Você está na página 1de 15

Available online at www.sciencedirect.

com

Bioresource Technology 99 (2008) 6587–6601

Assessment of four biodiesel production processes using HYSYS.Plant


Alex H. West, Dusko Posarac, Naoko Ellis *
Department of Chemical and Biological Engineering, The University of British Columbia, 2360 East Mall, Vancouver, BC, Canada V6T 1Z3

Received 24 February 2006; received in revised form 31 October 2007; accepted 3 November 2007
Available online 29 January 2008

Abstract

Four continuous biodiesel processes were designed and simulated in HYSYS. The first two employed traditional homogeneous alkali
and acid catalysts. The third and fourth processes used a heterogeneous acid catalyst and a supercritical method to convert a waste veg-
etable oil feedstock into biodiesel. While all four processes were capable of producing biodiesel at high purity, the heterogeneous and
supercritical processes were the least complex and had the smallest number of unit operations. Material and energy flows, as well as sized
unit operation blocks, were used to conduct an economic assessment of each process. Total capital investment, total manufacturing cost
and after tax rate-of-return were calculated for each process. The heterogeneous acid catalyst process had the lowest total capital invest-
ment and manufacturing costs, and had the only positive after tax rate-of-return.
Ó 2007 Elsevier Ltd. All rights reserved.

Keywords: Biodiesel; Process design; Process simulation; Economic assessment

1. Introduction and background applications in highly sensitive environments, such as mar-


ine ecosystems and mining enclosures.
Recent concerns over diminishing fossil fuel supplies As shown in Eq. (1), biodiesel (alkyl ester) is usually
and rising oil prices, as well as adverse environmental produced by the transesterification of a lipid feedstock.
and human health impacts from the use of petroleum fuel Transesterification is the reversible reaction of a fat or oil
have prompted considerable interest in research and devel- (which is composed of tri-glycerides) with an alcohol to
opment of fuels from renewable resources, such as biodiesel form fatty acid alkyl esters and glycerol. Stoichio-
and ethanol. Biodiesel is a very attractive alternative fuel, metrically, the reaction requires a 3:1 molar alcohol-to-oil
as it is derived from a renewable, domestic resource and ratio, but excess alcohol is usually added to drive the equi-
can, therefore, reduce reliance on foreign petroleum librium toward the products side
imports. Biodiesel reduces net carbon dioxide emissions
by 78% on a life-cycle basis when compared to conven- CH2–OOC–R1 R1–COO–R’ CH2–OH
tional diesel fuel (Tyson, 2001). It has also been shown to | Catalyst |
have dramatic improvements on engine exhaust emissions. CH–OOC–R2 + 3R’OH ⇔ R2–COO–R’ + CH–OH
For instance, combustion of neat biodiesel decreases car-
bon monoxide (CO) emissions by 46.7%, particulate matter | |
emissions by 66.7% and unburned hydrocarbons by 45.2% CH2–OOC–R3 R3–COO–R’ CH2–OH
(Schumacher et al., 2001). Additionally, biodiesel is biode- ð1Þ
gradable and non-toxic, making it useful for transportation
Transesterification can be alkali-, acid- or enzyme-cata-
lyzed; however, enzyme-catalysts are rarely used, as they
*
Corresponding author. Tel.: +1 604 822 1243; fax: +1 604 822 6003. are less effective (Ma and Hanna, 1999). The reaction
E-mail address: nellis@chml.ubc.ca (N. Ellis). can also take place without the use of a catalyst under

0960-8524/$ - see front matter Ó 2007 Elsevier Ltd. All rights reserved.
doi:10.1016/j.biortech.2007.11.046
6588 A.H. West et al. / Bioresource Technology 99 (2008) 6587–6601

Nomenclature

ATROR after tax rate-of-return NRTL non-random two liquid


CAC auxiliary facility cost FP pressure factor
CBM bare module capital costs Cp purchase cost
FBM bare module factor K1, K2 purchase cost parameters
B1, B2 bare module factor parameters TCI, CTCI total capital investments
A capacity parameter TMC total manufacturing cost
CCF contingency fee CTM total module cost
CFC fixed capital cost wt.% weight per cent
FFA free fatty acid CWC working capital cost
FM materials factor

conditions in which the alcohol is in a supercritical state one step acid-catalyzed process was the most economically
(Demirbas, 2002; Saka and Kusdiana, 2001). attractive (Zhang et al., 2003b). Haas et al. (2006) devel-
Currently, the high cost of biodiesel production is the oped a versatile process simulation model to estimate bio-
major impediment to its large-scale commercialization diesel production costs; however, the model was limited to
(Canakci and Van Gerpen, 2001). The high cost is largely a traditional alkali-catalyzed production method. In order
attributed to the cost of virgin vegetable oil as feedstock. to determine whether the supercritical methanol or the het-
Exploring methods to reduce the production cost of biodie- erogeneous acid-catalyst process is a promising alternative
sel has been the focus of much recent research. One method to the standard homogeneous catalytic routes, our aim was
involves replacing a virgin oil feedstock with a waste cook- to develop a process flowsheet and simulation, conduct an
ing oil feedstock. The costs of waste cooking oil are esti- economic analysis of each process based on the material
mated to be less than half of the cost of virgin vegetable and energy balance results reported by HYSYS, and carry
oils (Canakci and Van Gerpen, 2001). Furthermore, utiliz- out sensitivity analyses to optimize each process. Addition-
ing waste cooking oil has the advantage of removing a sig- ally, it was desired to automate the sizing and economic
nificant amount of material from the waste stream – as of calculations, whence they were incorporated into each sim-
1990, it was estimated that at least 2 billion pounds of ulation by way of the spreadsheet tool available in HYSYS.
waste grease was produced annually in the United States The material and energy flows, as well as some unit param-
(Canakci and Van Gerpen, 2001). eters were imported directly into the spreadsheet, thereby
In the last few years, a number of new production meth- allowing the sizing and economic results to be updated
ods have emerged from laboratory/bench-scale research automatically when any changes were made to the process
aimed at reducing the cost of biodiesel (Canakci and Van flowsheet. Additional comparison is made to the simula-
Gerpen, 2003; Delfort et al., 2003; Demirbas, 2002). One tion work by Zhang et al. (2003a) in order to ensure that
such method uses alcohol in its supercritical state, and the present simulations provide comparable results.
eliminates the need for a catalyst. Additionally, the super- The homogeneous alkali-catalyzed system has been well
critical process requires only a short residence time to reach studied, and optimum conditions at 1 atm pressure (60 °C,
high conversion (Kusdiana and Saka, 2004). Another 1 wt.% catalyst, six-to-one alcohol-to-oil molar ratio), are
option is to use a solid catalyst (Abreu et al., 2005; Furuta known (Freedman et al., 1984). In order to prevent sapon-
et al., 2004; Suppes et al., 2004) which simplifies down- ification during the reaction, the free fatty acid (FFA) and
stream purification of the biodiesel. The catalyst can be water content of the feed must be below 0.5 wt.% and
separated by physical methods such as a hydrocyclone in 0.05 wt.%, respectively (Freedman et al., 1984). Because
the case where a multiphase reactor is used. Alternatively, of these limitations, only pure vegetable oil feeds are
a fixed bed reactor would eliminate the catalyst removal appropriate for alkali-catalyzed transesterification without
step entirely. extensive pre-treatment.
Zhang et al. (2003a) developed a HYSYS-based process A homogeneous acid-catalyzed process can be employed
simulation model to assess the technological feasibility of to take advantage of cheaper feedstocks, such as waste
four biodiesel plant configurations – a homogeneous cooking oil and animal-based tallow. The acid-catalyzed
alkali-catalyzed pure vegetable oil process; a two step pro- process can tolerate up to 5 wt.% FFA, but is sensitive to
cess to treat waste vegetable oil; a single step homogeneous water content greater than 0.5 wt.%. The disadvantage of
acid-catalyzed process to treat waste vegetable oil; and a this method is that it is extremely slow at mild conditions:
homogeneous acid-catalyzed process using hexane extrac- to achieve a 98% conversion, reaction for 48 h at 60 °C at
tion to help purify the biodiesel. All four configurations an alcohol-to-oil ratio of 30:1 was required (Canakci and
were deemed technologically feasible, but a subsequent Van Gerpen, 1999). At higher temperatures and pressures
economic analysis of the four designs revealed that the (e.g., 100 °C and 3.5 bar) reaction times can be substan-
A.H. West et al. / Bioresource Technology 99 (2008) 6587–6601 6589

Table 1 sis, complete process simulations were performed. Despite


Catalysts and reaction parameters for heterogeneously catalyzed reactions some expected differences between a process simulation
of soybean oil at 1 atm
and real-life operation, process simulators are commonly
Reaction parameters used to provide reliable information on process operation,
Catalyst type Molar Temperature Conversion Time owing to their vast component libraries, comprehensive
ratio (°C) (h) thermodynamic packages and advanced computational
WO3/ZrO2 (Furuta et al., 40:1 >250 >90% 4 methods. HYSYS (HYprotech SYStem) Plant NetVer 3.2
2004) (ASPEN Tech, Cambridge MA) was used to conduct the
SO4/SnO2 (Furuta et al., 40:1 300 65% 4
2004)
simulation. HYSYS was selected as a process simulator
SO4/ZrO2 (Furuta et al., 40:1 300 80% 4 for both its simulation capabilities and its ability to incor-
2004) porate calculations using the spreadsheet tool. The first
SnO (Abreu et al., 2005) 4.15:1 60 94.7 3 steps in developing the process simulation were selecting
the chemical components for the process, as well as a ther-
modynamic model. Additionally, unit operations and oper-
tially reduced (down to 8 h) to achieve similar conversion ating conditions, plant capacity and input conditions must
(99%) (Goff et al., 2004). all be selected and specified. The unit operations, plant
To utilize the advantages of the alkali- and acid-cata- capacity and input conditions for the base cases, i.e.,
lyzed processes, a process that combines both schemes homogeneous acid- and alkali-catalyzed processes, as well
can be created. FFA waste vegetable oil can be pre-treated as distillation column operating conditions, were selected
via esterification with methanol in an acid-catalyzed envi- based on the research done by Zhang et al. (2003a) to
ronment to convert the FFA to methyl-ester. The acid-free ensure that each of the four processes simulated in this
oil feedstock can then be treated by a base catalyst in the work could be compared in a consistent manner.
manner described above for the homogeneous alkali-cata- HYSYS library contained information for the following
lyzed system. This process has the advantages that it can components used in the simulation: methanol, glycerol, sul-
use a cheaper feedstock such as waste vegetable oil, and furic acid, sodium hydroxide, and water. Waste Canola oil
employ alkali-catalyzed transesterification which is fast was represented by triolein and oleic acid. Accordingly,
and efficient (Canakci and Van Gerpen, 2001). methyl-oleate, available in the HYSYS component library,
A process using a heterogeneous acid-catalyst is appeal- was taken as the product of the transesterification reaction.
ing because of the ease of separation of a solid catalyst. Lote- Where a simulation required a feedstock with some amount
ro et al. (2005) reports this advantage, coupled with the of free fatty acids, oleic acid, also available in the HYSYS
ability of the acid functionality to process low cost, high free library, was specified as the free fatty acid present.
fatty acid feedstocks, will yield the most economical biodie- Components not available in the HYSYS library were
sel production method. As outlined in Table 1, a number of specified using the ‘‘Hypo Manager” tool. Calcium oxide,
solid phase catalysts have been identified that hold potential calcium sulfate, phosphoric acid, sodium phosphate and
for use. Research concerning heterogeneous catalysts is still triolein were all specified in this manner. Specification of
in the catalyst screening stage. Studies regarding reaction a component requires input of a number of properties, such
kinetics, as well as improving reaction parameters have yet as normal boiling point, density, molecular weight, as well
to be conducted. In addition, studies to determine the effects as the critical properties of the substance. Since triolein is a
of free fatty acid concentration and water on the perfor- crucial component and is involved in operations requiring
mance of the catalyst have been scarce. data for liquid and vapour equilibria, great care was taken
Supercritical transesterification is also a potential alter- in specifying the values as accurately as possible. Values for
native to the standard homogenous catalytic routes. Super- density, boiling point, critical temperature, pressure and
critical transesterification using methanol has been shown volume were obtained from the ASPEN Plus component
to give nearly complete conversion in small amount of time library and were input as 915 kg/m3, 846 °C, 1366 °C,
(15 min) (Warabi et al., 2004). High temperatures (up to 470 kPa, 3.09 m3/kmol, respectively.
350 °C) and large alcohol-to-oil ratios (42:1) are required Owing to the presence of polar compounds such as
to achieve the high levels of conversion that have been methanol and glycerol in the process, the non-random
reported (Kusdiana and Saka, 2001). In addition to the two liquid (NRTL) thermodynamic/activity model was
high conversion and reaction rates, supercritical transeste- selected for use as the property package for the simulation.
rification is appealing as it can tolerate feedstocks with very Since some binary interaction parameters were not avail-
high contents of FFAs and water, up to 36 wt.% and able in the simulation databanks, they were estimated using
30 wt.%, respectively (Kusdiana and Saka, 2004). the UNIFAC vapour–liquid equilibrium and UNIFAC
liquid–liquid equilibrium models where appropriate.
2. Process simulation Plant capacity was specified at 8000 metric tonnes/year
biodiesel (the same as in Zhang et al., 2003a). This trans-
To assess the technological feasibility and obtain mate- lated to vegetable oil feeds of roughly 1000 kg/h for each
rial and energy balances for a preliminary economic analy- process configuration.
6590 A.H. West et al. / Bioresource Technology 99 (2008) 6587–6601

The process units common to all configurations included alkali-catalyzed case), while vacuum operation in the
reactors, distillation columns, pumps and heat exchangers. methyl-ester purification columns was necessary to keep
The homogeneous acid- and alkali-catalyzed processes the temperatures of the distillate and bottoms streams at
included liquid–liquid extraction columns to separate the suitably low levels, as biodiesel and glycerol are subject
catalyst and glycerol from the biodiesel. In contrast to to degradation at temperatures greater than 250 °C and
the base case scenarios, a gravity separation unit was 150 °C, respectively (Goodrum, 2002; Newman, 1968).
included in the supercritical methanol and heterogeneous
acid-catalyst processes. In spite of the availability of kinetic 3. Process design
data for the alkali-catalyzed, homogeneous acid-catalyzed,
and supercritical processes (Freedman et al., 1986; Kusdi- Four continuous processes were simulated. The first was
ana and Saka, 2001), the reactors were modeled as conver- based on an alkali-catalyzed reaction to convert a waste
sion reactors since kinetic information for the vegetable oil feedstock containing 5 wt.% FFA that was
heterogeneous acid-catalyzed process was unavailable. pre-treated with an acid-catalyzed reaction (Fig. 1a). The
The reactors were assumed to operate continuously for second was based on a homogeneous acid-catalyzed pro-
all cases. Lab-scale reaction conditions and conversion cess using a waste cooking oil feedstock, containing
data available for all processes (Abreu et al., 2005; Warabi 5 wt.% free fatty acids. The third configuration employed
et al., 2004; Zhang et al., 2003a), were assumed to be a heterogeneous acid-catalyst, (tin(II) oxide), in a multi-
appropriate for large-scale production, and set as the oper- phase reactor fed with waste vegetable oil, while the final
ating conditions for each reactor. The following conver- process used a supercritical methanol treatment of waste
sions were assumed for each process: 97%, 97%, 94% and vegetable oil to produce biodiesel. Process flowsheets are
98% for the alkali, acid, heterogeneous and supercritical presented in Figs. 1–4.
cases, respectively. The mono- and di-glyceride intermedi- The processes all followed the same general scheme. The
ates were neglected during the reaction (Zhang et al., vegetable oil was transesterified in the first step, and then
2003a). sent for downstream purification. Downstream purification
Multi-stage distillation was used to recover the excess consisted of the following steps: methanol recovery by dis-
methanol, as well as in the final purification of biodiesel. tillation; glycerol separation; catalyst neutralization and
Distillation columns were specified to meet or exceed the removal (where appropriate) and methyl-ester purification
ASTM standard for biodiesel purity, i.e., 99.65 wt.%. by distillation. Table 2 gives details for the unit operations
Reflux ratios for the heterogeneous acid-catalyzed and in each process. Tables 3–6 present the feed and product
supercritical cases were calculated by determining the min- material flow details for each process.
imum reflux ratio using a shortcut distillation column, and As illustrated in Table 2, there are also a number of key
then multiplying by 1.5 to obtain the optimum reflux ratio differences between the processes. The first difference is
(McCabe et al., 2001). The methanol recovery columns with regards to the catalyst removal method. The solid cat-
were able to operate at ambient pressures (except in the alyst in Process III is removed by a hydrocyclone before

Fig. 1. (a) Pre-treated alkali-catalyzed process flowsheet. Pre-treatment component. (b) Pre-treated alkali-catalyzed process flowsheet. Transesterification
component.
A.H. West et al. / Bioresource Technology 99 (2008) 6587–6601 6591

Fig. 1 (continued)

Fig. 2. Homogeneous acid-catalyzed process flowsheet.

methanol recovery; whereas the liquid phase catalyst in The acid-catalyst in Process II was removed as a solid pre-
Processes I and II is removed by washing the product cipitate in separator X-100 after neutralization in reactor
stream with water in a liquid–liquid extraction column. CRV-101. As in the homogeneous acid-catalyst process,
6592 A.H. West et al. / Bioresource Technology 99 (2008) 6587–6601

Fig. 3. Heterogeneous acid-catalyzed process flowsheet.

Fig. 4. Supercritical alcohol process flowsheet.

the alkali-catalyst had to be neutralized before it could be separator by gravity settling. Krawczyk (1996) initially pro-
disposed of. The heterogeneous catalyst in Process III posed gravity separation to remove glycerol; however,
required no neutralization step and it was discarded as a Zhang et al. (2003a) indicated from their simulation that
waste product. However, it has the potential advantage satisfactory separation could not be achieved by gravity
of being recycled. alone. In the present work, gravity separation was used
The second major difference is in the separation of glyc- to separate the biodiesel from the glycerol, and a satisfac-
erol from the biodiesel. In Processes I and II, glycerol is tory separation was achieved. Note, however, that the cal-
removed by washing the product stream with water, and culations for this unit operation are based on parameters
collected in the bottoms product. In Processes III and IV, that have been estimated by HYSYS and, therefore, may
glycerol is separated from the biodiesel in a three-phase not truly represent a real system. Additional experimental
A.H. West et al. / Bioresource Technology 99 (2008) 6587–6601 6593

Table 2
Summary of unit operating conditions for each process
Pre-treated alkali-catalyzed Acid-catalyzed Heterogeneous acid-catalyzed Supercritical process
Transesterification
Catalyst H2SO4a/NaOH H2SO4 SnO N/A
Reactor type CSTRa/CSTR CSTR Multiphase CSTR
Temperature (°C) 70a/60 80 60 350
Pressure (kPa) 400a/400 400 101.3 20  103
Alcohol-to-oil ratio 6:1a/6:1 50:1 4.5:1 42:1
Residence time (h) 1a/4 4 3 0.333
Conversion (%) 100a/95 97 94 98
Methanol recovery
Reflux ratio 5a/2 2 3.99 3.42
Number of stages 5a/6 6 14 12
Condenser/reboiler pressure (kPa) 20a/30 101.3/111 40/50 101.3/105.3
%Recovery 94a/94 99.2 99.9 99.3
Distillate flowrate (kg/h) 201a/119.3 1687 66.33 1239.7
Distillate purity (%) 99.5a/100 100 99.9 99.99
Catalyst removal
Glycerol washa/N/A N/A hydrocyclone N/A
Glycerol separation
Water washing Water washing Gravity Gravity
Water flowrate 11 kg/h 46 kg/h – –
Catalyst neutralization
Neutralizing agent H3PO4 CaSO4 N/A N/A
Biodiesel recovery
Reflux ratio 2 2 2 2
Number of stages 5 10 8 8
Condenser/reboiler pressure (kPa) 10/15 10/15 101.3/111.3 101.3/111.3
%Recovery 99.9 98.65 99.9 99.9
Final purity 99.97 99.65 99.9 99.65
a
Indicates operating conditions for the pre-treatment unit in the pre-treatment case.

Table 3
Feed and product stream information for the pre-treated alkali-catalyzed process
Feed streams Product streams
101 105 103 106 111 112 113
Temperature (°C) 25.0 25.0 25.0 Temperature (°C) 70 35.1 71.8 60.0
Pressure (kPa) 101.3 101.3 101.3 Pressure (kPa) 400 20 30 400
Molar flow (kgmol/h) 0.641 1.34 0.10 Molar flow (kgmol/h) 9.28 7.17 1.97 1.34
Mass flow (kg/h) 20.6 1050.00 10.00 Mass flow (kg/h) 1310 229.1 138.3 1053
Component mass fraction Component mass fraction
Methanol 1.000 0.000 0.000 Methanol 0.1853 0.9963 0.1052 0.0000
Triolein 0.000 0.950 0.000 Glycerol 0.0000 0.0000 0.7956 0.0000
H2SO4 0.000 0.000 1.000 Triolein 0.7532 0.0000 0.0000 0.9372
Oleic acid 0.000 0.0500 0.0000 Methyl-oleate 0.0505 0.0000 0.0000 0.0628
H2SO4 0.0074 0.0000 0.0709 0.0000
Water 0.0036 0.0037 0.0285 0.0000
Streams represent only the pre-treatment portion.

data is needed to verify the applicability and results of the 4. Equipment sizing
gravity separator, in order to use the unit block with con-
fidence. In practice, a gravity separation unit has been used Process equipment was sized according to principles
on a pilot plant-scale to separate glycerol and biodiesel outlined in the literature (Seider et al., 2003; Turton
(Canakci and Van Gerpen, 2003). All processes produced et al., 2003). The principal dimensions of each unit are pre-
biodiesel at a higher purity than required by the ASTM sented in Table 7. The equipment sizing calculations were
standard of 99.65 wt.%. conducted using the Spreadsheet tool available within
6594 A.H. West et al. / Bioresource Technology 99 (2008) 6587–6601

Table 4
Feed and product stream information for the alkali-catalyzed transesterification step of the pre-treated process
Feed streams Product streams
101 105-PVO 103 401A 401 402 501 502
Temperature (°C) 25.0 25.0 25.0 Temperature (°C) 167.8 167.5 463.9 42.8 148.6
Pressure (kPa) 101.3 101.3 101.3 Pressure (kPa) 10 10 15 20 30
Molar flow (kgmol/h) 3.61 1.19 0.25 Molar flow (kgmol/h) 0.12 3.38 0.06 0.65 1.20
Mass flow (kg/h) 115.71 1050.00 10.00 Mass flow (kg/h) 4.57 1001.8 52.77 13.79 105.12
Component mass fraction Component mass fraction
Methanol 1.0000 0.0000 0.0000 Methanol 0.6114 0.0001 0.0000 0.3432 0.0001
Triolein 0.0000 0.9372 0.0000 Glycerol 0.0005 0.0000 0.0000 0.0002 0.9865
NaOH 0.0000 0.0000 1.0000 Triolein 0.0000 0.0001 0.9967 0.0000 0.0014
Glycerol 0.0000 0.0000 0.0000 Methyl-oleate 0.2125 0.9997 0.0033 0.0000 0.0002
Methyl-oleate 0.0000 0.0628 0.0000 NaOH 0.0000 0.0000 0.0000 0.0000 0.0000
H2SO4 0.0000 0.0000 0.0000 H3PO4 0.0000 0.0000 0.0000 0.0000 0.0000
Water 0.0000 0.0000 0.0000 Na3PO4 0.0000 0.0000 0.0000 0.0000 0.0000
Water 0.1755 0.0000 0.0000 0.6565 0.0119

Table 5
Feed and product stream information for the homogeneous acid-catalyzed process
Feed streams Product streams
101 103 105 401A 401 402 501 502
Temperature (°C) 25 25 25 Temperature (°C) 130.7 234.3 702.2 23.4 226.6
Pressure (kPa) 101.3 101.3 101.3 Pressure (kPa) 35 45 55 10 15
Molar flow (kgmol/h) 3.78 1.53 1.17 Molar flow (kgmol/h) 0.65 3.42 0.05 6.59 1.10
Mass flow (kg/h) 121.2 150.06 1030.00 Mass flow (kg/h) 20.42 1002.98 33.22 155.64 101.69
Component mass fraction Component mass fraction
Methanol 1.000 0.000 0.000 Methanol 0.957 0.001 0.000 0.520 0.000
Triolein 0.000 0.000 0.950 Glycerol 0.001 0.000 0.000 0.009 0.993
H2SO4 0.000 1.000 0.000 Triolein 0.000 0.001 0.889 0.000 0.007
Oleic acid 0.000 0.000 0.050 H2SO4 0.000 0.000 0.000 0.000 0.000
Methyl-oleate 0.007 0.998 0.111 0.003 0.000
CaO 0.000 0.000 0.000 0.000 0.000
Water 0.035 0.000 0.000 0.468 0.000
CaSO4 0.000 0.000 0.000 0.000 0.000
Oleic acid 0.000 0.000 0.000 0.000 0.000

Table 6
Feed and product stream information for the heterogeneous acid-catalyzed process
Feed streams Product streams
Methanol 101 SnO 103 Triolein 105 302 Glycerol out 401 402
Temperature (°C) 25.0 25.0 25.0 Temperature (°C) 25.0 203.2 535.5
Pressure (kPa) 101.3 101.3 101.3 Pressure (kPa) 50 101.3 111.3
Molar flow (kgmol/h) 3.38 0.04 1.31 Molar flow (kgmol/h) 1.22 3.38 0.07
Mass flow (kg/h) 108.3 10.54 1050.00 Mass flow (kg/h) 100.4 989.6 59.80
Component mass fraction Component mass fraction
Methanol 1.0000 0.0000 0.0000 Methanol 0.0004 0.0000 0.0000
Triolein 0.0000 0.0000 0.9500 Glycerol 0.9625 0.0001 0.0001
Tin(II) oxide 0.0000 1.0000 0.0000 Triolein 0.0064 0.0000 0.9835
Oleic acid 0.0000 0.0000 0.0500 Methyl-oleate 0.0002 0.9995 0.0165
Tin(II) oxide 0.0000 0.0000 0.0000
Oleic acid 0.0000 0.0000 0.0000
Water 0.0304 0.0002 0.0000

HYSYS. Key variables for unit sizing were imported from were encoded within the spreadsheet. Therefore, any alter-
the flowsheet directly to the spreadsheet. Sizing equations ations to the flowsheet, such as component fractions, com-
A.H. West et al. / Bioresource Technology 99 (2008) 6587–6601 6595

Table 7
Feed and product stream information for the supercritical methanol process
Feed streams Product streams
101 Methanol 103 Triolein 302 Glycerol out 401 402
Temperature (°C) 25 25 Temperature (°C) 25 134.5 463.7
Pressure (kPa) 100 100 Pressure (kPa) 105.3 101.3 111.3
Molar flow (kgmol/h) 3.67 1.31 Molar flow (kgmol/h) 1.44 3.62 0.03
Mass flow (kg/h) 117.8 1050.00 Mass flow (kg/h) 110.1 1039.4 20.83
Component mass fraction Component mass fraction
Methanol 1.0000 0.0000 Methanol 0.0501 0.0030 0.0000
Triolein 0.0000 0.9500 Glycerol 0.9180 0.0006 0.0000
Oleic acid 0.0000 0.0500 Triolein 0.0012 0.0000 0.9947
Methyl-oleate 0.0033 0.9960 0.0052
Oleic acid 0.0000 0.0000 0.0000
Water 0.0272 0.0003 0.0000

ponent flowrates, changes to the desired recovery in the dis- process vessels with an aspect ratio of 2. They were sized
tillations columns are automatically calculated and imple- to allow for continuous operation, with a residence time
mented, thus eliminating tedious recalculations steps. of 1 h.

4.1. Reactor vessels 4.4. Hydrocyclone

Reactors were sized for continuous operation by divid- The initial dimensions of the hydrocyclone (used to sep-
ing the residence time requirement by the feed flowrate arate the solid catalyst from the product stream in Process
for each process. Residence times were: 4 h, 4 h, 3 h and III) were calculated by the unit block in HYSYS. Those
20 min for the pre-treated alkali-catalyzed, acid-catalyzed, dimensions were then manipulated slightly to obtain com-
heterogeneous acid-catalyzed and supercritical processes, plete removal of the catalyst in the hydrocyclone
respectively. The vessels were specified to have an aspect underflow.
ratio of three-to-one.
5. Economic assessment
4.2. Columns
Since each process was capable of producing biodiesel at
Distillation column diameters were sized by two meth- the required purity, it was of interest to conduct an eco-
ods. An initial diameter was estimated from the F-Factor nomic assessment to determine process viability, and deter-
Method (Luyben, 2002). If the column diameter was calcu- mine if any one process was advantageous over the others.
lated to be greater than 0.90 m (2.95 ft) it was specified as a As with the sizing calculations, all the economic calculations
tray tower, and thus calculated from the flooding velocity were performed within the HYSYS spreadsheet. Addition-
using the Fair correlation (Seider et al., 2003). Columns ally, the values presented for the economic analysis are the
with diameters calculated at less than 0.9 m were specified values obtained after performing sensitivity analyses and
as a packed tower. The diameter of each packed column optimization of each process. The details for the sensitivity
was calculated from the flooding velocity obtained from analyses and optimization studies are presented in Section 6
the Leva correlation (Seider et al., 2003). of this paper. All parameters necessary to determine mate-
Tray tower height was calculated by multiplying the rial and energy costs were imported to the spreadsheet from
number of actual stages by the tray spacing, and then the flowsheet. Costing equations were incorporated directly
increasing the result by 20% to provide height for the con- into the spreadsheet as well. Individual unit costs were cal-
denser and reboiler. culated, as well as figures for each process in its entirety.
Packed tower height was calculated by multiplying the Incorporating the economic calculations into the simulation
height equivalent of a theoretical plate (HETP) by the num- allowed for automatic recalculation of process economics
ber of stages calculated for the tower. HETP was assumed should any process parameters, such as component flow-
to equal the column diameter (Seider et al., 2003). As for rates or unit operating conditions be changed. By integrat-
the height of a tray tower, the packed height was increased ing sizing and economic calculations into each simulation,
by 20%. The liquid–liquid extraction columns were sized the potential to perform economic sensitivity analyses is
according to the work of Zhang et al. (2003a). automatically built-in to each simulation.

4.3. Gravity separators 5.1. Basis of calculations

The gravity separators in the heterogeneous acid-cata- Each process was based on a plant capacity of 8000 ton-
lyzed and supercritical processes were designed as vertical nes/year biodiesel production. Operating hours were set at
6596 A.H. West et al. / Bioresource Technology 99 (2008) 6587–6601

7920 h/year (assuming 330 operating days/year). The feed- 5.3. Total manufacturing cost
stock was assumed free of water and solid impurities. Low
and high pressure steam were used as heating media, while Direct manufacturing expenses were calculated based on
water was used for cooling. the price and consumption of each chemical and utility.
Each process was evaluated based on total capital Chemical and utility prices are presented in Table 9 and
investment (TCI), total manufacturing cost (TMC), and material flow information was obtained from HYSYS.
after tax rate-of-return (ATROR). The assessment per- Operator salary was estimated at $47,850/year, and it
formed in this work is classified as a ‘‘study estimate” (Tur- was assumed that an operator worked 49 weeks/year, and
ton et al., 2003) with a range of expected accuracy from there were three 8-h shifts per day for the continuous plant
+30% to 20%. While the results of such a study will likely (Zhang et al., 2003b). Table 10 presents a breakdown of the
not reflect the final cost of constructing a chemical plant, components of the total manufacturing cost as well as the
the technique is useful for providing a relative means to results for each process. After tax rate-of-return is a general
compare competing processes. criterion for economic performance of a chemical plant. It
is defined as the percentage of net annual profit after taxes,
5.2. Total capital investment relative to the total capital investment. Net annual profit
after taxes (ANNP) is half the net annual profit (ANP)
Table 8 gives a breakdown of the total capital invest- assuming a corporate tax rate of 50%. The results for after
ment. It also presents the costs for the individual unit oper- tax rate-of-return for each process are shown in Table 10.
ations in each process. As shown in Table 8, the transesterification reactor
Bare module capital costs (CBM) consist of the purchase forms a large part of the capital cost, especially for Pro-
cost of a piece of equipment, multiplied by the bare module cesses II and IV. The reactor in Process II was required
factor. Purchase costs were estimated for each piece of to contain a large material flow at a long residence time.
equipment based on a capacity equation (Turton et al., The presence of sulfuric acid as the catalyst required a
2003) stainless steel reactor, resulting in a substantially higher
2 reactor cost. Consequently the reactor in Process II was
log10 C p ¼ K 1 þ K 2 log10 ðAÞ þ K 3 ½log10 ðAÞ ð2Þ
much more expensive than in all other processes. The
where Ki is a constant specific to the unit type and A is supercritical reactor was required to withstand a high pres-
the capacity of the unit. Bare module cost was calculated sure, and was constructed from stainless steel to prevent
from oxidation and corrosion, hence its high cost. Distillation
columns also contributed a significant part to the capital
C BM ¼ C p F BM ð3Þ
cost of each process. Tower costs for the methyl-ester puri-
where FBM is given by fication tower were roughly equal between the processes, as
each tower was handling approximately the same material
F BM ¼ B1 þ B2 F M F P ð4Þ
flows and producing biodiesel at equal purities. The meth-
where B1 and B2 are constants specific to the unit type, and anol recovery columns in Processes I and III were the least
FM and FP are the material and pressure factors, respec- expensive, as they had the smallest material flow require-
tively. The constants, Ki and Bi, as well as the pressure ments. In spite of Process IV having the smallest number
and material factors were obtained from the literature of unit operations, Process III had the smallest total capital
(Turton et al., 2003). Eqs. (2)–(4) were encoded within investment. This is due to the fact that Process IV required
the costing spreadsheet to allow for automatic cost updates a more expensive reactor in order to withstand the high
when process parameters are changed. pressures and corrosive conditions associated with the

Table 8
Equipment sizes for various process units in all processes
Type Description Pre-treated alkali-catalyzed Acid-catalyzed Heterogeneous acid-catalyzed Supercritical process
Reactor Pre-treatment 0.8  2.4
Transesterification 1.8  5.4 2.1  6.3 1.2  3.64 0.96  2.9
Neutralization 0.36  1.1 0.5  1.5 N/A N/A
Columns Methanol recovery 0.6  10a/0.46  3 0.9  8.6 0.31  7.4 1  8.8
Fame purification 0.9  9.5 1  8.5 0.9  6.6 1  6.6
Water washing 1  10a/0.8  10 1  10 N/A N/A
Glycerol purification N/A 0.5  3.7 N/A N/A
Separators Gravity separators N/A N/A 1.2  2.4 1.1  2.4
Hydrocyclone N/A N/A 0.112  1.35 N/A
Dimensions are diameter x height (m).
a
Indicates units that are pre-treatment process components.
A.H. West et al. / Bioresource Technology 99 (2008) 6587–6601 6597

Table 9
Equipment costs, fixed capital costs and total capital investments for each process
Type Description Pre-treated alkali-catalyzed Acid-catalyzed Heterogeneous acid-catalyzed Supercritical process
Reactor Esterification 0.087 – – –
Transesterification 0.292 0.680 0.075 0.639
Neutralization 0.027 0.036 0 0
Columns Pre-treat methanol recovery 0.143 – – –
Methanol recovery 0.038 0.152 0.028 0.167
Fame purification 0.102 0.076 0.095 0.146
Glycerol washing .231 – – –
Water washing 0.084 0.113 0 0
Glycerol purification 0 0.028 0 0
Other Gravity separators 0 0 0.057 0.058
Heat exchangers 0 0.079 0.079 0.109
Pumps 0.014 0.010 0.014 0.141
Others (hydrocyclone, etc.) 0 0 0.015 0
Total bare module cost, CBM 0.93 1.17 0.37 1.26
Contingency fee, CCF = 0.18CBM 0.17 0.22 0.07 0.23
Total module cost, CTM = CBM + CCF 1.10 1.38 0.43 1.49
Auxiliary facility cost, CAC = 0.3CBM 0.279 0.35 0.11 0.38
Fixed capital cost, CFC = CTM + CAC 1.38 1.73 0.54 1.87
Working capital CWC = 0.15CFC 0.21 0.26 0.08 0.28
Total capital investment CTCI = CFC + CWC 1.59 1.99 0.63 2.15
Costs are reported as $ millions.

Table 10 mostly in the lower costs calculated for the methanol recov-
Conditions for the economic assessment of each process (Zhang et al., ery column and methyl-ester purification column, due to
2003b)
the differences in sizing.
Item Specification Price ($/tonne) Results for the total manufacturing cost of each process
Chemicals are shown in Table 11. The direct manufacturing cost rep-
Biodiesel 600 resents between 71% and 77% of the total manufacturing
Calcium oxide 40
Glycerolb 92 wt.% 1200
cost in each process. The largest proportion of the direct
85 wt.% 750 manufacturing cost is due to the oil feedstock, namely
39% for Process I, and around 43% for the other processes.
Methanol 99.85% 180
Phosporic acid 340 Process III has the lowest total manufacturing cost. This is
Sodium hydroxide 200 due to both the ability of the process to use low cost waste
Sulfuric acid 60 vegetable oil, as well as the lower utility costs of the process
Tin (II) oxide 600 resulting from the smaller material streams handled in the
Pure canola oil 500
process. The total manufacturing cost of Process IV is
Waste cooking oil 200
slightly more than that of Process III, owing to the large
Utilities energy requirements necessary to separate the methanol
Cooling water 400 kPa 6 °C $0.007/m3
Electricity $0.062/kWh
from the product stream. In spite of the use of an efficient
Low pressure steama 601.3 kPa 160 °C $7.78/GJ reaction scheme in Process I, the total manufacturing costs
High pressure steama 4201.3 kPa 254 °C $19.66/GJ are higher than in Process II because of the pre-treatment
a
Value from Turton et al., 2003. step: the additional catalyst and methanol consumed to
b
Glycerol is used as a solvent in Process I, and its cost is assumed to be convert the FFA, and the large costs associated with using
750 $/tonne. glycerol as a washing agent in the first liquid washing col-
umn add considerable expense to the process. This use of
glycerol for washing essentially negates the benefit of the
supercritical state of the alcohol, as well as the larger meth- glycerol production credits that the other processes are able
anol recovery tower. The total capital investment for Pro- to realize.
cess I was less expensive than Process II even though it Except for Process III, all processes had a negative after
had a greater number of required unit operations due to tax rate-of-return. Process I had the lowest ATROR, at
the waste oil pre-treatment step. Again this is due to 22%, while Processes II and IV had ATRORs at 9%
decreased material flows and reactor costs. The total capi- and 1%, respectively. The ATROR for Process III was
tal investment for Process I in the present work was calcu- 58%, indicating that the process could earn a profit without
lated to be $1.59 million, less than the value reported by any government subsidies. The value for ATROR reported
Zhang et al. (2003b) of $2.68 million. The difference lies by Zhang et al. (2003b) for the pre-treated Process I was
6598 A.H. West et al. / Bioresource Technology 99 (2008) 6587–6601

Table 11
Total manufacturing cost and after tax rate-of-return for each process
Process I Process II Process III Process IV
Direct manufacturing cost
Oil feedstock 1.66 1.63 1.66 1.66
Methanol 0.18 0.30 0.16 0.17
Catalyst 0.70 0.10 0.05 0.00
Operating labour 0.58 0.58 0.58 0.58
Supervisory labour 0.09 0.09 0.09 0.09
LP steam 0.13 0.36 0.05 0.39
HP steam 0.26 0.25 0.28 0.33
Electricity 0.00 0.00 0.00 0.00
Cooling water 0.01 0.02 0.01 0.02
Liquid waste disposal 0.22 0.09 0.05 0.02
Solid waste disposal 0.01 0.06 0.02 0.00
Maintenance & Repairs (M&R), 6% of CFC 0.08 0.11 0.03 0.11
Operating supplies, 15% of M&R 0.01 0.02 0.00 0.02
Lab charges, 15% of operating labour 0.09 0.09 0.09 0.09
Patents and royalties, 3% TMC 0.16 0.15 0.12 0.14
Subtotal ADMC 4.19 3.84 3.19 3.61
Indirect manufacturing cost
Overhead, packaging and storage, 60% of sum of operating labour, supervision and maintenance 0.45 0.46 0.42 0.47
Local taxes, 1.5% of CFC 0.02 0.03 0.01 0.03
Insurance, 0.5% of CFC 0.01 0.01 0.00 0.01
Subtotal, AIMC 0.46 0.47 0.42 0.47
Depreciation 10% of CFC 0.14 0.18 0.05 0.19
General expenses
Administrative costs, 25% of overhead 0.11 0.12 0.10 0.12
Distribution and selling, 10% of TMC 0.52 0.48 0.39 0.46
Research & Development, 5% of TMC 0.26 0.24 0.19 0.23
Subtotal 0.89 0.84 0.69 0.81
Total production cost 5.78 5.37 4.45 5.19
Glycerine credit 0.58 0.61 0.57 0.60
Total manufacturing cost, ATE 5.20 4.76 3.88 4.59
Revenue from biodiesel 4.77 4.76 4.70 4.92
Net annual profit 0.43 0.005 0.82 0.33
Income taxes, AIT 50% of ANP 0.21 0.003 0.41 0.17
Net annual after tax profit, ANNP 0.21 0.003 0.41 0.17
After tax rate-of-return, I = [ANNP  ABD]/CTC (%) 22.2 8.71 58.76 0.90
Costs are reported as $millions.

51% was reasonably close to the value reported in this Since the conversion data for the heterogeneous acid-
work. Additionally, our rate-of-return for Process II catalyzed and supercritical processes were taken from
(9%) was in close agreement with the value reported for bench-scale research, the economics of scale may not be
the acid-catalyzed case by Zhang et al. (2003b) of 15%. accurately reflected. Thus the effect of reduced conversion
The relative economic order of Processes I and II (i.e., that on the overall process economics was examined for each
Process II has an ATROR greater than that of Process I) as process. As shown in Fig. 5, conversion in the heteroge-
presented in this work is also in agreement with that of neous acid-catalyzed process must drop to approximately
Zhang et al. (2003b). As predicted by Lotero et al. 85%, while conversion in the supercritical and homoge-
(2005), the heterogeneous acid-catalyzed process was by neous acid-catalyzed processes must increase to almost
far the most economically attractive process. 100% before there is any overlap in the after tax rate-of-
return. From this, it is clear that even at reduced reactor
6. Sensitivity analyses and optimization conversion, the heterogeneous process will still be advanta-
geous over the supercritical and homogeneous acid-cata-
Sensitivity analyses were conducted to determine the lyzed processes.
effect on the process of variables that had some degree of Sensitivity analyses were performed for all processes to
uncertainty; and to identify any operating specifications determine the effect of changing methanol recovery in the
within an individual process that could be modified to methanol recovery distillation column on the ATROR. In
improve the process. all cases except the pre-treatment case, increasing the meth-
A.H. West et al. / Bioresource Technology 99 (2008) 6587–6601 6599

75

45

15

-15

ATROR (%) -45

-75

-105

-135

-165
80 85 90 95 100 105
Reaction Conversion (%)
Base Catalyzed Homog. Acid Cat. Heterog. Acid Cat. Supercritical

Fig. 5. After tax rate-of-return vs. reaction conversion for all processes.

anol recovery caused an increase in the ATROR, due to was offset by the increase in revenue that results from
decreased methanol consumption in all cases. Methanol higher methanol recovery. As shown in Fig. 7, the addition
acts as a cosolvent (Chiu et al., 2005) increasing the solubil- of the vacuum system resulted in a decrease in ATROR.
ity of biodiesel in the glycerol phases. Therefore, reducing However, as the methanol recovery was increased under
the amount of methanol entering the three phase separator vacuum operation the ATROR increased, indicating the
(HAC and SC processes) reduced the amount of biodiesel potential for optimization of the column operating condi-
lost in the glycerol stream, thereby boosting ATROR for tions to maximize the ATROR. Similar analyses were
both processes. The effect of methanol recovery on conducted for the homogeneous acid-catalyzed and super-
ATROR for the HAC process is illustrated in Fig. 6. Meth- critical processes, but it was found that vacuum operation
anol recovery is limited to about 85%, as the bottoms did not provide any economic benefits, as the methanol
stream temperature should not exceed 150 °C to prevent recovery was already greater than 99% and the bottoms
glycerol decomposition. In order to increase the methanol temperature was within the allowable limit at ambient pres-
recovery, the distillation column was operated under vac- sure operation. The HYSYS optimizer tool was used to
uum conditions. The effect of vacuum pressure (and there- vary the HAC methanol recovery in order to maximize
fore cost of the vacuum system) on the ATROR was the ATROR, according to the following constraints: bot-
investigated to determine if the cost of the vacuum system toms temperature <150 °C; 1 kPa < column pressure

58 180

57 160
Bottoms Temperature (°C)

56 140

120
55
ATROR (%)

100
54
80
53
60
52
40
51 20
50 0
0.75 0.8 0.85 0.9 0.95 1
Methanol Recovery (%)

ATROR Temperature

Fig. 6. ATROR vs. methanol recovery in the methanol recovery column, HAC process.
6600 A.H. West et al. / Bioresource Technology 99 (2008) 6587–6601

57
160
56

55

Bottoms Temperature (°C)


150

54
140
ATROR (%)

53

52 130

51
120
50
110
49

48 100
10 20 30 40 50 60 70 80 90 100 110
Operating Pressure (kPa)

ATROR Temperature

Fig. 7. ATROR vs. operating pressure in the methanol recovery column, HAC process.

<100 kPa; and 85.0% < methanol recovery <99.9%. An sizes and carbon steel construction of most of the process
optimum was found at a pressure equal to 40 kPa and equipment. Raw materials consumed in the process
methanol recovery of 99.9%. Upon optimization the bot- account for a major portion of the total manufacturing
toms temperature decreased from 149.9 °C to 145.5 °C, cost. Accordingly, Processes II, III and IV have much
methanol recovery increased from 85% to 99.9% and the lower manufacturing costs than Process I. The high solvent
ATROR increased slightly from 53.7% to 54.2%. In addi- (glycerol for washing) and catalyst costs of Process I, the
tion to the financial incentive, including a vacuum system large excesses of methanol in Processes II and IV resulted
reduces methanol consumption and eliminates 79,200 kg/ in much higher manufacturing costs than in Process III
year of methanol from the waste stream, greatly reducing making it the only process to produce a net profit. The
the environmental impact of the process. after tax rate-of-return for Process III was 59%, while Pro-
Lastly, the effect of vacuum operation in the FAME dis- cesses I, II and IV had rates-of-return of 22%, 9% and
tillation columns was investigated for the heterogeneous 1%, respectively.
acid-catalyzed and the supercritical processes, to determine Sensitivity analyses were conducted to identify any unit
if vacuum operation would result in a net savings due to a operations where operating conditions could be modified
decrease in the heating and cooling duties on the column. to improve the process. Increasing methanol recovery led
Column heating and cooling loads did decrease; however, to a greater ATROR. Accordingly, methanol recovery
the savings in utilities cost were not enough to offset the was set as high as possible (>99%) before the glycerol deg-
cost of the vacuum system, and inclusion of a vacuum sys- radation temperature (150 °C) was exceeded in the homo-
tem therefore decreased the ATROR in both cases. Since geneous acid-catalyzed and supercritical processes. Use of
the upper temperature limit of glycerol (150 °C) has not the optimizer function indicated a vacuum system could
been exceeded at ambient operation, a vacuum system be installed in the HAC process to increase methanol
was deemed unnecessary for FAME distillation in both recovery and consequently the ATROR, while keeping
processes. Vacuum operation for FAME distillation was the bottoms stream within the temperature limit.
needed in the homogeneous acid-catalyzed process to keep An analysis of the effect of reaction conversion on
the temperature of the distillate below 250 °C. ATROR revealed that even at reduced reaction conversion
(i.e., between 85% and 93%), the ATROR of the HAC pro-
7. Conclusion cess is greater than at 100% conversion of the homoge-
neous acid and supercritical processes.
Four continuous processes to produce biodiesel at a rate Therefore Process III, the heterogeneous acid-catalyzed
of 8000 tonnes/year were designed and simulated in process, is clearly advantageous over the other processes,
HYSYS: all were capable of producing ASTM grade as it had the highest rate-of-return, lowest capital invest-
biodiesel. Subsequently, an economic assessment revealed ment, and technically, was a relatively simple process.
that the heterogeneous acid-catalyzed process has the low- Further research in developing the heterogeneous acid-
est total capital investment, owing to the relatively small catalyzed process for biodiesel production is warranted.
A.H. West et al. / Bioresource Technology 99 (2008) 6587–6601 6601

Acknowledgement Haas, M.J., McAloon, A.J., Yee, W.C., Foglia, T.A., 2006. A process
model to estimate biodiesel production costs. Bioresource Technology
97, 671–678.
The authors acknowledge the financial support of the Krawczyk, T., 1996. Biodiesel. INFORM 7, 801–822.
Natural Sciences and Engineering Research Council of Kusdiana, D., Saka, S., 2001. Kinetics of transesterification in
Canada. rapeseed oil to biodiesel fuel as treated in supercritical methanol.
Fuel 80, 693–698.
References Kusdiana, D., Saka, S., 2004. Effects of water on biodiesel fuel production
by supercritical methanol treatment. Bioresource Technology 91, 289–
295.
Abreu, F.R., Alves, M.B., Macedo, C.C.S., Zara, L.F., Suarez, P.A.Z.,
Lotero, E., Liu, Y.J., Lopez, D.E., Suwannakarn, K., Bruce, D.A.,
2005. New multi-phase catalytic systems based on tin compounds
Goodwin, J.G., 2005. Synthesis of biodiesel via acid catalysis.
active for vegetable oil transesterification reaction. Journal of Molec-
Industrial & Engineering Chemistry Research 44, 5353–5363.
ular Catalysis A: Chemical 227, 263–267.
Luyben, W.L., 2002. Plantwide Dynamic Simulators in Chemical Pro-
Canakci, M., Van Gerpen, J., 1999. Biodiesel production via acid
cessing and Control. Marcel Dekker, New York.
catalysis. Transactions of the ASAE 42, 1203–1210.
Ma, F.R., Hanna, M.A., 1999. Biodiesel production: a review. Biore-
Canakci, M., Van Gerpen, J., 2001. Biodiesel production from oils and
source Technology 70, 1–15.
fats with high free fatty acids. Transactions of the ASAE 44, 1429–
McCabe, W.J., Smith, J.C., Harriott, P., 2001. Unit Operations of
1436.
Chemical Engineering, sixth ed. McGraw Hill, Boston.
Canakci, M., Van Gerpen, J., 2003. A pilot plant to produce biodiesel
Newman, A.A., 1968. Glycerol. CRC Press, Cleveland.
from high free fatty acid feedstocks. Transactions of the ASAE 46,
Saka, S., Kusdiana, D., 2001. Biodiesel fuel from rapeseed oil as prepared
945–954.
in supercritical methanol. Fuel 80, 225–231.
Chiu, C.W., Goff, M.J., Suppes, G.J., 2005. Distribution of methanol and
Schumacher, L.G., Marshall, W., Krahl, J., Wetherell, W.B., Grabowski,
catalysts between biodiesel and glycerin phases. AIChE Journal 51,
M.S., 2001. Biodiesel emissions data from Series 60 DDC engines.
1274–1278.
Transactions of the ASAE 44, 1465–1468.
Delfort, B., Hillion, G., Le Pennec, D., Chodorge, J.A., Bournay, L.,
Seider, W.D., Seader, D., Lewin, D.R., 2003. Process Design Principles:
2003. Biodiesel production by a continuous process using a heteroge-
Synthesis, Analysis and Evaluation, second ed. John Wiley & Sons,
neous catalyst. Abstracts of Papers, 226th ACS National Meeting,
Chichester, UK.
New York, NY, United States, September 7–11, 2003, FUEL-066.
Suppes, G.J., Dasari, M.A., Doskocil, E.J., Mankidy, P.J., Goff, M.J.,
Demirbas, A., 2002. Biodiesel from vegetable oils via transesterification in
2004. Transesterification of soybean oil with zeolite and metal
supercritical methanol. Energy Conversion and Management 43,
catalysts. Applied Catalysis A: General 257, 213–223.
2349–2356.
Turton, R., Bailie, R.C., Whiting, W.B., Shaeiwitz, J.A., 2003. Analysis,
Freedman, B., Pryde, E.H., Mounts, T.L., 1984. Variables affecting the
Synthesis, and Design of Chemical Processes, 2nd ed. Prentice Hall,
yields of fatty esters from transesterified vegetable-oils. Journal of the
Upper Saddle River, New Jersey.
American Oil Chemists Society 61, 1638–1643.
Tyson, K.S., 2001. Biodiesel: Handling and Use Guidelines. National
Freedman, B., Butterfield, R.O., Pryde, E.H., 1986. Transesterification
Renewable Energy Laboratory, Golden, CO.
kinetics of soybean oil. Journal of the American Oil Chemists Society
Warabi, Y., Kusdiana, D., Saka, S., 2004. Reactivity of triglycerides and
63, 1375–1380.
fatty acids of rapeseed oil in supercritical alcohols. Bioresource
Furuta, S., Matsuhashi, H., Arata, K., 2004. Biodiesel fuel production
Technology 91, 283–287.
with solid superacid catalysis in fixed bed reactor under atmospheric
Zhang, Y., Dube, M.A., McLean, D.D., Kates, M., 2003a. Biodiesel
pressure. Catalysis Communications 5, 721–723.
production from waste cooking oil: 1. Process design and technological
Goff, M.J., Bauer, N.S., Lopes, S., Sutterlin, W.R., Suppes, G.J., 2004.
assessment. Bioresource Technology 89, 1–16.
Acid-catalyzed alcoholysis of soybean oil. Journal of the American Oil
Zhang, Y., Dube, M.A., McLean, D.D., Kates, M., 2003b. Biodiesel
Chemists Society 81, 415–420.
production from waste cooking oil: 2. Economic assessment and
Goodrum, J.W., 2002. Volatility and boiling points of biodiesel from
sensitivity analysis. Bioresource Technology 90, 229–240.
vegetable oils and tallow. Biomass & Bioenergy 22, 205–211.

Você também pode gostar