property values decrease (heating value, melting point, cetane
number, viscosity, and oxidation stability), while other increase
(density, bulk modulus, fuel lubricity, and iodine value) [90].
4.2. Viscosity
Of the transport properties with reported data for biodiesel
fuels, the viscosity is one of the most sensitive to the compositional
differences [91,92]. While most of the biodiesel fuel properties
compare favorably with those of PDDFs, higher viscosity can affect
the fuel injection parameters, particularly at low engine operat-
ing temperature. The viscosity of commercially produced biodiesel
fuels (4.04.6mm
2
/s at 40

C and 0.1MPa) is outside the allowed


142 G. Anitescu, T.J. Bruno / J. of Supercritical Fluids 63 (2012) 133149
Fig. 9. Process diagramfor (a) thermallystressedbiodiesel fuel productionbyasupercritical transesterication(SC-TE) method[7,21] (top) and(b) pre-treatedalkali-catalyzed
transesterication process. With Elsevier permission from [8] (bottom).
range by ASTM standard D975 for PDDFs (2.53.2mm
2
/s at 40

C)
[91]. Limited experimental data for biodiesel fuels produced by
a SC-TE process at 400425

C show that this property can be


improved through the thermal decomposition of some of the reac-
tion products to smaller molecular fuel components [24,31,52].
4.3. Volatility
Volatility is a critical uid property for fuel replacement
purposes andproper engineoperation. It is verysensitivetocompo-
sitional variabilityof the biodiesel fuel producedbythe TEreactions
and can be determined by the advanced distillation curve (ADC)
technique [9395]. This technique provides temperature and vol-
ume measurements of low uncertainty and composition-explicit
data that allows precise analysis of each fraction of the distilla-
tion. Thefuel compositionversus extent of distillationcantherefore
be determined for a variety of complex fuels, including biodiesel
fuels [9496]. This approach provides VLE measurements of low
uncertainty, and dened thermodynamic state points can be mod-
eled with an appropriate equation of state (EOS). These distillation
curves can be easily compared with those of selected fuels that
are considered typical (e.g., diesel fuel #2 (DF2) and commercial
soybean biodiesel fuel) as shown in Fig. 11 [87].
The experimental results (Fig. 11) show that biodiesel fuel
samples obtained from SC-TE processing exhibit higher volatility
compared to commercial biodiesel fuels produced by a conven-
G. Anitescu, T.J. Bruno / J. of Supercritical Fluids 63 (2012) 133149 143
4
7
8
9
10
3
4
5
6
7
9
1
2
2
5 6
3
1
10
12
8
Oil
Methanol
Biodiesel
11
Fig. 10. CHEMCAD simulation process diagramfor biodiesel fuel production. The properties of each streamare generated by the model for a given set of input conditions [7].
tional catalytic method [87]. This volatility is very close to that of
the DF2 at the start of vaporization, while commercial biodiesel
fuel starts boiling at a temperature higher by more than 100

C.
Biodiesel fuel samples processed at 400

C by the one-step SC-TE


method [21] showed no signicant thermal decomposition over
the full ADC temperature range. In contrast, the distillation curve
of the commercial biodiesel fuel sample exhibits a sharp increase
in temperature at volumetric fractions higher than 60% (vol/vol)
as a result of signicant thermal decomposition [87]. The distil-
lation proles of various test fuels show some differences, which
became more noticeable at the medium-temperature (T50) and
high-temperature (T90). These points were higher for the more
unsaturated fuels. Those increases indistillationtemperatures may
be partly due to the slight differences in chain length and density
among fuels [90].
4.4. Thermal conductivity
Although thermal conductivity is involved in heat-transfer phe-
nomena during SC-TE processing and fuel combustion, available
data are at present limited for biodiesel fuels and their compo-
200
220
240
260
280
300
320
340
360
380
400
420
440
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
L
i
q
u
i
d
T
e
m
p
e
r
a
t
u
r
e
(
o
C
)
BD com
CF_BD
Diesel Fuel
Volume Fracon Dislled
Fig. 11. Distillation curves for biodiesel fuels obtained under supercritical trans-
esterication (SC-TE) conditions from chicken fat (CF-BD) compared to those of
commercial biodiesel fuel (BD com) and diesel fuel #2 [87] (the fuel volume mea-
sured at 25

C and 0.83MPa).
nents [54,57,71]. The thermal conductivity of methyl oleate as a
main component of biodiesel fuels was recently measured over a
wide range of temperature (27327

C) and pressure (0.160MPa)


[97]. The experimental data were used to develop correlations for
thermal conductivity that are estimated to have a relative uncer-
tainty of 4%at a 95%condence level. Work is inprogress ontriolein
as a representative triglyceride for biodiesel fuel processing.
4.5. Density and speed of sound
The densities of the test biodiesel fuels increase with the degree
of unsaturation [90]. Values of this property for neat biodiesel fuels
lay within the range limited in the European standard EN 14214
(860900kg/m
3
) [4]. Other important fuel properties that are more
difcult to measure can be correlated with density. Speed of sound
in a uid is an important property, because it is related to the
isothermal compressibility. This property is used for EOS devel-
opment, and fuel level indicators in many modern aircraft function
via a speed-of-sound measurement. The isentropic bulk moduli of
pure methyl esters and several biodiesel fuels as a function of the
density and speed of sound were obtained at elevated pressures
[98100]. It was found that the speed of sound and the isentropic
bulk modulus of biodiesel fuels tend to increase as the degree of
unsaturation increases.
4.6. Fit-for-purpose properties
Since biodiesel fuels obtained by SC-TE at T <350

C do not differ
signicantly fromthose fromBAC-TE, their t-for-purpose proper-
ties are similar and must meet the ASTM D6751 specications [3].
Several of these properties are discussed below and included in
Table 3 for comparison. For biodiesel fuels obtained at T >350

C,
there are very limited data for most of these properties [31].
4.6.1. Higher heating value (HHV)
Because the HHV values of the most common saturated and
unsaturated long-chain FAMEs vary within a narrow range, the
test fuels as well as commercial biodiesel fuels have similar val-
ues of this property [88]. More detailed insights on the saturated
andunsaturatedbiodiesel fuels throughfuel testing showthat HHV
underwent a slight decrease with the degree of unsaturation [90].
4.6.2. Cetane number
Cetane number is very sensitive to the FAME degree of unsatu-
ration. The variation of the estimated cetane number of test fuels
with their unsaturated index is practically linear [90]. The linearity
between the degree of unsaturation and cetane number shows that
144 G. Anitescu, T.J. Bruno / J. of Supercritical Fluids 63 (2012) 133149
Table 3
Comparative fuel properties linked to fuel combustion quality assessment (averaged values).
Property DF2 B20 B100 SC B100
a
Chemical formula C
14
H
27
C
15
H
29
O
0.4
C
19
H
36
O
2
C
18
H
34
O
2
Molecular weight 195 215 296 282
Carbon (wt.%) 86.9 85.1 77.3 77
Hydrogen (wt.%) 13.1 12.6 11.7 12
Oxygen (wt.%) 0.0 2.1 11.0 11
Specic gravity 0.856 0.862 0.886 0.880
Boiling temperature
b
(

C) 216343 <343 340440 230420


Cetane no. (min 40 required) 43.3 46.0 47.5
Cloud point
c
(

C) 17 14 10 4
Pour point
c
(

C) 21 15 1 0
Flash point (min 52 required) (

C) 62 90 146
Heat of combustion (MJ/kg) 44.0 41.4 37.2
Viscosity
d
(mm
2
/s)/40

C 2.7 2.9 4.1 3.8


a
Biodiesel fuel obtained by SC-TE of chicken fat with methanol [50].
b
The last value is for 90% distilled.
c
Limits are not imposed but the values must be reported.
d
A range of 1.96.0mm
2
/s is required for biodiesel fuels.
biodiesel fuels with an unsaturated index higher than 200 are not
suitable as neat fuels for modern diesel engines. A compendium
of experimental cetane number data including oils and biodiesel
fuels is available in the public domain [101]. A recent paper also
provides the cetane numbers for biodiesel fuels obtain by SC-TE at
400

C [31].
4.7. Properties from computational techniques
In addition to experimental data for biodiesel esters, thermo-
chemical property data estimated by use of various computational
techniques have beenrecently reported[100,102,103]. These prop-
erties include density and speed of sound data of the ve primary
FAMEs found in soybean/rapeseed biodiesel fuels. The data can
be used to develop EOS for the individual FAMEs, and then a
model can be developed to estimate the thermodynamic proper-
ties of ester mixtures. This approach was applied to soy-derived
biodiesel fuels. The predictions for density, speed of sound, and
bubble point were found to agree well with values measured from
two soy biodiesel fuel samples [96]. Studies on physical proper-
ties of biodiesel fuels and individual methyl esters [92,104106]
determinedcritical properties, vapor pressure, heat of vaporization,
density, surface tension, and liquid viscosity.
A wide range of detailed property data for several FAMEs is
contained in the Design Institute for Physical Properties (DIPPR)
database [54,107]. This database includes critical properties, vapor
pressure, liquid density, liquid surface tension, and latent heat of
vaporization, as well as liquid and vapor heat capacities, viscosities,
and thermal conductivities. A brief selection of these properties
is shown in Table 3 along with limited properties determined for
a biodiesel fuel obtained from chicken fat with methanol via the
SC-TE method [50].
4.8. Blended PDDF-biodiesel fuels
Although biodiesel fuels (B100) can be directly combusted
within slightly modied engines, this is not a realistic option. For
reasons of availability, it is likely that biodiesel fuels will be blended
with PDDFs for the foreseeable future. Hence, more attention is
expectedto be devotedtowardthe implications to renery upgrad-
ing. In comparison, a biodiesel fuel contains 94% of the energy
of the same mass of a PDDF, but has a higher cetane number and
lubricity. Typically, 20%(vol/vol) biodiesel fuels are added to PDDFs
to forma B20 blend. B100 and the PDDF must meet their respective
ASTM specications before blending [3]. Selected properties for a
B20 blended fuel are included in Table 3 [108]. Blends of over 20%
biodiesel with diesel fuel should be evaluated on a case-by-case
basis until further experience is available [35].
5. Combustion of biodiesel fuels produced by SC-TE
methods
5.1. Overview
The combustion of biodiesel fuels is signicantly different from
that of PDDFs, due to property differences: higher viscosity, lower
volatility, higher cetane number, higher cloud points, etc. This dif-
ference is due mainly to the oxygen content of the former fuels and
their unsaturated hydrocarbon content. While exhaust emissions
from biodiesel combustion typically have less particulate matter,
due to a higher combustion temperature, this same effect leads to
higher NO
x
content of the exhaust gases. Particular compositions of
biodiesel fuels (higher on saturated palmitic/stearic FAMEs) render
poor behavior of these fuels in cold-weather conditions. However,
PDDF-FAME blends mitigate this disadvantage and also lead to a
more complete combustion, with reduced CO and PM emissions.
In a SC state, any fuel (including biodiesel fuels) will combust
with higher efciency and cleaner exhaust emissions than a fuel
injected as a liquid at near ambient temperatures. The emission
benet of combusting SC fuels is apparent when compared to con-
ventional methods. This advantage mayleadtothe implementation
of a new concept of coupling the SC combustion with the one-
step SC-TE process in a diesel engine, as shown in the last section.
Computational uid-dynamic simulations assist thestudies onow
visualization and extend the study of phenomena for which exper-
imental techniques are either unavailable or limited.
5.2. Spray/jet characteristics and fuel properties
The spray/jet characteristics of fuels injected into a combustion
chamber are major factors affecting combustion phenomena and
hence the engine efciency and emission cleanliness [109]. The
main issue with fuel sprays (for liquid fuels) and jets (for gaseous
and SC fuels) is their level of mixing with air under tight engine
conditions of very short thermodynamic cycles (on a scale of a
few milliseconds). Obviously, in the case of sprays, the mixing
is much lower compared to that in jets. Fuel sprays require a
sequence of complex events such as fuel atomization, heating,
vaporization, diffusion, and reaction with the oxygen in the air
charge. In computational uid-dynamic models for biodiesel
fuel combustion, the physical properties of the liquid phase
needed for fuel spray characterization are density, vapor pressure,
G. Anitescu, T.J. Bruno / J. of Supercritical Fluids 63 (2012) 133149 145
Fig. 12. Spray/jet shadowgraph images of a liquid fuel (left), fuelCO
2
(center) and a supercritical fuelCO
2
(right) injected at 30MPa (upstreampressure) into air at ambient
conditions [114]. The total length of the visual eld is 6.5cm.
surface tension, viscosity, thermal conductivity, heat of vaporiza-
tion, and heat capacity. The properties of the gas-phase species
needed are diffusivity, thermal conductivity, viscosity, and heat
capacity.
Recent research is devoted toward replacing the combustion of
fuel sprays with that of the homogeneous fuelair mixtures such as
in homogeneous charge compression ignition (HCCI) and SC meth-
ods [110113]. Superior ignition-combustion characteristics can
be achieved by injecting the fuel in the form of a SC uid. Fig. 12
shows shadowgraph images of a biodiesel fuel injected in open air
from30MPa and two temperatures (ambient and SC), taken with a
high-speeddigital camera(at 1000frames/s) [114]. Whenthefuel is
injected under SC conditions, it achieves the highest level of mixing
with air upon injection.
The main parameters affecting sprays (the spray penetration
rate and spread angle) are the injector nozzle orice diame-
ter, injection pressure and fuel temperature, and ambient gas
temperature and pressure (density). Both experimental and mod-
eling results are available in the public domain [109111]. The
Ohnesorge number, Oh, is often used to describe the disper-
sion of liquids in gases (e.g., spray technology) [115]. It is
dened as
Oh =

(oL)
1]2
=
viscous forces
(inertia surface tension)
1]2
(12)
Here, is the dynamic viscosity, is the liquid density, o is the
surface tension, and L is the characteristic length scale (typically
drop diameter).
To simulate the injection of a SC fuel into a SC environment, the
species, temperature, and velocity distributions were obtained by
a nite-volume solution of the species, enthalpy, NavierStokes,
and turbulent energy equations [113]. The time-dependent gov-
erning equations written in the cylindrical (z, r) coordinate system
for axisymmetric ow are:

t
+
u
z
+
v
r
r
+
v
r
r
= 0 (13)
]t +u]z +v
r
]r = ]z(1

]z)
+]r(1

]r) v
r
]r +(1

]r)]r +S

(14)
Eq. (13) is the continuity equation, and Eq. (14) represents the
momentum, species, or energy equation depending on the variable
represented by . The symbols u and v
r
are the axial and radial
velocities, respectively, 1

are transport coefcients, and S

are
the source terms of the governing equations. The fuel exiting the
nozzle with an internal diameter I.D. has Reynolds numbers that
are characteristic of turbulent jet ows:
Rc =
u(l.D.)

(15)
The rate of turbulent mass transport is several orders of magnitude
greater than that of the concentration-driven (molecular-diffusive)
mass transport with the mass diffusivity D. Thus, a constant
Schmidt number of unity can be used for simplicity:
Sc =

uD
= 1 (16)
On the basis of the above observations, it would be desirable
to implement an experimental method to determine a fuel qual-
ity by studying its sprays/jets. Of particular interest will be to map
the fuel concentration gradient in the air upon injection. This is,
however, a difcult task, due to the harsh conditions in combus-
tion chambers. An easier task would be to determine the spray/jet
cone angle of the fuels injected into a controlled atmosphere. As
shown in Fig. 12, a larger spray/jet cone angle indicates a better
fuelair mixingprocess. This methodcouldcomplement thevolatil-
ity property of fuel determined by the advanced distillation curve
method.
5.3. Biodiesel fuel properties and combustion quality
The quality of fuel combustion is obviously closely connected
with fuel properties. Under the conditions of an engine cycle, the
fuel has to followa complex sequence of events fromfuel injection
to the evacuation of the exhaust gases. The fuel properties needed
include volatility, viscosity, thermal stability, reactivity with air,
ignition delay, ignition temperature, heat of combustion, density,
critical PT, thermal conductivity, rate of heat release, andC
p
. Other
physical properties that are neededtodevelopmodels for fuel com-
bustion are transport properties (viscosity, thermal conductivity,
species diffusivity) of the chemical species in the gas phase. The
most important of these gases are the fuel components and O
2
,
along with the intermediate and nal products, such as OH

, H

, O

,
CO
2
and H
2
O, of the combustion reaction [116]. Recent progress in
improving knowledge of the physical properties of surrogate com-
ponents, their mixtures and real fuels is discussed in the literature
[58,117].
The inuence of biodiesel fuel properties on injection, spray,
and engine characteristics with the aim to reduce harm-
ful emissions has been extensively discussed in the literature
[1,58,90,111,118,119]. Most of the SC-TE reaction products were
146 G. Anitescu, T.J. Bruno / J. of Supercritical Fluids 63 (2012) 133149
Fig. 13. Conceptual design of a retrotted diesel engine running on supercritical biodiesel fuel produced in situ by supercritical transesterication (SC-TE) reactions [53].
rened in order to meet the required specications for biodiesel
fuels. As a result of this post-processing step, the resulting fuels
were close to those obtained by the BAC-TE methods. Standing
apart fromthese laborious post-processed fuels are those obtained
at 400

C due to the thermal decomposition of the polyunsatu-


rated FAMEs, mainly to lower molecular esters and hydrocarbons
[21,24,5053]. Although there are no experimental data yet for a
comprehensive characterization of these fuels, the knowledge in
the eld points out to higher combustion efciency and cleaner
emissions in this case.
5.4. In-engine biodiesel fuel generation and SC combustion
Vegetable oil feedstocks exhibit better behavior at lowtempera-
tures than that of most of biodiesel fuels. The transition fromliquid
tosolidstates occurs at lower temperatures for theformer, whilefor
the latter solidication occurs at higher temperatures and includes
a two-phase heterogeneous transition [1]. This different behavior
is due to the uniform distribution of the saturated and unsatu-
rated FAs in triglyceride molecules in the feedstocks. In cooling of
biodiesel fuels, the saturatedandunsaturatedFAMEs are separated,
with the former solidifying rst.
Based on the above observation, a new method was designed
to inject into combustion chambers the SC-TE products that were
generated in a retrotted engine fuel system(Fig. 13) [53,112]. The
SC-TE products could, in this case, include higher concentrations of
water andglycerol thanthoseimposedbycurrent standards. Theoil
or the alcohol could be preheated by being used as a coolant for the
engine block. The uid properties needed for alcohol, lipid feed-
stock, oilalcohol mixtures, and SC-TE reaction products include
critical PT properties, density, viscosity, Reynolds number, heat
capacity, thermal conductivity, solubility, thermal stability, and
volatility. This approach could solve the emission problems asso-
ciated with the conventional combustion while improving engine
efciency and TE-associated costs.
6. Summary
The connection between uid properties and process-
ing/combustion, leading to optimized steps of biodiesel fuel
production and combustion, is discussed. The review provides
information and state-of-the-art knowledge on feedstock proper-
ties, SC-TE reaction characteristics, t-for-purpose properties of
biodiesel fuels, and SC injection/combustion. Superior results with
SC-TE processing/combustion are outlined and compared to con-
ventional methods. The support for this statement is presented in
a ChemCad application of the SC-TE process. Emerging knowledge
on uid properties involved in SC-TE reactions and SC combustion
is outlined.
Despitetherecent progress, muchremains unknownabout non-
catalytic SC-TE processes. The reaction mechanism, pathways, and
detailed kinetics have not yet been resolved. The phase behavior
of this reacting system is extremely important, and recent litera-
ture is focused on this topic [21,7280]. In early SC-TE studies, very
high alcohol-to-oil ratios are used to ensure the existence of a sin-
gle phase at reaction conditions. With a better understanding of
the phase behavior, one could reduce the amount of excess alcohol
used and thereby reduce the processing cost. Moreover, advanced
understanding of the phase behavior would allow adequate reac-
tion models to be developed.
Several directions for newor additional researchthat wouldlead
to important advances in this eld are suggested. Among them,
a better understanding of the reaction kinetics and phase behav-
ior (e.g., mapping liquid-liquid-vapor, liquid-vapor, and SC regions
during the TE reactions) is needed. There are critical needs for
uid properties versus biodiesel fuel processing (properties versus
conditions, reaction, and reactor design) and combustion (prop-
erties versus spray characteristics, burning efciency, pollution,
etc.). The SC-TE process requires higher temperatures, pressures,
and alcohol-to-oil ratios. Consequently, a process well engineered
for unspent reactant recycle and energy recovery is a prerequisite.
There is a needfor additional work onthe comparative process eco-
nomics so that the precise tradeoffs are better understood. Finally,
a new conceptual design of in situ generation of biodiesel fuels
coupled with SC fuel injection and combustion is presented. Such
designs may be able to mitigate emissions of conventional combus-
tion, while improving engine efciency and TE-associated costs.
Abbreviation list
^
r
H(T) heat of reaction (reaction enthalpy)
^
c
H

heat of combustion
C
p,i
constant-pressure molar heat capacity of a uid i
A, B, C uid constants in Eq. (12)
ADC advanced distillation curve
ASTM American Society for Testing and Materials
B100 neat biodiesel fuel (100% biodiesel)
B20 blend of 20% biodiesel fuel and 80% petroleum diesel fuel
(vol/vol)
c
i
specic heat capacity of a uid i
CN cetane number
CP cloud point
DF2 diesel fuel #2
DG diglyceride
DIPPR Design Institute for Physical Properties
G. Anitescu, T.J. Bruno / J. of Supercritical Fluids 63 (2012) 133149 147
DSC differential scanning calorimetry
E
a
energy of activation
EOS equation of state
FA fatty acid
FAMEs fatty acid methyl esters
FAEEs fatty acid ethyl esters
FFA free fatty acid
FID ame ionization detector
FP ash point
GC gas chromatography
GP glycerol products
HC hydrocarbon(s)
HCCI homogeneous charge compression ignition
L liquid phase
m mass
MG monoglyceride
MSD mass selective detector
NIST National Institute of Standards and Technology
NO
x
nitrogen oxides (noxes)
NRTL non-random two-liquid (thermodynamic model)
PM particulate matter
PP pour point
PR PengRobinson
PVT pressurevolumetemperature
Q heat
q
v0
total volumetric ow rate at the reactor inlet
REFPROP NIST Reference Fluid Thermodynamic and Transport
Properties database (Standard Reference Database 23)
SC supercritical
SC-TE supercritical transesterication
TDC top dead center (the upmost position of an engine piston)
TDE NIST ThermoData Engine (Standard Reference Database
103)
TE transesterication
TG triglyceride
UHC unburned hydrocarbons
V reactor volume, vapor phase
VLE vaporliquid equilibrium
x mass fraction
X mole fraction
Greek symbols
average heat ow rate;
q kinematic viscosity
dynamic viscosity

i
stoichiometric numbers
extent of the reaction
,
0
uid density
z residential time
heat ow rate
Appendix A. Supplementary data
Supplementary data associated with this article can be found, in
the online version, at doi:10.1016/j.supu.2011.11.020.
References
[1] G. Knothe, J. van Gerpen, J. Krahl, The Biodiesel Handbook, American Oil
Chemists Society (AOCS) Press, Champaign, IL, 2005.
[2] L.T.T. Dinh, Y. Guo, M.S. Mannan, Sustainability evaluation of biodiesel
production using multicriteria decision-making, Environmental Progress &
Sustainable Energy 28 (2009) 3846.
[3] ASTM-International, Standard Specication for Biodiesel Fuel Blendstock
(B100) for Middle Distillate Fuels. ASTM Int. Specif. ASTM D6751, 2008.
[4] http://www.worldenergy.net/products/biodiesel/us specs.php (retrieved
02.06.11).
[5] http://www.epa.gov/otaq/fuels/renewablefuels/index.htm (retrieved
02.06.11).
[6] J. Hill, E. Nelson, D. Tilman, S. Polasky, D. Tiffany, Environmental, economic,
and energetic costs and benets of biodiesel and ethanol biodiesels, Proceed-
ings of the National Academy of Sciences of the United States of America 103
(2006) 1120611210.
[7] A. Deshpande, G. Anitescu, P.A. Rice, L.L. Tavlarides, Supercritical biodiesel
production and power cogeneration: technical and economic feasibilities,
Bioresource Technology 101 (2010) 18341843.
[8] A.H. West, D. Posarac, N. Ellis, Assessment of four biodiesel production pro-
cesses using HYSYS Plant, Bioresource Technology 99 (2008) 65876601.
[9] Y. Zhang, M.A. Dube, D.D. Mclean, M. Kates, Biodiesel production from waste
cooking oil: 2. Economic assessment and sensitivity analysis, Bioresource
Technology 90 (2003) 229240.
[10] M.J. Haas, A.J. McAloon, W.C. Yee, T.A. Foglia, A process model to estimate
biodiesel production costs, Bioresource Technology 97 (2006) 671678.
[11] J.M.N.V. Kasteren, A.P. Nisworo, A process model to estimate the cost of
industrial scale biodiesel production from waste cooking oil by supercrit-
ical transesterication, Resources Conservation and Recycling 50 (2007)
442458.
[12] A. Elbheri, W. Coyle, E. Dohlman, H. Kmak, J. Ferrell, Z. Haq, J.
Houghton, B. Stokes, M. Buford, W. Nieh, M. Patton-Mallory, D. Perla,
L. Draucker, R. Dodder, P.O. Kaplan, D. Okamuro, J. Steiner, H. Shapouri,
D. OToole, The Economics of Biomass Feedstocks in the United States,
Biomass Research and Development Board, Technical Report, 2008, October,
http://www.usbiomassboard.gov/pdfs/7 Feedstocks Literature Review.pdf.
[13] K. Arai, R.L. Smith Jr., T.M. Aida, Decentralized chemical processes with super-
critical uid technology for sustainable society, Journal of Supercritical Fluids
47 (2009) 628636.
[14] M. Poliakoff, P. Licence, Sustainable technology: green chemistry, Nature 450
(2007) 810812.
[15] M.O. McLinden, T.J. Bruno, M. Frenkel, M.L. Huber, Standard reference data for
the thermophysical properties of biofuels, ASTMInternational 7 (2010) 118.
[16] G. Anitescu, T.J. Bruno, Liquid biofuels: integrating feedstock selection, pro-
cessing, rening, blending, storage and combustion with uid properties a
review, Energy & Fuels, (2011) doi:10.1021/ef201392s.
[17] T. Pinnarat, P.E. Savage, Assessment of noncatalytic biodiesel synthesis using
supercritical reaction conditions, Industrial Engineering Chemistry Research
47 (2008) 68016808.
[18] R. Sawangkeaw, K. Bunyakiat, S. Ngamprasertsith, A review of laboratory-
scale research on lipid conversion to biodiesel with supercritical methanol
(20012009) review article, Journal of Supercritical Fluids 55 (2010) 113.
[19] J-S. Lee, S. Saka, Biodiesel production by heterogeneous catalysts and super-
critical technologies, Bioresource Technology 101 (2010) 71917200.
[20] S. Glisic, D. Skala, The problems in design and detailed analyses of energy
consumption for biodiesel synthesis at supercritical conditions, Journal of
Supercritical Fluids 49 (2009) 293301.
[21] G. Anitescu, A. Deshpande, L.L. Tavlarides, Integratedtechnologyfor supercrit-
ical biodiesel production and power cogeneration, Energy & Fuels 22 (2008)
13911399.
[22] P. Olivares-Carrillo, J. Quesada-Medina, Synthesis of biodiesel from soybean
oil using supercritical methanol in a one-step catalyst-free process in batch
reactor, Journal of Supercritical Fluids 58 (2011) 378384.
[23] K. Bunyakiat, S. Makmee, R. Sawangkeaw, S. Ngamprasertsith, Continuous
production of biodiesel via transesterication from vegetable oils in super-
critical methanol, Energy & Fuels 20 (2006) 812817.
[24] W. Iijima, Y. Kobayashi, K. Takekura, K. Taniwaki, The non-glycerol process of
biodiesel fuel treated in supercritical methanol, ASAE/CSAE Annual Interna-
tional Meeting, Ottawa, Ontario, Canada, August 14, 2004, Paper no. 046073.
[25] N. Aimaretti, D.L. Manuale, V.M. Mazzieri, C.R. Vera, J.C. Yori, Batch study
of glycerol decomposition in one-stage supercritical production of biodiesel,
Energy & Fuels 23 (2009) 10761080.
[26] W.B. Schulte, Biodiesel production from tall oil and chicken fat via supercrit-
ical methanol treatment, Master thesis, University of Arkansas (2007).
[27] H. He, T. Wang, S. Zhu, Continuous productionof biodiesel fuel fromvegetable
oil using supercritical methanol process, Fuel 86 (2007) 442447.
[28] M.N. Varma, G. Madras, Synthesis of biodiesel from castor oil and linseed oil
in supercritical uids, Industrial Engineering Chemistry Research 46 (2007)
16.
[29] I. Vieitez, C. da Silva, G.R. Borges, F.C. Corazza, J.V. Oliveira, M.A. Grompone,
I. Jachmanian, Continuous production of soybean biodiesel in supercritical
ethanolwater mixtures, Energy & Fuels 22 (2008) 28052809.
[30] H. Han, W. Cao, J. Zhang, Preparationof biodiesel fromsoybeanoil usingsuper-
critical methanol and CO
2
as a co-solvent, Process Biochemistry 40 (2005)
31483151.
[31] R. Sawangkeaw, S. Teeravitud, K. Bunyakiat, S. Ngamprasertsith, Biofuel pro-
duction from palm oil with supercritical alcohols: effects of the alcohol to oil
molar ratios on the biofuel chemical composition and properties, Bioresource
Technology 102 (2011) 1070410710.
[32] G. Madras, C. Kolluru, R. Kumar, Synthesis of biodiesel in supercritical uids,
Fuel 83 (2004) 20292033.
[33] P. Hegel, G. Mabe, S. Pereda, E.A. Brignole, Phase transitions ina biodiesel reac-
tor using supercritical methanol, Industrial Engineering Chemistry Research
46 (2007) 63606365.
148 G. Anitescu, T.J. Bruno / J. of Supercritical Fluids 63 (2012) 133149
[34] C.-S. Choi, J.-W. Kim, C.-J. Jeong, H. Kim, K.-P. Yoo, Transesterication kinetics
of palm olein oil using supercritical methanol, Journal of Supercritical Fluids
58 (2011) 365370.
[35] J.Z. Yin, M. Xiao, J.B. Song, Biodiesel fromsoybeanoil insupercritical methanol
with co-solvent, Energy Conversion and Management 49 (2008) 908912.
[36] C.W. Wang, J.F. Zhou, W. Chen, W.G. Wang, Y.X. Wu, J.F. Zhang, R.A. Chi, W.Y.
Ying, Effect of weak acids as a catalyst on the transesterication of soybean
oil in supercritical methanol, Energy & Fuels 22 (2008) 34793483.
[37] V. Rathore, G. Madras, Synthesis of biodiesel from edible and non-edible oils
insupercritical alcohols andenzymatic synthesis insupercritical carbondiox-
ide, Fuel 86 (2007) 26502659.
[38] R. Sawangkeaw, K. Bunyakiat, S. Ngamprasertsith, Effect of co-solvents on
production of biodiesel via transesterication in supercritical methanol,
Green Chemistry 9 (2007) 679685.
[39] E-S. Song, J-W. Lim, H-S. Lee, Y-W. Lee, Transesterication of RBD palm
oil using supercritical methanol, Journal of Supercritical Fluids 44 (2008)
356363.
[40] S. Hawash, N. Kamal, F. Zaher, O. Kenawi, G. Diwani, Biodiesel fuel from Jat-
ropha oil via non-catalytic supercritical methanol transesterication, Fuel 88
(2009) 579582.
[41] C. Silva, T. Weschenfelder, S. Rovani, F. Corazza, M. Corazza, C. Dariva, J.
Oliveira, Continuous production of fatty acid ethyl esters from soybean oil
in compressed ethanol, Industrial Engineering Chemistry Research 46 (2007)
53045309.
[42] S. Saka, D. Kusdiana, Biodiesel fuel from rapeseed oil as prepared in super-
critical methanol, Fuel 80 (2001) 225231.
[43] P.D. Patil, V.G. Gude, S. Deng, Transesterication of camelina sativa oil using
supercritical and subcritical methanol with cosolvents, Energy & Fuels 24
(2010) 746751.
[44] M.N. Varma, P.A. Deshpande, G. Madras, Synthesis of biodiesel insupercritical
alcohols and supercritical carbon dioxide, Fuel 89 (2010) 16411646.
[45] S. Lim, S-S. Hoong, L-K. Teong, S. Bhatia, Supercritical uid reactive extrac-
tion of Jatropha curcas L. seeds with methanol: a novel biodiesel production
method, Bioresource Technology 101 (2010) 71697172.
[46] P. Valle, A. Velez, P. Hegel, G. Mabe, E.A. Brignole, Biodiesel production using
supercritical alcohols with a non-edible vegetable oil in a batch reactor, Jour-
nal of Supercritical Fluids 54 (2010) 6170.
[47] I. Vieitez, M.J. Pardo, C. da Silva, C. Bertoldi, F. de Castilhos, J.V. Oliveira,
M.A. Grompone, I. Jachmanian, Continuous synthesis of castor oil ethyl esters
under supercritical ethanol, Journal of Supercritical Fluids 56(2011) 271276.
[48] C.M. Trentin, A.P. Lima, I.P. Alkimim, C. da Silva, F. de Castilhos, M.A. Mazutti,
J.V. Oliveira, Continuous catalyst-free production of fatty acid ethyl esters
from soybean oil in microtube reactor using supercritical carbon dioxide as
co-solvent, Journal of Supercritical Fluids 56 (2011) 283291.
[49] C-Y. Lin, C-L. Fan, Fuel properties of biodiesel produced fromCamellia oleifera
Abel oil through supercriticalmethanol transesterication, Fuel 90 (2011)
22402244.
[50] V.F. Marulanda, G. Anitescu, L.L. Tavlarides, Biodiesel fuels through a contin-
uous ow process of chicken fat supercritical transesterication, Energy &
Fuels 24 (2010) 253260.
[51] T. Cong, G. Anitescu, L.L. Tavlarides, Supercritical transesterication of tri-
olein as model compound for biodiesel synthesis: a kinetic study, (2012) in
preparation.
[52] V. Marulanda, G. Anitescu, L.L. Tavlarides, Investigations on supercritical
transesterication of chicken fat for biodiesel production fromlow-cost lipid
feedstocks, Journal of Supercritical Fluids 54 (2010) 5360.
[53] G. Anitescu, Supercritical uidtechnology appliedtothe productionandcom-
bustion of diesel and biodiesel fuels, Ph.D. thesis, Syracuse University (2008).
[54] R.L. Rowley, W.V. Wilding, J.L. Oscarson, N.A. Zundel, T.L. Marshall, T.E.
Daubert, R.P. Danner, DIPPR Data Compilation of Pure Compound Proper-
ties, Design Institute for Physical Properties, New York, NY, 2004 (retrieved
08.06.11) http://www.aiche.org/dippr/.
[55] http://aspentech.com/products/aspen-plus.aspx (retrieved 19.05.11).
[56] CHEMCAD Version 5.6.4, Chemstations Inc., Houston, TX, 2007, http://www.
chemstations.com/content/documents/Technical Articles/BiodieselWhite
Paper.pdf.
[57] E.W. Lemmon, M.L. Huber, M.O. McLinden, NIST Standard Reference Database
23, Reference Fluid Thermodynamic and Transport Properties REFPROP,
Version9.0, StandardReference Data Program, National Institute of Standards
and Technology, Gaithersburg, MD, 2010.
[58] W.J. Pitz, C.J. Mueller, Recent progress in the development of diesel surrogate
fuels, Progress in Energy and Combustion Science 37 (2011) 330350.
[59] M.L. Huber, E. Lemmon, A. Kazakov, L.S. Ott, T.J. Bruno, Model for the thermo-
dynamic properties of a biodiesel fuel, Energy & Fuels 37 (2009) 37903797.
[60] G. Mowry, A portable, automated and environmentally friendly biodiesel
processing system, International Journal of Sustainable Energy (2011),
doi:10.1080/1478646X.2010.542816.
[61] J. Satyarthi, D. Srinivas, P. Ratnasamy, Estimation of free fatty acid content in
oils, fats, and biodiesel by
1
H NMR spectroscopy, Energy & Fuels 23 (2009)
22732277.
[62] W. Chen, C. Wang, W. Ying, W. Wang, Y. Wu, J. Zhang, Continuous production
of biodiesel via supercritical methanol transesterication in a tubular reactor.
Part 1: Thermophysical and transitive properties of supercritical methanol,
Energy & Fuels 23 (2009) 526532.
[63] P.T. Pienkos, A. Darzins, The promise and challenges of microalgal-derived
biofuels, Biofuels, Bioproducts & Biorening 23 (2009) 43440.
[64] A. Singh, S.I. Olsen, A critical review of biochemical conversion, sustainabil-
ity and life cycle assessment of algal biofuels, Applied Energy 88 (2011)
35483555.
[65] A.F. Clarens, E.P. Resurreccion, M.A. White, L.M. Colosi, Environmental life
cycle comparison of algae to other bioenergy feedstocks, Environmental Sci-
ence & Technology 44 (2010) 18131819.
[66] R.B. Levine, T. Pinnarat, P.E. Savage, Biodiesel production from wet algal
biomass through in situ lipid hydrolysis and supercritical transesterication,
Energy & Fuels 24 (2010) 52355243.
[67] R.F. Service, Algaes second try, Science 333 (2011) 12381239.
[68] G. Knothe, Designer biodiesel: optimizing fatty ester composition to improve
fuel properties, Energy & Fuels 22 (2008) 13581364.
[69] G.W.H. Hhne, W.F. Hemminger, H.-J. Flammersheim, Differential Scanning
Calorimetry, 2nd ed., Springer-Verlag, Berlin/Heidelberg, 2003.
[70] H.W. Xiang, A. Laesecke, M.L. Huber, A new reference correlation for the vis-
cosity of methanol, Journal of Physical Chemistry Reference Data 35 (2006)
15971620.
[71] http://www.chempro.in/palmoilproperties.htm.
[72] D.G.B. Boocock, S.K. Konar, H. Sidi, Phase diagrams for oil/methanol/ether
mixtures, Journal of the American Oil Chemists Society 73 (1996) 12471251.
[73] W. Zhou, S.K. Konar, D.G.B. Boocock, Ethyl esters from the single-phase base-
catalyzed ethanolysis of vegetable oils, Journal of the American Oil Chemists
Society 80 (2003) 367371.
[74] W. Zhou, D.G.B. Boocock, Phase behavior of the base-catalyzed transesteri-
cation of soybean oil, Journal of the American Oil Chemists Society 83 (2006)
10411045.
[75] S. Glisic, O. Montoya, A. Orlovic, D. Skala, Vaporliquid equilibria of
triglyceridesmethanol mixtures and their inuence on the biodiesel synthe-
sis under supercritical conditions of methanol, Journal of Serbian Chemical
Society 72 (2007) 1327.
[76] S. Glisic, D. Skala, Phase transition at subcritical and supercritical conditions
of triglycerides methanolysis, Journal of Supercritical Fluids 54 (2010) 7180.
[77] H. Zhou, H. Lu, B. Liang, Solubility of multicomponent systems in the biodiesel
production by transesterication of Jatropha curcas L. oil with methanol,
Journal of Chemical and Engineering Data 51 (2006) 11301135.
[78] D. Geana, R. Steiner, Calculation of phase equilibrium in supercritical extrac-
tion of C54 triglyceride (rapeseed oil), Journal of Supercritical Fluids 8 (1995)
107118.
[79] T. Tsuji, M. Kubo, N. Shibasaki-Kitakawa, T. Yonemoto, Is excess methanol
addition required to drive transesterication of triglyceride toward complete
conversion? Energy & Fuels 23 (2009) 61636167.
[80] Z. Tang, Z. Du, E. Min, L. Gao, T. Jiang, B. Han, Phase equilibria of
methanoltriolein system at elevated temperature and pressure, Fluid Phase
Equilibria 239 (2006) 811.
[81] G.M. Acosta, R.L. Smith Jr., K. Arai, High-pressure PVT behavior of natural fats
and oils, trilaurin, triolein, and n-tridecane from 303K to 353K from atmo-
spheric pressure to 150MPa, Journal of Chemical and Engineering Data 41
(1996) 961969.
[82] R.J. Goldstein, Fluid Mechanics Measurements, Hemisphere, Washington DC,
1983.
[83] G. Anitescu, L.L. Tavlarides, D. Geana, Phase transitions and thermal behavior
of fueldiluent mixtures, Energy & Fuels 23 (2009) 30683077.
[84] H. He, S. Sun, T. Wang, S. Zhu, Transesterication kinetics of soybean oil for
production of biodiesel in supercritical methanol, Journal of the American Oil
Chemists Society 84 (2007) 399404.
[85] M.S. Graboski, R.L. McCormick, Combustion of fat and vegetable oil derived
fuels in diesel engines, Progress in Energy and Combustion Science 24 (1998)
125164.
[86] M. Lapuerta, O. Armas, J. Rodriguez-Fernandez, Effect of biodiesel fuels on
diesel engine emissions, Progress in Energy and Combustion Science 34
(2008) 198223.
[87] G. Anitescu, T.J. Bruno, Characterization of the biodiesel fuel obtained by
supercritical uid processing by the advanced distillation curve method,
(2012) in preparation.
[88] G. Knothe, Dependence of biodiesel fuel properties on the structure of fatty
acid alkyl esters, Fuel Processing Technology 86 (2005) 10591070.
[89] A.A. Refaat, Correlation between the chemical structure of biodiesel and its
physical properties, International Journal of Environmental Science andTech-
nology 6 (2009) 677694.
[90] P. Benjumea, J.R. Agudelo, A.F. Agudelo, Effect of the degree of unsaturation
of biodiesel fuels on engine performance, combustion characteristics, and
emissions, Energy & Fuels 25 (2011) 7785.
[91] M.E. Tat, J.H. van Gerpen, The kinematic viscosity of biodiesel and its blends
with diesel fuel, Journal of the American Oil Chemists Society 76 (1999)
15111513.
[92] W. Yuan, A.C. Hansen, Q. Zhang, Predicting the temperature dependent vis-
cosity of biodiesel fuels, Fuel 88 (2009) 11201126.
[93] T.J. Bruno, Improvements in the measurement of distillation curves part 1:
a composition-explicit approach, Industrial Engineering Chemistry Research
45 (2006) 43714380.
[94] L.S. Ott, T.J. Bruno, Variability of biodiesel fuel and comparison to petroleum-
derived diesel fuel: application of a composition and enthalpy explicit
distillation curve method, Energy & Fuels 22 (2008) 28612868.
[95] B.L. Smith, L.S. Ott, T.J. Bruno, Composition-explicit distillation curves of com-
mercial biodiesel fuels: comparison of petroleum derived fuel with B20 and
B100, Industrial Engineering Chemistry Research 47 (2008) 58325840.
G. Anitescu, T.J. Bruno / J. of Supercritical Fluids 63 (2012) 133149 149
[96] M.L. Huber, E.W. Lemmon, A. Kazakov, L.S. Ott, T.J. Bruno, Model for the
thermodynamic properties of a biodiesel fuel, Energy & Fuels 23 (2009)
37903797.
[97] R.A. Perkins, M.L. Huber, Measurement and correlation of the thermal con-
ductivities of biodiesel constituent uids: methyl oleate andmethyl linoleate,
Energy & Fuels 25 (2011) 23832388.
[98] M.E. Tat, J.H. van Gerpen, S. Soylu, M. Canakci, A. Monyem, S. Wormley, The
speed of sound and isentropic bulk modulus of biodiesel at 21

C from atmo-
spheric pressure to 35MPa, Journal of the American Oil Chemists Society 77
(2000) 285289.
[99] M.E. Tat, J.H. van Gerpen, Speed of sound and isentropic bulk modulus of alkyl
monoesters at elevated temperatures and pressures, Journal of the American
Oil Chemists Society 80 (2003) 12491256.
[100] A. Osmont, L. Catoire, P. Dagaut, Thermodynamic data for the modeling of
the thermal decomposition of biodiesel 1. Saturated and monounsaturated
FAMEs, Journal of Physical Chemistry A 114 (2009) 37883795.
[101] M.J. Murphy, J.D. Taylor, R.L. McCormick, Compendium of Experimental
Cetane Number Data. Technical Report NREL/SR-540-36805, National Renew-
able Energy Laboratory, 2004, September.
[102] M. Lapuerta, J. Rodrguez-Fernndez, F. Oliva, Determination of enthalpy of
formation of methyl and ethyl esters of fatty acids, Chemistry and Physics of
Lipids 163 (2010) 172181.
[103] L.S. Ott, M.L. Huber, T.J. Bruno, Density and speed of sound measure-
ments on ve fatty acid methyl esters at 83kPa and temperatures from
(278.15 to 338.15) K, Journal of Chemical and Engineering Data 53 (2008)
24122416.
[104] W. Yuan, A.C. Hansen, Q. Zhang, Predicting the physical properties of
biodiesel for combustion modeling, Transactions of the ASAE 46 (2003)
14871493.
[105] W.Q. Yuan, A.C. Hansen, Q. Zhang, Z.C. Tan, Temperature-dependent kine-
matic viscosity of selected biodiesel fuels and blends with diesel fuel, Journal
of the American Oil Chemists Society 82 (2005) 195199.
[106] W. Yuan, A.C. Hansen, Q. Zhang, Vapor pressure and normal boiling point
predictions for pure methyl esters and biodiesel fuels, Fuel 84 (2005)
943950.
[107] T.E. Daubert, R.P. Danner, H.M. Sibul, C.C. Stebbins, R.L. Rowley, W.V.
Wilding, Physical and Thermodynamic Properties of Pure Chemicals, 2011,
http://dippr.byu.edu/.
[108] National Biodiesel Board, Specication for Biodiesel Blends B6-B20,
ASTM 7467-10, 2011 (retrieved 17.08.11) http://www.biodiesel.org/pdf
les/fuelfactsheets/B20 Specication.pdf.
[109] L.M. Pickett, C.L. Genzale, G. Bruneaux, L.-M. Malbec, L. Hermant, C. Chris-
tiansen, J. Schramm, Comparison of diesel spray combustion in different
high-temperature, high-pressure facilities, SAE International 2010-01-2106,
140.
[110] R. Reitz, D. Arcoumanis, T. Kamimoto, B. Johannson, Special Issue on homoge-
neous charge compression ignition engines, International Journal of Engine
Research 8 (2007) 13.
[111] S. Um, S.W. Park, Numerical studyoncombustionandemissioncharacteristics
of homogeneous charge compression ignition engines fueled with biodiesel,
Energy & Fuels 24 (2010) 916927.
[112] G. Anitescu, R-H. Lin, L.L. Tavlarides, Preparation, injection and combustion
of supercritical fuels; Poster P-2 presented at Directions in Engine-Efciency
and Emissions Research (DEER) Conference, Dearborn, MI, August 36, 2009.
[113] T. Doungthip, J.S. Ervin, T.F. Williams, J. Bento, Studies of injection of jet fuel at
supercritical conditions, Industrial EngineeringChemistryResearch41(2002)
58565866.
[114] G. Anitescu, Syracuse University, unpublished data.
[115] W. Ohnesorge, Formation of drops by nozzles and the breakup of liquid jets,
Journal of Applied Mathematics and Mechanics 16 (1936) 355358.
[116] A.T. Holley, X.Q. You, E. Dames, H. Wang, F.N. Egolfopoulos, Sensitivity of
propagation and extinction of large hydrocarbon ames to fuel diffusion,
Proceedings of the Combustion Institute 32 (2009) 11571163.
[117] O. Herbinet, W.J. Pitz, C.K. Westbrook, Detailed chemical kinetic mechanism
for the oxidation of biodiesel fuels blend surrogate, Combustion and Flame
157 (2010) 893908.
[118] B. Kegl, Effects of biodiesel on emissions of a bus diesel engine, Bioresource
Technology 99 (2008) 863873.
[119] H.J. Curran, P. Gafuri, W.J. Pitz, C.K. Westbrook, A comprehensive modeling
study on n-heptane oxidation, Combustion and Flame 114 (1998) 149177.

Das könnte Ihnen auch gefallen