Vous êtes sur la page 1sur 28

Computers and Chemical Engineering 47 (2012) 2956

Contents lists available at SciVerse ScienceDirect

Computers and Chemical Engineering


journal homepage: www.elsevier.com/locate/compchemeng

Process synthesis of hybrid coal, biomass, and natural gas to liquids via
FischerTropsch synthesis, ZSM-5 catalytic conversion, methanol synthesis,
methanol-to-gasoline, and methanol-to-olens/distillate technologies
Richard C. Baliban, Josephine A. Elia, Vern Weekman, Christodoulos A. Floudas
Department of Chemical and Biological Engineering, Princeton University, Princeton, NJ 08544, USA

a r t i c l e

i n f o

Article history:
Received 2 April 2012
Received in revised form 24 June 2012
Accepted 25 June 2012
Available online 3 July 2012
Keywords:
Process synthesis with heat, power, and
water integration
Hybrid energy systems
Mixed-integer nonlinear optimization
FischerTropsch
Methanol to gasoline
Methanol to olens and distillate

a b s t r a c t
Several technologies for synthesis gas (syngas) rening are introduced into a thermochemical based
superstructure that will convert biomass, coal, and natural gas to liquid transportation fuels using
FischerTropsch (FT) synthesis or methanol synthesis. The FT efuent can be (i) rened into gasoline,
diesel, and kerosene or (ii) catalytically converted to gasoline and distillate over a ZSM-5 zeolite. Methanol
can be converted using ZSM-5 (i) directly to gasoline or to (ii) distillate via olen intermediates. A mixedinteger nonlinear optimization model that includes simultaneous heat, power, and water integration
is solved to global optimality to determine the process topologies that will produce the liquid fuels at
the lowest cost. Twenty-four case studies consisting of different (a) liquid fuel combinations, (b) renery capacities, and (c) superstructure possibilities are analyzed to identify important process topological
differences and their effect on the overall system cost, the process material/energy balances, and the
well-to-wheel greenhouse gas emissions.
2012 Elsevier Ltd. All rights reserved.

1. Introduction
The transportation sector relies heavily on petroleum as a primary energy source, and currently faces challenges over high crude
oil prices, volatility of the global oil market, and greenhouse gas
emissions. These issues may be further compounded if the concern of domestic peak oil production is to be realized within the
coming decades (Nashawi, Malallah, & Al-Bisharah, 2010; Zittel &
Schindler, 2007). As a result, a substantial effort has been taken in
the United States to focus on greater energy independence through
the utilization of a more diverse array of primary energy sources. Of
particular interest are liquid fuels that can be derived from domestic carbon-based feedstocks which can directly replace a proportion
of the fuels that are produced from crude imports from undesirable
or unstable governments (Weekman, 2010). The United States has
three major carbon-based feedstocks, namely coal, biomass, and
natural gas, which can be converted into liquid transportation fuels
through a variety of means. A recent review has highlighted the
process design alternatives that can produce gasoline, diesel, and
kerosene using any one or a combination of the three feedstocks
(Floudas, Elia, & Baliban, 2012).

Corresponding author. Tel.: +1 609 258 4595; fax: +1 609 258 0211.
E-mail address: oudas@titan.princeton.edu (C.A. Floudas).
0098-1354/$ see front matter 2012 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.compchemeng.2012.06.032

Each of the three feedstocks has signicant advantages


and disadvantages when used to generate liquid transportation
fuels. Coal is benecial because the delivered cost is generally
cheaper ($2.02.5/MM Btu) (Energy Information Administration,
2011) than natural gas ($4.85.8/MM Btu) (Energy Information
Administration, 2011) or biomass ($4.09.0/MM Btu) (Kreutz,
Larson, Liu, & Williams, 2008; Larson, Jin, & Celik, 2009; National
Academy of Sciences, 2009). The prices of all feedstocks are converted to 2009$ using the GDP deator index (US Government
Printing Ofce, 2009). However, the high carbon content of coal
may require that a signicant portion of the feedstock carbon is
converted to CO2 where it can either be vented, sequestered, or
converted back to CO using a non-carbon based source of hydrogen
(Agrawal, Singh, Ribeiro, & Delgass, 2007; Baliban, Elia, & Floudas,
2010, 2011, 2012; Elia, Baliban, & Floudas, 2010, 2012; Elia, Baliban,
Xiao, & Floudas, 2011). Natural gas provides a high hydrogen to
carbon ratio that can help to increase the conversion rates of feedstock carbon to nal liquid fuels. Recent prospects for shale gas
production have helped reduce the delivered cost of natural gas
and made this feedstock a more attractive choice for liquid fuels
production (Energy Information Administration, 2011). Biomass is
highly benecial since it is a renewable energy source that can
absorb atmospheric CO2 during photosynthesis (Lynd et al., 2009;
National Academy of Sciences, 2009; Water Science & Technology
Board, 2008). Though corn-based ethanol and soybean-based diesel
comprise a majority of the biofuels manufactured today, their use

30

R.C. Baliban et al. / Computers and Chemical Engineering 47 (2012) 2956

for fuel production has led to concerns regarding the impact on


the price and availability of these feedstocks as sources of food
(Lynd et al., 2009). Lignocellulosic plant sources (e.g., corn stover
or forest residue) are expected to be a more considerable source
of biofuels in the future, though an increase in crop production
will be required to generate an appropriate amount of sustainable
residue for fuels production (de Fraiture, Giordano, & Liao, 2008;
Department of Energy, 2005; National Research Council, 2008).
Current alternative energy processes in the literature focus on
using a single feedstock or a hybrid combination of feedstocks to
combine the benets of each feed (Floudas et al., 2012). Generally,
these designs have focused on generation of synthesis gas (syngas) from the raw materials which is subsequently converted to
the nal liquid fuels via the FischerTropsch (FT) process. Examples of hybrid systems are coal and biomass to liquids (Agrawal
et al., 2007; Chen, Adams, & Barton, 2011a, 2011b; Kreutz et al.,
2008), coal and natural gas to liquids (Adams & Barton, 2010; Cao
et al., 2008; Sudiro & Bertucco, 2007, 2009), biomass and natural gas to liquids, and coal, biomass, and natural gas to liquids
(Baliban, Elia, & Floudas, 2012; Baliban, Elia, Misener, & Floudas,
2012; Baliban et al., 2010, 2011; Elia et al., 2010). Polygeneration
systems have also been designed which are able to co-produce
electricity and other liquid fuels along with gasoline, diesel, and
kerosene (Adams & Barton, 2011; Baliban, Elia, & Floudas, 2012;
Baliban, Elia, Misener, et al., 2012; Baliban et al., 2011; Chen
et al., 2011a, 2011b; Chiesa, Consonni, Kreutz, & Williams, 2005;
Kreutz, Williams, Consonni, & Chiesa, 2005; Liu, Pistikopoulos, & Li,
2009, 2010a, 2010b; Sudiro, Bertucco, Ruggeri, & Fontana, 2008).
In addition to the FT process, liquid fuels can be generated from
syngas by rst converting the syngas to methanol and then converting the methanol to liquid hydrocarbons (Keil, 1999; Tabak &
Yurchak, 1990). The methanol to gasoline (MTG) (Mobil Research &
Development Corporation, 1978), methanol to olens (MTO), and
Mobil olens to gasoline/distillate (MOGD) (Tabak, Krambeck, &
Garwood, 1986; Tabak & Yurchak, 1990) processes developed by
Mobil in the 1970s and 1980s utilize a zeolite catalyst to convert the methanol to gasoline and distillate range hydrocarbons.
To date, the major process designs involving methanol conversion
are based on single feedstock systems including coal to liquids and
biomass to liquids (Mobil Research & Development Corporation,
1978; National Renewable Energy Laboratory, 2011). No hybrid
feedstock design for methanol synthesis and conversion to liquid
fuels has been considered.
The previous studies typically focus on process designs where
the topology of the process is xed. A process simulation is then
conducted to determine the heat and mass balances for the process
and an economic analysis is performed to determine the viability
of the plant (Adams & Barton, 2010, 2011; Baliban et al., 2010;
Bechtel Corp, 2003; Cao et al., 2008; Chen et al., 2011a, 2011b;
Chiesa et al., 2005; Elia et al., 2010; Kreutz et al., 2005, 2008; Larson
& Jin, 1999; Liu et al., 2009, 2010a, 2010b; Sudiro & Bertucco, 2007,
2009; Vliet, Faaij, & Turkenburg, 2009). Recently, optimizationbased process synthesis strategies have been developed where a
superstructure of process topologies is postulated and the design
which produces the liquid fuels at the lowest cost or highest prot
is selected (Baliban, Elia, & Floudas, 2012; Baliban, Elia, Misener,
et al., 2012; Baliban et al., 2011; Martin & Grossmann, 2011). A comprehensive process synthesis strategy was proposed which uses a
hybrid feedstock of coal, biomass, and natural gas to produce the
desired liquid fuels (CBGTL) (Baliban, Elia, & Floudas, 2012; Baliban
et al., 2011), though the methodology could be tailored to analyze any one feedstock or combination of feedstocks. A rigorous
global optimization framework (Baliban, Elia, Misener, et al., 2012)
was also introduced to theoretically guarantee that the cost (or
prot) achieved by the optimal design is within a small percentage of the best value possible. The process synthesis strategy is

capable of directly examining the technoeconomic, environmental, and process topological trade-offs between each design and
output the solution with the best economic value. The process synthesis framework was enhanced by including a simultaneous heat
and power integration (Baliban et al., 2011) using an optimizationbased heat-integration approach (Duran & Grossmann, 1986) and
a series of heat engines that can convert waste heat into electricity (Baliban et al., 2010, 2011; Elia et al., 2010). Additionally, the
process synthesis model integrated a comprehensive wastewater
treatment network (Baliban, Elia, & Floudas, 2012) that utilized a
superstructure approach (Ahmetovic & Grossmann, 2010a, 2010b;
Grossmann & Martn, 2010; Karuppiah & Grossmann, 2006) to
determine the appropriate topology and operating conditions of
process units that are needed to minimize wastewater contaminants and freshwater intake.
This manuscript introduces several distinct methods for conversion of syngas to liquid fuels into the CBGTL process superstructure
and investigates the tradeoffs that arise from these methods. The
previous superstructure converted the syngas into a raw FT hydrocarbon product using one of four FT units operating with either
a cobalt or iron catalyst and at high or low temperature (Baliban,
Elia, & Floudas, 2012; Baliban et al., 2011). The efuent was subsequently fractionated and upgraded using a series of hydrotreating
units, a wax hydrocracker, two isomerization units, a naphtha
reformer, an alkylation unit, and a gas separation plant (i.e., deethanizer). This study introduces two iron-based FT units that utilize
the forward watergas-shift reaction to produce the raw hydrocarbons using an input H2 /CO ratio that is less than the typical 2/1
ratio needed for FT synthesis. Catalytic conversion of the FT vapor
efuent over a ZSM-5 catalyst is considered as an alternative for
producing gasoline range hydrocarbons from the raw FT efuent.
Methanol synthesis and subsequent conversion to liquid hydrocarbons are also introduced into the superstructure. The methanol
may be catalytically converted using a ZSM-5 zeolite to (i) gasoline
range hydrocarbons or (ii) to distillate (i.e., diesel and kerosene)
via an intermediate coversion to olens. The mathematical modeling and cost functions needed to incorporate the above alternatives
into the superstructure are outlined in detail. The complete process
synthesis optimization model is then tested on a total of 24 case
studies which consist of two liquid product combinations, three
plant capacities, and four plant superstructures. Using low-volatile
bituminous coal (Illinois #6) and perennial biomass (switchgrass),
important topological differences between the case studies are discussed and the results of each component of the process synthesis
framework are illustrated.
2. CBGTL mathematical model for process synthesis with
simultaneous heat, power, and water integration
This section will discuss the enhancements to the previous
mathematical model for process synthesis and simultaneous heat,
power, and water integration that will incorporate a wide variety of designs for syngas conversion and hydrocarbon upgrading.
Modeling of these enhancements will be described in detail in the
following section and the complete mathematical model is listed in
Appendix A. The nomenclature used in the mathematical description below is outlined in Table 1. Note that this table represents a
subset of the comprehensive list of symbols that are needed for the
full mathematical model. The full list of symbols and mathematical
model are included for reference in Appendix A.
2.1. Conceptual design
The syngas conversion and hydrocarbon upgrading units proposed in this paper is based on an extension of the CBGTL

R.C. Baliban et al. / Computers and Chemical Engineering 47 (2012) 2956

Clean Gas
(to MEOHS)

31

SPFTM

CO2 Lean
Light Gases
(from CO2SEP)
SPCG

Clean Gas
(from AGR)

Hydrogen
(from XH2-LTFT)

SPFTLT

SPFTN
Hydrogen
(from XH2-HTFT)

SPFTHT

LTFT
Cobalt HT NomWax FT

Steam
(from SPSTM)
MTFTWGS-N
Iron MT fWGS
Nom-Wax FT

CO2
(from XCO2-HTFT)

Reformed Gases
(from SPATR)

Wax
(to WHC)

Wax
(to WHC)

Hydrogen
(from XH2-MTFTM)
CO2
(from XCO2-LTFT)

FT Hydrocarbons
(to SPFTH)

Hydrogen
(from XH2-HTFT)
FT Hydrocarbons
(to SPFTH)

MTFTWGS-M
Iron-MT fWGS
Min-Wax FT

Hydrogen
(from XH2-MTFTN)

PSA Offgas
(from SPPSA)

LTFTRGS
Iron HT rWGS
Nom-Wax FT

HTFT
Cobalt LT MinWax FT

Steam
(from SPSTM)

PSA Offgas
(from SPPSA)
Reformed Gases
(from SPATR)

FT Hydrocarbons
(to SPFTH)

HTFTRGS
Iron LT rWGS
Min-Wax FT

Hydrogen
(from XH2-HTFTRGS)

FT Hydrocarbons
(to SPFTH)

Fig. 1. FischerTropsch (FT) hydrocarbon production owsheet. Each of the six FT units has a distinct set of operating conditions including catalyst type (cobalt or iron),
temperature (low 240 C, medium 267 C, and high 320 C), and watergas-shift reaction extent (forward, reverse, or none). Each unit is designed to produce either a
minimal or nominal amount of wax (shown as a dashed line). The mathematical model will select at most two types of the six FT units to operate in a nal process topology.

MXHRC

FT Hydrocarbons
VLWS
Vapor / Liquid /
Water Separator

SPFTH

WSOS
Water Soluble
Oxygenates Sep.

Oxygenates

FT-ZSM5
ZSM-5 Unit

Distillate
(to DHT)

ZSM5F
ZSM-5 Product
Fractionation
Raw ZSM-5
HC Product
(to SPFTZSM)

Water Lean FT
Hydrocarbons
(to HRC)

VPOS
Vapor Phase
Oxygenates Sep.
MXFTWW

Oxygenates

Wastewater
(to MXSS)

Raw
Product

Sour Water
(to MXPUWW)

Fig. 2. First FischerTropsch (FT) hydrocarbon upgrading owsheet. The FT efuent may be passed through a series of stripper and ash units to separate the oxygenates
and aqueous phase from the hydrocarbons. Alternatively, the efuent may be passed over a ZSM-5 catalytic reactor to convert most of the hydrocarbons into gasoline range
species. The raw ZSM-5 product is then fractionated to remove any distillate or sour water from the gasoline product.

32

R.C. Baliban et al. / Computers and Chemical Engineering 47 (2012) 2956

C4 Gases

Hydrogen
(from SPH2)

C4I
C4 Isomerizer

INBUT
Input
Butane

Light Offgas

SGP
Saturated Gas
Plant

C4 Gases

Isomerized C4
Gases

OUTLPG
Output LPG

SPLPG

MXSGP

Light
Offgas

C345A
C3/C4/C5
Alkylizer

Light Gases
(to SPLG)

Offgas (from other units)

C3/C4/C5
Water Lean FT
Alkylate
Gases
Hydrogen
Hydrocarbons
(from SPH2)
gas
Light
Off
(from MXHRC)
HRC
Sour Water
Hydrocarbon
(to MXPUWW )
Recovery System
Offgas
Hydrogen
NHT
Distillate
Naphtha
(to SGP)
(from SPH2)
Naphtha
Offgas
Hydrotreater
Offgas
Sour Water
Kerosene
(to SGP)
Treated
(to SGP)
(to MXPUWW )
DHT
Naphtha
KHT
Distillate
Distillate
GBL
Kerosene
Hydrotreater
Gasoline
Hydrotreater
Kerosene
Reformate
Blender
Wax
Sour Water
OUTGAS
(from MXWAX)
Hydrogen
NRF
(to MXPUWW )
OUTKER
Output
(from SPH2)
Naphtha
Output
Gasoline
C5/C6
Diesel
Cracked
Reformer
Offgas
Kerosene
Isomerate
Gases
Naphtha
DBL
(to SGP)
Offgas
WHC
Offgas
C56I
Diesel
(to SGP)
Wax
(to SGP)
C5/C6
Blender
Hydrocracker
Isomerizer
/C
Gases
C
5 6
Diesel
OUTDIE
Hydrogen
Sour Water
Output
7 en
Hydrog
(from SPH2)
(to MXPUWW)
Diesel
(from SPH2)
Fig. 3. Second FischerTropsch (FT) hydrocarbon upgrading owsheet. The water lean FT efuent is fractionated and passed through a series of treatment units to recover
the gasoline, diesel, and kerosene products along with some LPG byproduct. Light gases (i.e., unreacted syngas and C1 C2 hydrocarbons) are collected and recycled back to
the process.

H2
(from SPH2)

Clean Gas
(from SPCG)

Offgas
(to SPLG)

MEOHS
Methanol
Synthesis

MTO-F
Olefin Fractionation

Raw Methanol
Product

MTO
Methanol to Olefins

MEOH-F
Methanol Flash
Raw
Olefins
MEDEG
Methanol
Degasser

Light Gases
(to SPLG)

OUTGAS
Output Gasoline
Wastewater
(to MXPUWW)
Distillate
(to DHT)

Olefins

Purified
Methanol
SPMEOH

Offgas
(to SPLG)

OGD
Olefins to Gasoline
and Distillate

MTG
Methanol to
Gasoline
Wastewater
(to MXPUWW)

Raw
Product

MTODF
Hydrocarbon
Fractionation

Kerosene
(to KHT)

Raw MTOD
HC Product
(to SPMTODHC)

Raw MTG HC
Product
(to SPMTGHC)

Fig. 4. Methanol synthesis and conversion owsheet. Clean syngas is initially converted to methanol and then split to either the methanol to gasoline (MTG) or methanol to
olens (MTO) processes. The two processes utilize a ZSM-5 zeolite to convert the methanol to either gasoline range hydrocarbons (MTG) or olens which are subsequently
oligomerized to gasoline and distillate range hydrocarbons (MOGD). The distillate is hydrotreated to form diesel or kerosene which the gasoline range hydrocarbons are sent
to an LPG-gasoline separation system.

R.C. Baliban et al. / Computers and Chemical Engineering 47 (2012) 2956

33

Raw ZSM-5 HC
Product
(from ZSM5F)

Raw MTG
HC Product
(from MTG)

SPMTGHC

C3/C4 gases

HCKO1
Mixed HC
Condensation
Knockout 1
CO2 Rich
Light gases
(to SPLG)
SPMTODHC

OUTLPG
Output LPG

INBUT
Input Butane

SPFTZSM

C3+ HC

LPG /
Alkylate

ALK-UN
HF Alkylation
Unit

STA-COL
Stabilizer Column

LPG-ALK
LPG / Alkylate
Splitter

OUTGAS
Output Gasoline

SP-COL
Splitter Column
Light gases

Crude
HC
DEETH
Deethanizer

Light HC
gas

Lean oil
recycle
ABS-COL
Absorber Column

Reflux

Raw MTOD
HC Product
(from MTODF)
HCKO2
Mixed HC
Condensation
Knockout 2

Light gases

CO2SEP
CO2 Separation Unit

CO2 Lean
Light gases
(to SPCG)

Recovered CO2
(to MXCO2C)

Fig. 5. LPG-gasoline product separation owsheet. The raw HC products from the FT-ZSM5 unit, the MTG unit, or the MOGD process are passed through a series of separation
units to recover a gasoline product and an LPG byproduct. Light gases are recycled back to the renery and CO2 recovery may be utilized in preparation for sequestration or
reaction with H2 via the reverse watergas-shift reaction.

renery superstructure detailed by Baliban, Elia, and Floudas


(2012), Baliban et al. (2011). All relevant thermodynamic information (i.e., chemical equilibrium constants, vapor-liquid equilibrium
constants, specic enthalpies, and heat capacities) for the units
and streams in the renery have been extracted from Aspen
Plus v7.3 using the PengRobinson equation of state with the
BostonMathias alpha function. The owsheets depicting the
extensions of the superstructure are shown in Figs. 15 of this
text and the complete superstructure is included as Supplementary
Material. In the gures, xed process units are represented in a
light blue color, variable process units in dark blue, splitter units
in gold, and mixer units in green. Fixed process streams will have
a gray color while the variable process streams are in blue. Note
that some units (e.g., compressors, pumps, heat exchangers) are not
included in the gures for clarity, though these units are thoroughly

Table 1
Mathematical model nomenclature.
Symbol

Denition

Indices
s
u

Species index
Process unit index

Sets
(u, u )
(u, u , s)
Ir
u UFT

Stream from unit u to unit u


Species s within stream (u, u )
Set of all iron-based FT units

Parameters
KuWGS
KuMSN

Watergas-shift equilibrium constant for unit u


Methanol synthesis equilibrium constant for unit u

Variables
S
Nu,u
 ,s
S
xu,u
 ,s

Molar ow of species s from unit u to unit u


Molar concentration of species s from unit u to unit u

modeled in the CBGTL renery, as noted in previous studies


(Baliban, Elia, & Floudas, 2012; Baliban, Elia, Misener, et al., 2012;
Baliban et al., 2011).
The CBGTL superstructure is designed to co-feed biomass, coal,
or natural gas to produce gasoline, diesel, and kerosene. Syngas
is generated via gasication from biomass (Supp. Fig. S1) or coal
(Supp. Fig. S2) or auto-thermal reforming of natural gas (Supp. Fig.
S10). Co-feeding of the coal, biomass, or natural gas in a single
gasier unit was not considered in this study due to the lack of
(i) technical maturity of the process design and (ii) cost and operating data for co-fed units. Synergy for co-fed biomass and coal
gasication and simultaneous reforming the natural gas using the
gasier quench heat (Adams & Barton, 2011) may be important to
reduce the capital cost required for synthesis gas production, and
the authors note that the optimization model is capable of including the technoeconomic benet of co-fed gasication if cost and
operational data become available.
The synthesis gas is either (i) converted into hydrocarbon products in the FischerTropsch (FT) reactors (Fig. 1; Supp. Fig. S5) or (ii)
into methanol via methanol synthesis (Fig. 4; Supp. Fig. S8). The FT
wax will be sent to a hydrocracker to produce distillate and naphtha
(Supp. Fig. S7) while the FT vapor efuent may be (a) fractionated
and upgraded into gasoline, diesel, or kerosene or (Fig. 3; Supp.
Fig. S7) (b) catalytically converted to gasoline via a ZSM-5 zeolite
(Fig. 2; Supp. Fig. S6). The methanol may be either (a) catalytically
converted to gasoline via the ZSM-5 catalyst (Figs. 45; Supp. Figs.
S8S9) or (b) catalytically converted to olens via the ZSM-5 catalyst and subsequently fractionated to distillate and gasoline (Fig. 4;
Supp. Fig. S8).
Acid gases including CO2 , H2 , and NH3 are removed from the
syngas via a Rectisol unit prior to conversion to hydrocarbons or
methanol (Supp. Fig. S3). Incorporation of other acid gas removal
technologies (e.g., amine absorption, pressure-swing absorption,

34

R.C. Baliban et al. / Computers and Chemical Engineering 47 (2012) 2956

vacuum-swing absorption, membrane separation) and their relative capital/operating cost as a function of input ow rate and acid
gas concentration is the subject of an ongoing study. The sulfur-rich
gases are directed to a Claus recovery process (Supp. Fig. S4) and
the recovered CO2 may be sequestered (Supp. Fig. S3) or reacted
with H2 via the reverse watergas-shift reaction. The CO2 may
be directed to either the gasiers (Supp. Figs. S1S2), the reverse
watergas-shift reactor (Supp. Fig. S3), or the iron-based FT units
(Supp. Fig. S5). Recovered CO2 is not sent to the cobalt-based FT
units to ensure a maximum molar concentration of 3% and prevent
poisoning of the catalyst. Hydrogen is produced via pressure-swing
absorption or an electrolyzer unit while oxygen can be provided by
the electrolyzer or a distinct air separation unit (Supp. Fig. S11). A
complete water treatment network (Supp. Figs. S12S13) is incorporated that will treat and recycle wastewater from various process
units, blowdown from the cooling tower, blowdown from the boilers, and input freshwater. Clean output of the network includes (i)
process water to the electrolyzers, (ii) steam to the gasiers, autothermal reactor, and watergas-shift reactor, and (iii) discharged
wastewater to the environment.
The efuent of each reactor in the CBGTL renery is based
on either (i) known extents of reaction, (ii) thermodynamicallylimited equilibrium, or (iii) a specied composition from a literature
source. Reaction system (i) is used in the gasiers, the tar cracker,
and the combustor units (e.g., fuel combustor, gas turbine, Claus
combustor) and the extents of reaction are based on known information from literature (gasiers/cracker) or from the operating
conditions of the unit (i.e., complete combustion using a stoichiometric excess of oxygen). Reaction system (ii) is used for
the watergas-shift reaction (i.e., gasiers, WGS reactor, FT units,
methanol synthesis, auto-thermal reactor), methanol synthesis,
and steam reforming in the auto-thermal reactor. Reaction system
(iii) is used for the FT units, the ZSM-5 hydrocarbon conversion, the
MTG reactor, the MTO reactor, and the MOGD reactor. The efuent
composition of these units is based on known commercial data or
pilot plant data for the units operating at a specied set of conditions (i.e., temperature, pressure, and feed composition). The CBGTL
process is designed to ensure that the appropriate conditions are
met within the reactor to ensure that the efuent composition that
is assumed is valid. Binary decision variables (y) are included within
the mathematical model to logically dene the existence of specic
process units (Eqs. (A.17)(A.21)). That is, if y = 0 for a particular
unit, then no heat/mass ow will be allowed through the unit and
the unit will effectively be removed from the process topology. If
y = 1 for a unit, then the heat/mass ow through the unit will be
governed by the proper operation of the unit.
2.2. FischerTropsch units
The four FT units considered in previous studies (Baliban, Elia,
& Floudas, 2012; Baliban et al., 2011) utilized either a cobalt or iron
catalyst and operated at high or low temperature. The two cobaltbased FT units would not facilitate the watergas-shift reaction
and therefore required a minimal level of CO2 input to the units
to improve the perpass conversion of CO. The two iron-based FT
units were assumed to facilitate the reverse watergas-shift reaction and therefore could consume CO2 within the unit using H2 to
produce the CO necessary for the FT reactions. A key synergy of
the reaction conditions in the latter units was the heat needed for
the reverse watergas-shift reaction that is provided by the highly
exothermic FT reaction. Though the reverse watergas-shift reaction is typically unfavorable at the lower operating temperatures
of the FT units, the reaction may be indeed facilitated through the
use of an appropriate amount of input hydrogen.
In this study, the set of possible FT units is expanded to consider
iron-based systems that will facilitate the forward watergas-shift

reaction within the units. These FT units will require a lower H2 /CO
ratio for the FT reaction because steam in the feed will be shifted to
H2 through consumption of CO. These units may be benecial since
certain syngas generation units (e.g., coal gasiers) will produce
a gas that generally has a H2 /CO ratio that is much less than the
2/1 requirement for FT synthesis (Baliban et al., 2010; Kreutz et al.,
2008). The downside of the new FT units will be the high quantity of
CO2 that is produced as a result of the watergas-shift reaction. The
framework developed for the CBGTL superstructure will directly
examine the benets and consequences for each of the six FT units
to determine which technology produces a renery with a superior
design.
Fig. 1 shows the owsheet for FT hydrocarbon production within
the superstructure. Clean gas from the acid gas removal (AGR) unit
is mixed with recycle light gases from a CO2 separator (CO2 SEP) and
split (SPCG ) to either the low-wax FT section (SPFTM ), the nominalwax FT section (SPFTN ), or methanol synthesis (MEOHS). The FT
units will operate at a pressure of 20 bar and within the temperature range of 240320 C. The cobalt-based FT units operate at
either low temperature (LTFT; 240 C) or high temperature (HTFT;
320 C) and must have a minimal amount of CO2 in the input stream.
Two iron-based FT units will facilitate the reverse watergas-shift
(rWGS) reaction and will operate at low (LTFTRGS; 240 C) and high
temperature (HTFTRGS; 320 C). The other two iron-based FT units
will use the forward reverse watergas-shift (fWGS) units, operate
at a mid-level temperature (267 C), and produce either minimal
(MTFTWGS-M) or nominal (MTFTWGS-N) amounts of wax. The
operating conditions of the FT units are summarized in Table 2.
Hydrogen may be recycled to any of the FT units to either shift
the H2 /CO ratio or the H2 /CO2 ratio to the appropriate level. Steam
may alternatively be used as a feed for the two iron-based fWGS
FT units to shift the H2 /CO ratio. CO2 may be recycled back to the
iron-based rWGS FT units to be consumed in the WGS reaction.
Similarly, the pressure-swing absorption (PSA) offgas which will
be lean in H2 may be recycled to the iron-based rWGS FT units for
consumption of the CO or CO2 . The efuent from the auto-thermal
reactor (ATR) will contain a H2 /CO ratio that is generally above 2/1,
and is therefore favorable as a feedstock for FT synthesis (National
Academy of Sciences, 2009). However, the concentration of CO2
within the ATR efuent will prevent the stream from being fed to
the cobalt-based units. The two streams exiting the FT units will
be a waxy liquid phase and a vapor phase containing a range of
hydrocarbons. The wax will be directed to a hydrocracker (WHC)
while the vapor phase is split (SPFTH ) for further processing.
Modeling of the four original FT units has been previously
described (Baliban, Elia, & Floudas, 2012; Baliban et al., 2011) and
is included in the Supplementary Material. The efuent from the
two additional FT units (iron-based FT fWGS) is based off of the
slurry phase FT units developed by Mobil Research and Development Corporation in the 1980s (Mobil Research & Development
Corporation, 1983, 1985). A H2 /CO ratio of 2/3 is desired for the
input feed (Mobil Research & Development Corporation, 1983,
1985), so a sufcient amount of steam must be added to the feed to
promote the forward watergas-shift reaction. The decomposition
of carbon from CO to hydrocarbons and CO2 is outlined in Table
42 of the minimal-wax FT report (Mobil Research & Development
Corporation, 1983) and Table VIII-2 of the nominal-wax FT report
(Mobil Research & Development Corporation, 1985), and a 90% conversion of CO in the inlet stream is assumed (Mobil Research &
Development Corporation, 1983, 1985). The syngas species exiting
the four iron-based FT reactors will be constrained by watergasshift equilibrium, as noted in Eq. (1) where (u, u ) is the stream
exiting the FT unit u.
S
Nu,u
 ,H

2O

S
WGS
S
S
Nu,u
Nu,u
 ,H Nu,u ,CO
 ,CO = Ku
2

Ir
u UFT

(1)

R.C. Baliban et al. / Computers and Chemical Engineering 47 (2012) 2956

35

Table 2
Operating conditions for the process units involved in methanol synthesis and conversion to liquid hydrocarbon fuels.
Unit

Temperature ( C)

Pressure (bar)

Conv.

LT cobalt FT synthesis
LT iron FT synthesis
MT iron FT synthesis (low wax)
MT iron FT synthesis (high wax)
HT cobalt FT synthesis
HT iron FT synthesis
ZSM-5 FT upgrading
Methanol synthesis
Methanol-to-gasoline
Methanol-to-olens
Olens-to-gasoline/distillate

240
240
267
267
320
320
408
300
400
482
300

20
20
20
20
20
20
16
50
12.8
1.1
50

80% of CO
80% of CO
90% of CO
90% of CO
80% of CO
80% of CO
100% of hydrocarbons
3040% of CO
100% of methanol
100% of methanol
100% of olens

The mathematical model will select at most two types of


FischerTropsch units to operate in the nal process design. This
constraint is added because two different kinds of FT units will be
able to supply a range of hydrocarbon species that is diverse enough
to provide a target composition of liquid products without adding
unnecessary complexity to the renery design (de Klerk, 2011).

2.3. FischerTropsch product upgrading


The vapor phase efuent from FT synthesis will contain a mixture of C1 C30+ hydrocarbons, water, and some oxygenated species.
Fig. 2 details the process owsheet used to process this efuent
stream. The stream will be split (SPFTH ) and can pass through a
series of treatment units designed to cool the stream and knock out
the water and oxygenates for treatment. Initially, the water-soluble
oxygenates are stripped (WSOS) from the stream. The stream is
then passed to a three-phase separator (VLWS) to remove the
aqueous phase from the residual vapor and any hydrocarbon liquid. Any oxygenates that are present in the vapor phase may be
removed using an additional separation unit (VSOS). The water lean
FT hydrocarbons are then sent to a hydrocarbon recovery column
for fractionation and further processing (Fig. 3). The oxygenates and
water removed from the stream are mixed (MXFTWW ) and sent to
the sour stripper mixer (MXSS ) for treatment.
The FT hydrocarbons split from SPFTH may also be passed over
a ZSM-5 catalytic reactor (FT-ZSM5) operating at 408 C and 16 bar
(Mobil Research & Development Corporation, 1983) to be converted
into mostly gasoline range hydrocarbons and some distillate (Mobil
Research & Development Corporation, 1983, 1985). The ZSM-5 unit
will be able to convert the oxygenates to additional hydrocarbons,
so no separate processing of the oxygenates will be required for the
aqueous efuent. The composition of the efuent from the ZSM-5
unit is shown in Table 43 of the minimal-wax FT reactor Mobil study
(Mobil Research & Development Corporation, 1983) and in Table
VIII-3 of the nominal-wax FT reactor Mobil study (Mobil Research &
Development Corporation, 1985). For this study, the ZSM-5 efuent
composition is assumed to be equal to the composition outlined in
the minimal-wax FT reactor study (Mobil Research & Development
Corporation, 1983). This is modeled mathematically using an atom
balance around the ZSM-5 unit and the efuent composition outlined in Table 43 of the Mobil study (Mobil Research & Development
Corporation, 1983). The raw product from FT-ZSM5 is fractionated
(ZSM5F) to separate the water and distillate from the gasoline product. The water is mixed with other wastewater knockout (MXPUWW )
and the distillate is hydrotreated (DHT) to form a diesel product.
The raw ZSM-5 HC product is sent to the LPG-gasoline separation
section for further processing (Fig. 5).
The water lean FT hydrocarbons leaving MXFTWW are sent to a
hydrocarbon recovery column (HRC), as shown in Fig. 3. The hydrocarbons are split into C3 C5 gases, naphtha, kerosene, distillate,
wax, offgas, and wastewater (Baliban et al., 2010; Bechtel, 1998).

The upgrading of each stream will follow a detailed Bechtel design


(Bechtel, 1992, 1998) which includes a wax hydrocracker (WHC),
a distillate hydrotreater (DHT), a kerosene hydrotreater (KHT), a
naphtha hydrotreater (NHT), a naphtha reformer (NRF), a C4 isomerizer (C4 I), a C5 /C6 isomerizer (C56 I), a C3 /C4 /C5 alkylation unit
(C345 A), and a saturated gas plant (SGP).
The kerosene and distillate cuts are hydrotreated in (KHT) and
(DHT), respectively, to remove sour water and form the products
kerosene and diesel. Any additional distillate or kerosene produced
in other sections of the renery will also be directed to these units
for processing. The naphtha cut is sent to a hydrotreater (NHT) to
remove sour water and separate C5 C6 gases from the treated naphtha. The wax cut is sent to a hydrocracker (WHC) where nished
diesel product is sent to the diesel blender (DBL) along with the
diesel product from (DHT). C5 C6 gases from (NHT) and (WHC) are
sent to an isomerizer (C56 I). Hydrotreated naphtha is sent to the
naphtha reformer (NRF). The C4 isomerizer (C4 I) converts in-plant
and purchased butane to isobutane, which is fed into the alkylation unit (C345 A). Purchased butane is added to the isomerizer such
that 80 wt% of the total ow entering the unit is composed of nbutane. Isomerized C4 gases are mixed with the C3 C5 gases from
the (HRC) in (C345 A), where the C3 C5 olens are converted to highoctane gasoline blending stock. The remaining butane is sent back
to (C4 I), while all light gases are mixed with the offgases from other
unit and sent to the saturated gas plant (SGP). C4 gases from (SGP)
are recycled back to the (C4 I) and a cut of the C3 gases are sold as
byproduct propane.
2.4. Methanol synthesis and conversion
The clean gas split (SPCG ) from the acid gas recovery unit may
be directed to a methanol synthesis unit (MEOHS) for conversion
of the syngas to methanol (National Renewable Energy Laboratory,
2011). The syngas exiting the acid gas recovery unit is heated up to
300 C prior to entering the MEOHS unit. The MEOHS unit operates
at a temperature of 300 C, a pressure of 51 bar, and will assume
equilibrium between the watergas-shift reaction (Eq. (2)) and
the methanol synthesis reaction (Eq. (3)) in the efuent stream
(MEOHS, u) (National Renewable Energy Laboratory, 2011).
S
NMEOHS,u,H

2O

S
WGS
S
S
NMEOHS,u,CO
= KMEOHS
NMEOHS,u,H
NMEOHS,u,CO
2

(2)

S
xMEOHS,u,CH

3 OH

MSN
S
= KMEOHS
xMEOHS,u,H

2

S
xMEOHS,u,CO

(3)

Note that the equations for watergas-shift equilibrium (Eqs.


(1) and (2)) utilize molar species ow rates while the methanol
synthesis equilibrium (Eq. (3)) and the steam reforming equilibrium (Eqs. (A.98)(A.101)) utilize molar species concentrations. The
conservation of total moles across the watergas-shift equilibrium

36

R.C. Baliban et al. / Computers and Chemical Engineering 47 (2012) 2956

allows for the use of either species molar ow rates or molar concentrations in the equilibrium reaction without a need for a total
molar ow rate variable. The mathematical model was formulated
using molar ow rates because the bilinear terms for calculation of
the concentration variables are not required for all syngas species.
The remainder of the chemical equilibrium equations do not conserve the amount of total moles, so the use of species molar ow
rates would require a total molar ow rate variable to balance the
equation. In this study, it was found to be computationally benecial to use species concentration variables to reduce the presence
of trilinear or quadrilinear terms that would arise with the use of
species molar ow rates. Note that the equilibrium constants used
in Eqs. (3) and (A.98)(A.101) have been modied from the values
extracted from Aspen to account for the increased pressure of the
units.
The state-of-technology conditions for methanol synthesis
used in this study will require a CO2 input concentration of 38% for
methanol synthesis (National Renewable Energy Laboratory, 2011),
though there could exist a potential synergy from a higher CO2
input concentration (Toyir et al., 2009). However, an increased level
of H2 may also need to be input to the reactor for consumption via
the reverse watergas-shift reaction. H2 generated via pressureswing absorption may not be appropriate if the H2 -lean offgas is
primarily used as plant fuel. Alternatively, H2 provided by electrolysis of water with a non-carbon-based form of electricity (e.g., wind
or solar) will have a high capital cost of electrolyzers coupled with a
relatively high cost of renewable-based electricity. This may offset
the reduction in capital that is achieved if a CO2 capture technology
is not needed for the synthesis gas. The technoeconomical benets
of higher levels of CO2 input to the methanol synthesis reactor will
be the subject of a future investigation. The raw methanol efuent
is cooled to 35 C and sent to a ash unit (MEOH-F) to remove over
95% of the entrained methanol through vaporliquid equilibrium.
The vapor phase is split and mostly recycled (split fraction: 95%) to
the methanol synthesis reactor to increase the yield of methanol.
The methanol leaving the MEOH-F unit is degassed (MEDEG) via
distillation to remove any light vapors. The MEDEG unit is operated as a split unit with a steam utility requirement derived through
simulation.
The puried methanol is split (SPMEOH ) to either the methanolto-gasoline (MTG) (Mobil Research & Development Corporation,
1978; National Renewable Energy Laboratory, 2011) process
or to the methanol-to-olens (MTO) and Mobil olens-togasoline/distillate (MOGD) (Keil, 1999; Tabak et al., 1986; Tabak
& Krambeck, 1985; Tabak & Yurchak, 1990) processes, both of
which were developed by Mobil Research and Development in the
1970s and 1980s. More recently, the National Renewable Energy
Laboratory performed a full design, simulation, and economic analysis of a biomass-based MTG process (National Renewable Energy
Laboratory, 2011). The MTG process will catalytically convert the
methanol to gasoline range hydrocarbons using a ZSM-5 zeolite and
a uidized bed reactor. The MTG efuent is outlined in Table 3.4.2
of the Mobil study (Mobil Research & Development Corporation,
1978) and in Process Flow Diagram P850-A1402 of the NREL study
(National Renewable Energy Laboratory, 2011). Due to the high
level of component detail provided by NREL for both the MTG unit
and the subsequent gasoline product separation units, the composition of the MTG reactor used in this study is based on the NREL
report. The MTG unit will operate adiabatically at a temperature
of 400 C and 12.8 bar. The methanol feed will be heated to 330 C
and input to the reactor at 14.5 bar. The MTG efuent will contain
44 wt% water and 56 wt% crude hydrocarbons, of which 2 wt% will
be light gas, 19 wt% will be C3 C4 gases, and 19 wt% will be C5+
gasoline (National Renewable Energy Laboratory, 2011). The crude
hydrocarbons are directed to the LPG-gasoline separation section
(Fig. 5), from which 82 wt% will be gasoline, 10 wt% will be LPG, and

the balance will be recycle gases. This is modeled mathematically in


the process synthesis model by using an atom balance around the
MTG unit and assuming a 100% conversion of the methanol entering the MTG reactor (Mobil Research & Development Corporation,
1978; National Renewable Energy Laboratory, 2011).
Any methanol entering the MTO process unit is heated to 400 C
at 1.2 bar. The MTO uidized bed reactor operates at a temperature of 482 C and a pressure of 1.2 bar (Tabak & Yurchak, 1990).
The exothermic heat of reaction within the MTO unit is controlled through generation of low-pressure steam. 100% of the
input methanol is converted into olen efuent containing 1.4 wt%
CH4 , 6.5 wt% C2 C4 parafns, 56.4 wt% C2 C4 olens, and 35.7 wt%
C5 C11 gasoline (Tabak & Yurchak, 1990). The MTO unit is modeled
mathematically using an atom balance and a typical composition
seen in the literature (Tabak & Yurchak, 1990). The MTO product is
fractionated (MTO-F) to separate the light gases, olens, and gasoline fractions. The MTO-F unit is a rigorous distillation column that
is designed so that approximately 100% of the C1 C3 parafns are
recycled back to the renery, 100% of the C4 parafns and 100% of
the olens are directed to the MOGD unit, 100% of the gasoline is
combined with the remainder of the gasoline generated in the process, and 100% of the water generated in the MTO unit is sent for
wastewater treatment. Note that the MTO-F unit is modeled within
the process synthesis model as a separator unit with the appropriate utilities (i.e., low-pressure steam and cooling water) that are
extracted from simulation of the distillation column.
The separated olens are sent to the MOGD unit where a xed
bed reactor is used to convert the olens to gasoline and distillate over a ZSM-5 catalyst. The gasoline/distillate product ratios
can range from 0.12 to >100, and the ratio chosen in this study was
0.12 to maximize the production of diesel. The MOGD unit operates at 400 C and 1 bar and will utilize steam generation to remove
the exothermic heat of reaction within the unit. The MOGD unit is
modeled with an atom balance and will produce 82% distillate, 15%
gasoline, and 3% light gases (Tabak & Yurchak, 1990). The product
will be fractionated (MTODF) to remove diesel and kerosene cuts
from the gasoline and light gases. The operational ratio of kerosene
to total distillate reported in the literature for the MOGD process
is about 30%, though this number may be increased by tailoring
the operating conditions within the MTO and MOGD units to yield
the appropriate range of hydrocarbons. The MTODF unit will be
modeled as a separator unit where 100% of the C11 C13 species
are directed to the kerosene cut and 100% of the C14+ species are
directed to the diesel cut.
2.5. LPG-gasoline separation
The gasoline range hydrocarbons produced by the FT-ZSM5 unit,
the MTG unit, or the MOGD process must be sent to the LPGgasoline separation owsheet depicted in Fig. 5. Each hydrocarbon
stream is split (SPFTZSM , SPMTGHC , and SPMTODHC , respectively) and
sent to a hydrocarbon knockout unit (35 C, 10 bar) for light gas
removal via vaporliquid equilibrium. The rst knock-out unit
(HCKO1) will not incorporate additional CO2 separation, so the CO2
rich light gases recovered from HCKO1 will be recycled back to the
process (SPLG ). The second knock-out unit (HCKO2) will separate
out CO2 from the recovered light gases via a 1-stage Rectisol unit
(CO2 SEP) for sequestration or recycle back to additional process
units (MXCO2C ). The CO2 lean light gases will be recycled back to
the process.
The crude liquid hydrocarbons recovered from the two knockout units is sent to a deethanizer (DEETH) to remove any C1 -C2
hydrocarbons. The light HC gases are sent to an absorber column
(ABS-COL) where a lean oil recycle is used to strip the C3+ HCs
from the input. The liquid bottoms from the ABS-COL are then
reuxed back to the deethanizer. The C3+ HCs from the bottom of

R.C. Baliban et al. / Computers and Chemical Engineering 47 (2012) 2956

the deethanizer are sent to a stabilizer column (STA-COL) where


the C3 /C4 hydrocarbons are removed and alkylated (ALK-UN) to
produce iso-octane and an LPG byproduct. Additional iso-butane
(INBUT ) may be fed to the alkylation unit for increased alkylate production. The bottoms from the stabilizer column is sent to a splitter
column (SP-COL) to recover a lean oil recycle from the column top
for use in the absorber column. Light and heavy gasoline fractions
are recovered from the column top and bottom, respectively. The
LPG/alkylate from the alkylation unit is split (LPG-ALK) into an LPG
byproduct (OUTLPG ) and an alkylate fraction which is blended with
the gasoline fractions from the splitter column (OUTGAS ). Each of
the distillation units is modeled mathematically as a splitter unit
where the split fraction of each species to an output stream is
given by the information in the Process Flow Diagrams P850-A1501
and P850-A1502 from the NREL study (National Renewable Energy
Laboratory, 2011). All low pressure steam and cooling water needed
for each of the units is derived for each of the units in the NREL
study. The total amount of process utility that is needed per unit
ow rate from the top or bottom of the column is calculated, and
this ratio is used as a parameter in the process synthesis model to
determine the actual amount of each utility needed based on the
unit ow rate. The alkylate was modeled as iso-butane (National
Renewable Energy Laboratory, 2011) and the alkylation unit was
modeled using a species balance where the key species, butene,
was completely converted to iso-butane. Butene is used as the limiting species in this reaction because it is generally present in a far
smaller concentration than iso-butane.
2.6. Unit costs
The total direct costs, TDC, for the CBGTL renery hydrocarbon production and upgrading units are calculated using estimates
from several literature sources (Mobil Research & Development
Corporation, 1978, 1983, 1985; National Energy Technology
Laboratory, 2007; National Renewable Energy Laboratory, 2011)
using the cost parameters in Table 3 and Eq. (4)
TDC = (1 + BOP) Co

S
So

sf

(4)

where Co is the installed unit cost, So is the base capacity, Sr is


the actual capacity, sf is the cost scaling factor, and BOP is the
balance of plant (BOP) percentage (site preparation, utility plants,
etc.). The BOP is estimated to be 20% of the total installed unit cost.
All numbers are converted to 2009 dollars using the GDP ination
index (US Government Printing Ofce, 2009). Detailed cost estimates were not available for the MTO or OGD process units, so
the cost associated with these units was estimated from the cost
of an atmospheric MTG unit provided by Mobil (Mobil Research &
Development Corporation, 1978). Note that not all units in Figs. 15
are represented in Table 3. Some of the units shown in Table 3
represent the cost of that unit plus any auxillary units needed for
proper unit operation. Specically, (a) the three FT aqueous phase
knock-out units are included in the cost of the hydrocarbon recovery column (Bechtel, 1998), (b) the cost of the FT ZSM-5 fractionator
is included in the cost of the FT ZSM-5 unit (Mobil Research &
Development Corporation, 1983, 1985), (c) the MTO fractionator
is included in the cost of the MTO unit (National Renewable Energy
Laboratory, 2011), and (d) the OGD fractionator was included in the
cost of the OGD unit (Mobil Research & Development Corporation,
1978).
The total overnight capital, TOC, for each unit is calculated as the
sum of the total direct capital, TDC, plus the indirect costs, IC. The
IC include engineering, startup, spares, royalties, and contingencies
and is estimated to 32% of the TDC. The TOC for each unit must be
converted to a levelized cost to compare with the variable feedstock
and operational costs for the process. Using the methodology of

37

Kreutz et al. (2008), the capital charges (CC) for the renery are calculated by multiplying the levelized capital charge rate (LCCR) and
the interest during construction factor (IDCF) by the total overnight
capital (Eq. (5)).
CC = LCCR IDCF TOC

(5)

Kreutz et al. (2008) calculates an LCCR value of 14.38%/year and


IDCF of 1.076. Thus, a multiplier of 15.41%/year is used to convert
the overnight capital into a capital charge rate. Assuming an operating capacity (CAP) of 330 days/year and operation/maintenance
(OM) costs equal to 5% of the TOC, the total levelized cost (CostU )
associated with a unit is given in Eq. (6).
Cost U
u =

 LCCR IDCF
CAP

OM
365

 

TOC u
Prod LHV Prod

(6)

The levelized costs for the units described for hydrocarbon production and upgrading are added to the complete list of CBGTL
process units given in Baliban, Elia, and Floudas (2012).
2.7. Objective function
The objective function for the model is given in Eq. (7). The
summation represents the total cost of liquid fuels production and
includes contributions from the feedstocks cost (CostF ), the electricity cost (CostEl ), the CO2 sequestration cost (CostSeq ), and the
levelized unit investment cost (CostU ). Each of the terms in Eq. (7)
is normalized to the total lower heating value in GJ of products
produced. For each case study, the capacity and ratio of liquid fuel
products is xed, so the normalization denominator in Eq. (7) will
be a constant parameter. Note that other objective functions (e.g.,
maximizing the net present value) can be easily incorporated into
the model framework.
MIN

 

Cost Fs + Cost El + Cost Seq +

uUIn (u,s)S U

Cost U
u

(7)

uUInv

The process synthesis model with simultaneous heat, power,


and water integration represents a large-scale non-convex mixedinteger non-linear optimization (MINLP) model that was solved to
global optimality using a branch-and-bound global optimization
framework that was previously described (Baliban, Elia, Misener,
et al., 2012). The MINLP model contains 32 binary variables, 11,104
continuous variables, 10,103 constraints, and 351 non-convex
terms (i.e., 285 bilinear terms, 1 trilinear term, 1 quadrilinear term,
and 64 power functions). At each node in the branch-and-bound
tree, a mixed-integer linear relaxation of the mathematical model
is solved using CPLEX (CPLEX, 2009) and then the node is branched
to create two children nodes. The solution pool feature of CPLEX
is utilized during the solution of the relaxed model to generate a
set of distinct points (150 for the root node and 10 for all other
nodes), each of which is used as a candidate starting point to solve
the original model. For each starting point, the current binary variable values are xed and the resulting NLP is minimized using
CONOPT. If the solution to the NLP is less than the current upper
bound, then the upper bound is replaced with the NLP solution
value. At each step, all nodes that have a lower bound that is within
an  tolerance of the current upper bound ((LBnode /UB) 1 ) are
eliminated from the tree. For a more complete coverage of branchand-bound algorithms, the reader is directed to the textbooks of
Floudas (Floudas, 1995, 2000) and reviews of global optimization
methods (Floudas, Akrotirianakis, Caratzoulas, Meyer, & Kallrath,
2005; Floudas & Gounaris, 2009; Floudas & Pardalos, 1995).
3. Computational studies
The proposed process synthesis model was used to analyze twenty-four distinct case studies using perennial biomass

38

R.C. Baliban et al. / Computers and Chemical Engineering 47 (2012) 2956

Table 3
CBGTL renery upgrading unit reference capacities, costs (2009$), and scaling factors.
Description

Co (MM$)

So

SMax

Units

Scale basis

sf

Ref.

FischerTropsch unit
Hydrocarbon recovery column
Distillate hydrotreater
Kerosene hydrotreater
Naphtha hydrotreater
Wax hydrocracker
Naphtha reformer
C5 C6 isomerizer
C4 isomerizer
C3 C5 alkylation unit
Saturated gas plant
FT ZSM-5 reactor
Methanol synthesis
Methanol degasser
Methanol-to-gasoline unit
Methanol-to-olens unit
Olens-to-gasoline/diesel unit
CO2 separation unit
Deethanizer
Absorber column
Stabilizer column
Splitter column
HF alkylation unit
LPG/alkylate splitter

$12.26
$0.65
$2.25
$2.25
$0.68
$8.42
$4.70
$0.86
$9.50
$52.29
$7.83
$4.93
$8.22
$3.82
$5.80
$3.48
$3.48
$5.39
$0.58
$0.91
$1.03
$1.01
$8.99
$1.06

23.79
1.82
0.36
0.36
0.26
1.13
0.43
0.15
6.21
12.64
4.23
10.60
35.647
11.169
10.60
10.60
10.60
8.54
5.13
0.96
4.57
3.96
0.61
0.61

60.0
25.20
81.90
81.90
81.90
72.45
94.50
31.50

kg/s
kg/s
kg/s
kg/s
kg/s
kg/s
kg/s
kg/s
kg/s
kg/s
kg/s
kg/s
kg/s
kg/s
kg/s
kg/s
kg/s
kg/s
kg/s
kg/s
kg/s
kg/s
kg/s
kg/s

Feed
Feed
Feed
Feed
Feed
Feed
Feed
Feed
Feed
Feed
Feed
Feed
Feed
Feed
Feed
Feed
Feed
Feed
Feed
Feed
Feed
Feed
Feed
Feed

0.72
0.70
0.60
0.60
0.65
0.55
0.60
0.62
0.60
0.60
0.60
0.65
0.65
0.70
0.65
0.65
0.65
0.62
0.68
0.68
0.68
0.68
0.65
0.68

b,c

a
b
c
d
e

d
d
d
d
d
d
d
d
d
d
b,c
e
e
a,e
a
a
a
a,e
a,e
a,e
a,e
a,e
a,e

Mobil Research and Development Corporation (1978).


Mobil Research and Development Corporation (1983).
Mobil Research and Development Corporation (1985).
Bechtel Corporation (1998).
National Renewable Energy Laboratory (2011).

(switchgrass), low-volatile bituminous coal (Illinois #6), and natural gas as feedstocks. A global optimization framework was used
for each case study, and termination was reached if all nodes in
the branch-and-bound tree have been processed or if 100 CPU
hours have passed (Baliban, Elia, Misener, et al., 2012). The ultimate and proximate analysis of the biomass and coal feedstocks and
the molar composition of the natural gas feedstock are presented
as Supplementary Information. To examine the effects of potential
economies of scale on the nal liquid fuels price, three distinct plant
capacities were examined to represent a small, medium, or large
capacity hybrid energy plant. Based on current petroleum renery
capacities (Energy Information Administration, 2009), representative sizes of 10 thousand barrels per day (TBD), 50 TBD, and 200 TBD
were chosen, respectively. A minimal carbon conversion threshold
of 40% was enforced for all of the case studies, and no upper bound
was used for the amount of CO2 that is vented or sequestered. This
threshold value was imposed to provide a comparative baseline
between all of the case studies, and does not have an effect on the
overall process topologies. If no lower threshold value is imposed,
then the overall conversion for each study will range between 34%
and 39%, which is consistent with the results of a previous study
(Baliban, Elia, Misener, et al., 2012). In general, raising the conversion rate produce more liquid fuels and decrease the byproduct
electricity output from the plant, and for a more in-depth analysis,
the reader is directed to the previous study (Baliban, Elia, Misener,
et al., 2012). The overall greenhouse gas emission target for each
case study is set to have a 50% reduction from petroleum based
processes (Baliban, Elia, & Floudas, 2012; Baliban et al., 2011). The
current case studies do not include the cost of a carbon tax for any
GHG emissions, though the process synthesis framework could be
readily extended include a cost for the total lifecycle emissions.
Four superstructure combinations will be investigated to analyze the effect of plant topology on the nal liquid fuels cost. These
superstructures will consider (1) only FischerTropsch synthesis
with fractionation of the vapor efuent, (2) only FischerTropsch
synthesis with ZSM-5 catalytic upgrading of the vapor efuent, (3)
only methanol synthesis with either the MTG or MOGD process,

and (4) a comprehensive superstructure allowing all possibilities


from (1), (2), or (3). Note that in superstructures (1), (2), and (4),
any wax efuent from the FischerTropsch units will be converted
to naphtha and diesel via a wax hydrocracker. Two sets of liquid
fuels products (i.e., gasoline/diesel/kerosene and gasoline/diesel)
will be considered to determine the effect of these products on the
optimal plant topology and overall costs. The ratio of liquid fuel
production will be equal to the total 2010 United States demand
(Energy Information Administration, 2011). Note that the process
superstructure is also capable of analyzing a variable concentration of output fuels (e.g., max diesel). Each of the 24 case studies
discussed below has a label PCN where P is the type of products
produced (GDK gasoline/diesel/kerosene, GD gasoline/diesel),
C is the plant capacity (S small, M medium, L large), and N is
the superstructure number dened above.
The cost parameters (Baliban, Elia, & Floudas, 2012; Baliban
et al., 2011) used for CBGTL process are listed in Table 4. The costs for
feedstocks (i.e., coal, biomass, natural gas, freshwater, and butanes)
include all costs associated with delivery to the plant gate. The
products (i.e., electricity and propane) are assumed to be sold from
the plant gate, and do not include the costs expected for transport
to the end consumer. The cost of CO2 capture and compression will
be included in the investment cost of the CBGTL renery while the
cost for transportation, storage, and monitoring of the CO2 is shown
in Table 4.
Once the global optimization algorithm has completed, the
resulting process topology provides (i) the operating conditions
and working uid ow rates of the heat engines, (ii) the amount
of electricity produced by the engines, (iii) the amount of cooling water needed for the engines, and (v) the location of the pinch
points denoting the distinct subnetworks. Given this information,
the minimum number of heat exchanger matches necessary to
meet specications (i), (ii), (iii), and (iv) are calculated as previously described (Baliban, Elia, & Floudas, 2012; Baliban et al.,
2011; Floudas, 1995; Floudas, Ciric, & Grossmann, 1986). Upon
solution of the minimum matches model, the heat exchanger topology with the minimum annualized cost can be found using the

R.C. Baliban et al. / Computers and Chemical Engineering 47 (2012) 2956

39

Table 4
Cost parameters (2009$) for the CBGTL renery.
Item

Cost

Item

Cost

Coal (LV bituminous)


Natural gas
Butanes
Electricity

$93.41/short ton ($3/GJ)


$5.39/TSCF1 ($5.5/GJ)
$1.84/gallon
$0.07/kWh

Biomass (Switchgrass)
Freshwater
Propanes
CO2 TS&M2

$139.97/dry metric ton ($8/GJ)


$0.50/metric ton
$1.78/gallon
$10/metric ton

1
2

TSCF thousand standard cubic feet


TS&M transportation, storage, and monitoring

superstructure methodology (Elia et al., 2010; Floudas, 1995;


Floudas et al., 1986). The investment cost of the heat exchangers is
added to the investment cost calculated within the process synthesis model to obtain the nal investment cost for the superstructure.
3.1. Optimal process topologies
The information detailing the optimal process topologies
for all case studies is shown in Table 5. Three possible temperature options were used for the biomass gasier (900 C,
1000 C, 1100 C), the coal gasier (1100 C, 1200 C, 1300 C),
the auto-thermal reactor (700 C, 800 C, 950 C), and the reverse
watergas-shift unit (400 C, 500 C, 600 C). For all 24 case studies, the biomass and coal solid/vapor fueled gasiers were utilized
in the optimal process design. Thus, each gasier employed a
vapor phase recycle stream as a fuel input along with the solid
coal or biomass. Recycle of some of the unreacted synthesis gas
to the gasiers helped to consume some CO2 generated in the
process and reduce the overall process emissions by converting the CO2 to CO for additional liquid fuels production. For the
biomass gasier, the 900 C unit is always selected for superstructure 1 and only selected for superstructure 3 at high capacity
levels. For all other case studies, the 1100 C unit is selected.
For the coal gasier, the 1300 C unit was always selected for
superstructures 1, 2, and 3 and the 1100 C unit was selected for
superstructure 4.
Selection of the gasier operating temperatures in the optimal
topology represents a balance between (i) the levels of oxidant
input to the gasier, (ii) the extent of consumption of CO2 via the
reverse watergas-shift reaction, and (iii) the level of waste heat
generated from syngas cooling. Lower gasier temperatures will
have less favorable conditions for CO2 consumption due to lower
values of the watergas-shift equilibrium constant and a smaller
amount of waste heat for use in steam generation and ultimately
electricity production. However, lower temperatures will require
lower levels of O2 for combustion within the gasier which reduces
the investment and utility cost for oxygen generation and may
increase the overall efciency of the gasier. The alternative disadvantages with a higher O2 in the higher temperature gasiers are
balanced by an increase in the CO2 reduction potential and the additional waste-heat generated. The operating temperature selected
in the 24 case studies reects the trade-offs between emissions
reduction, electricity production, and overall process efciency for
the entire renery.
The auto-thermal reformer temperature was selected to be
950 C for twelve of the case studies and 800 C for the remaining twelve studies (see Table 5). A 950 C unit is always used for
superstructure 1, used for superstructure 2 in the medium and large
plants, and used in superstructure 4 for the large plants. Selection of the temperature for the auto-thermal reformer will have
similar topological effects as the gasiers, though the overall conversion of CH4 will also increase with increasing temperature. The
use of the highest temperature reformer is benecial since approximately 90% of the input CH4 can be converted to syngas using a
H2 O/CH4 ratio of approximately 1.21.5. Ultimately, this will also
decrease the working capacity of the FT synthesis or methanol

synthesis units because the input CH4 is an inert species that will
not be separated until downstream of these units. The selection
of the 800 C units for the remaining studies generally converts
8285% of the CH4 , though the decrease in the oxygen requirement to the unit provides an economic benet to the decreased
conversion of the natural gas.
A dedicated reverse watergas-shift unit was not selected for
either product composition and plant capacity that used superstructures 1, 2, or 3. For each of these case studies, the proper syngas
ratio requirements for the FT and methanol synthesis was met via
light gas recycle to either the gasiers or the auto-thermal reactor
units. For the case studies using superstructure 4, a 600 C reverse
watergas-shift unit was utilized to both consume CO2 generated
in the process and shift the syngas ratios for conversion. All of the
case studies generated H2 using pressure-swing absorption and O2
using air separation. The H2 was utilized mostly for product upgrading and for injection, with the balance being sent to the reverse
watergas-shift units to consume some CO2 . Note that H2 separation is required for hydrotreating and hydrocracking within the
product upgrading section. Electrolyzers were not utilized in any
case study due to the high capital ($500/kW) and electricity costs
of the unit. The electricity input to the electrolyzers is assumed
to come from a non-carbon based source (e.g., wind/solar), which
was assumed to have a high cost (i.e., $0.10/kWh). Note that input
electricity from a carbon-based source (i.e., biomass/coal/natural
gas) is not considered because the process superstructure accounts
for H2 generation from pressure-swing absorption. A decrease in
the non-carbon based electricity cost may have an effect on the
electrolyzer use, as noted in a previous study (Baliban et al., 2011).
Both a gas and steam turbine are used in each case study to produce electricity for the process and to partially sell as a byproduct.
To reduce the GHG emissions from the processes, each case study
utilized CO2 capture and sequestration both upstream of synthesis
gas conversion and downstream of the gas turbine engine.
The case studies using superstructures 1 and 2 required FT
synthesis of the hydrocarbons, and each case study utilized an
iron-based catalyst within both the minimal-wax and nominalwax reactors. Additionally, the reverse watergas-shift reaction
was facilitated in most of the case studies, with the exception of
the minimal-wax reactor in superstructure 2 for the medium and
large capacities. In the former case studies, the iron-based units
can take advantage of the exothermic FT reaction to provide heat
for the endothermic reverse watergas-shift reaction (Baliban, Elia,
& Floudas, 2012; Baliban et al., 2011). In the latter studies, the
additional CO2 that is generated from the FT reactors is captured
and recycled back to the process to minimize the GHG emissions.
Due to the constraints of the process superstructure, upgrading
of the vapor phase FT efuent utilized a fractionation scheme for
superstructure 1 and the ZSM-5 catalyst for superstructure 2. For
superstructure 3, the syngas was converted to methanol rather
than hydrocarbons via the FT reaction. For all case studies using
this superstructure, both the methanol-to-gasoline and methanolto-olens/distillate processes are utilized to produce the liquid
fuels in the appropriate output ratios. In the case studies using
superstructure 4, the technologies used for liquid fuels production
are highly dependent on the plant capacity and the type of fuels

40

R.C. Baliban et al. / Computers and Chemical Engineering 47 (2012) 2956

Table 5
Topological information for the optimal solutions for the 24 case studies. Specically listed is the operating temperature of the biomass gasier (BGS), the coal gasier (CGS),
the auto-thermal reactor (ATR), and the reverse watergas-shift unit (RGS). The gasiers are also labeled as either solid/vapor (S/V) or solid (S) fueled, implying the presence
or absence of vapor-phase recycle process streams. The presence of a CO2 sequestration system (CO2SEQ) or a gas turbine (GT) is noted using yes (Y) or no (-). The minimum
wax and maximum wax FischerTropsch units are designated as either cobalt-based or iron-based units. The iron-based units will either facilitate the forward (fWGS)
or reverse watergas-shift (rWGS) reaction. The FT vapor efuent will be upgraded using fractionation into distillate and naphtha (Fract.) or ZSM-5 catalytic conversion.
The use of methanol-to-gasoline (MTG) and methanol-to-olens/olens-to-gasoline-and-diesel (MTO/MOGD) is noted using yes (Y) or no (-). The results for the complete
superstructure and medium sized capacity (M4) are shown in boldface.
Case study

GDK-S1

GDK-S2

GDK-S3

GDK-S4

GDK-M1

GDK-M2

GDK-M3

GDK-M4

GDK-L1

GDK-L2

GDK-L3

GDK-L4

BGS Temp. ( C)
BGS Type
CGS Temp. ( C)
CGS Type
RGS Temp. ( C)
ATR Temp. ( C)
Min Wax FT
Nom. Wax FT
FT Upgrading
MTG Usage
MOGD Usage
CO2SEQ Usage
GT Usage

900
S/V
1300
S/V

950
Ir. rWGS
Ir. rWGS
Fract.

Y
Y

1100
S/V
1300
S/V

800
Ir. rWGS
Ir. rWGS
ZSM-5

Y
Y

1100
S/V
1300
S/V

800

Y
Y
Y
Y

1100
S/V
1100
S/V
600
800

Ir. rWGS
Fract.
Y
Y
Y
Y

900
S/V
1300
S/V

950
Ir. rWGS
Ir. rWGS
Fract.

Y
Y

1100
S/V
1300
S/V

950
Ir. fWGS
Ir. rWGS
ZSM-5

Y
Y

1100
S/V
1300
S/V

800

Y
Y
Y
Y

1100
S/V
1100
S/V
600
800

Ir. rWGS
Fract.
Y
Y
Y
Y

900
S/V
1300
S/V

950
Ir. rWGS
Ir. rWGS
Fract.

Y
Y

1100
S/V
1300
S/V

950
Ir. fWGS
Ir. rWGS
ZSM-5

Y
Y

900
S/V
1300
S/V

800

Y
Y
Y
Y

1100
S/V
1100
S/V
600
950

Y
Y
Y
Y

Case study

GD-S1

GD-S2

GD-S3

GD-S4

GD-M1

GD-M2

GD-M3

GD-M4

GD-L1

GD-L2

GD-L3

GD-L4

BGS Temp. ( C)
BGS type
CGS Temp. ( C)
CGS type
RGS Temp. ( C)
ATR Temp. ( C)
Min wax FT
Nom. wax FT
FT upgrading
MTG usage
MOGD usage
CO2SEQ usage
GT usage

900
S/V
1300
S/V

950
Ir. rWGS
Ir. rWGS
Fract.

Y
Y

1100
S/V
1300
S/V

800
Ir. rWGS
Ir. rWGS
ZSM-5

Y
Y

1100
S/V
1300
S/V

800

Y
Y
Y
Y

1100
S/V
1100
S/V
600
800

Ir. rWGS
Fract.
Y

Y
Y

900
S/V
1300
S/V

950
Ir. rWGS
Ir. rWGS
Fract.

Y
Y

1100
S/V
1300
S/V

950
Ir. rWGS
Ir. rWGS
ZSM-5

Y
Y

1100
S/V
1300
S/V

800

Y
Y
Y
Y

1100
S/V
1100
S/V
600
800

Ir. rWGS
Fract.
Y

Y
Y

900
S/V
1300
S/V

950
Ir. rWGS
Ir. rWGS
Fract.

Y
Y

1100
S/V
1300
S/V

950
Ir. rWGS
Ir. rWGS
ZSM-5

Y
Y

900
S/V
1300
S/V

800

Y
Y
Y
Y

1100
S/V
1100
S/V
600
950

Ir. rWGS
ZSM-5
Y

Y
Y

produced. For the six studies with superstructure 4, the minimalwax FT unit was never utilized and the methanol-to-gasoline
process was always utilized. The nominal-wax iron-based rWGS
FT unit was used for the two small plants, the two medium plants,
and the large plant that does not produce kerosene. For the ve
case studies that used FT, the vapor phase was always converted
to gasoline-range hydrocarbons using ZSM-5. The MOGD process
was used to generate diesel and kerosene for all plant sizes in the
GDK case studies. In the GD case studies, the MOGD process was
not utilized and all diesel was generated from wax hydrocracking.
The results for the complete superstructure and medium sized
capacity (M4) are shown in boldface in Table 5. For each of these
cases, both the biomass and coal gasiers were solid/vapor fueled
units operating at 1100 C. A dedicated reverse watergas-shift unit
operating at 600 C is used and the auto-thermal reactor operates
at 800 C for both studies. The liquid fuels are produced via (i) catalytic ZSM-5 upgrading of the iron-based rWGS FT efuent, (ii) wax
hydrocracking, and (iii) methanol-to-gasoline for both studies and
by MOGD for the study requiring kerosene production.
3.2. Overall costs of liquid fuels
The overall cost of liquid fuel production (in $/GJ) is based on the
costs of feedstocks, capital investment, operation and maintenance
(O&M), and CO2 sequestration and can be partially defrayed using
byproduct sales of LPG and electricity. Feedstock costs are based
on the as-delivered price for (i) the three major carbon feedstocks
(coal, biomass, and natural gas), (ii) butanes needed for the isomerization process (Baliban et al., 2010, 2011; Bechtel, 1992), and
(iii) freshwater needed to make-up for process losses(Baliban et al.,
2012a). Table 6 outlines the breakdown of the cost contribution for

each case study, as well as the lower bound and the optimality gap
values. The total cost is also converted into a break-even oil price
(BEOP) in $/barrel based on the reners margin for gasoline, diesel,
or kerosene (Baliban et al., 2011; Kreutz et al., 2008), and represents the price of crude oil at which the CBGTL process becomes
economically competitive with petroleum based processes.
The overall cost values range between $17.33 and $18.79/GJ
for a small plant, $16.06$17.66/GJ for a medium plant, and
$14.76$16.20/GJ for a large plant. For a medium sized plant producing gasoline, diesel, and kerosene, the optimization model for
the complete superstructure (i.e., case study GDK-M4) selects a
topology with an overall cost of $16.25/GJ or $79.83/bbl crude oil
equivalent. The upper bound value found at the termination of
the global optimization algorithm is 4.56% above the lower bound
value of $15.51/GJ. When only gasoline and diesel are produced
in the general medium sized plant (GD-M4), the overall cost of
liquid fuel production for a medium sized plant with the most
general superstructure is $16.06/GJ or $78.74/bbl crude oil equivalent with a 5.35% optimality gap from its lower bound value of
$15.20/GJ. Negative values in the cost contributions from electricity and propane represent the prot gained from selling these
items as byproducts. In all of the 24 case studies, the selected
plant topologies are net producers of electricity and propane
(see Table 6).
For a given capacity level, Table 6 shows that the lowest overall
cost is achieved through the use of the most general superstructure topology. Additionally, the second lowest cost is consistently
found with superstructure 3, suggesting that the methanol synthesis/conversion process units generally yield a plant design with a
lower overall cost. However, the decrease in cost between superstructure 3 (only methanol) and superstructure 4 (methanol/FT)

R.C. Baliban et al. / Computers and Chemical Engineering 47 (2012) 2956

41

Table 6
Overall cost results for the 24 case studies. The case studies where the plant topologies produce gasoline, diesel, and kerosene are labeled as GDK, and the topologies that
produce gasoline and diesel are labeled as GD. The small (S), medium (M), and large (L) case studies are each labeled with the superstructure number, where (1) indicates that
only FischerTropsch synthesis with fractionation of the vapor efuent is considered, (2) only FischerTropsch synthesis with ZSM-5 catalytic upgrading of the vapor efuent,
(3) only methanol synthesis with either the MTG or MOGD process, and (4) a comprehensive superstructure allowing all possibilities from (1), (2), or (3). The contribution
to the total costs (in $/GJ) come from coal, biomass, natural gas, butanes, water, CO2 sequestration (CO2 . Seq.), and the investment. Propane is always sold as a byproduct
while electricity may be sold as a byproduct (negative value). The overall costs are reported in ($/GJ) and ($/bbl) basis, along with the lower bound values in ($/GJ) and the
optimality gap between the reported solution and the lower bound. The results for the complete superstructure and medium sized capacity (M4) are shown in boldface.
Case study

Contribution to cost
($/GJ of products)

Coal
Biomass
Natural gas
Butane
Water
CO2 Seq.
Investment
O&M
Electricity
Propane
Total ($/GJ)
Total ($/bbl)
Lower bound ($/GJ)
Gap

GDK-S1

GDK-S2

GDK-S3

GDK-S4

GDK-M1

GDK-M2

GDK-M3

GDK-M4

GDK-L1

GDK-L2

GDK-L3

GDK-L4

3.15
2.71
3.58
0.28
0.03
0.51
11.15
3.27
5.69
0.19
18.79
94.32
17.73
5.63%

3.18
2.69
3.48
0.31
0.02
0.50
10.81
3.17
5.43
0.15
18.59
93.18
17.86
3.92%

3.20
2.70
3.43
0.40
0.03
0.51
10.22
3.00
5.26
0.20
18.02
89.90
17.31
3.92%

2.91
2.74
4.14
0.28
0.02
0.50
10.03
2.94
5.96
0.15
17.46
86.72
16.92
3.10%

3.32
2.75
3.08
0.29
0.02
0.50
8.29
2.43
2.86
0.17
17.66
87.85
16.68
5.52%

3.31
2.81
3.02
0.25
0.02
0.49
8.16
2.40
2.82
0.14
17.51
87.00
16.54
5.55%

3.05
2.73
3.80
0.34
0.02
0.51
7.50
2.20
3.34
0.20
16.61
81.85
15.83
4.67%

3.16
2.70
3.53
0.33
0.02
0.50
7.65
2.25
3.72
0.17
16.25
79.83
15.51
4.56%

3.21
2.69
3.40
0.36
0.02
0.51
7.25
2.13
3.20
0.17
16.20
79.52
15.35
5.24%

3.08
2.73
3.78
0.25
0.02
0.51
7.44
2.18
3.92
0.21
15.86
77.58
15.32
3.40%

3.33
2.69
3.14
0.34
0.02
0.51
6.64
1.95
3.02
0.22
15.37
74.84
14.74
4.16%

3.37
2.67
3.02
0.36
0.02
0.51
6.70
1.97
3.48
0.19
14.95
72.40
14.40
3.67%

Contribution to cost
($/GJ of products)

Coal
Biomass
Natural gas
Butane
Water
CO2 Seq.
Investment
O&M
Electricity
Propane
Total ($/GJ)
Total ($/bbl)
Lower bound ($/GJ)
Gap

Case study

GD-S1

GD-S2

GD-S3

GD-S4

GD-M1

GD-M2

GD-M3

GD-M4

GD-L1

GD-L2

GD-L3

GD-L4

2.71
2.75
4.62
0.26
0.03
0.50
11.28
3.31
6.53
0.16
18.59
93.16
17.77
4.39%

3.38
2.65
2.98
0.26
0.03
0.50
11.09
3.26
5.72
0.20
18.06
90.14
17.28
4.30%

2.75
2.74
4.51
0.31
0.02
0.50
10.22
3.00
5.91
0.16
17.99
89.73
17.17
4.55%

2.72
2.75
4.59
0.32
0.02
0.50
10.07
2.96
6.43
0.16
17.33
85.99
16.41
5.32%

3.25
2.68
3.30
0.30
0.02
0.50
8.11
2.38
3.01
0.21
17.33
85.99
16.52
4.67%

3.13
2.68
3.56
0.38
0.03
0.50
8.38
2.46
3.65
0.17
17.30
85.82
16.40
5.20%

3.34
2.66
3.07
0.33
0.02
0.50
7.33
2.15
2.84
0.20
16.36
80.46
15.86
3.07%

3.23
2.66
3.30
0.36
0.02
0.50
7.48
2.20
3.56
0.14
16.06
78.74
15.20
5.35%

3.19
2.68
3.42
0.33
0.02
0.50
6.85
2.01
2.83
0.18
16.01
78.45
15.10
5.67%

3.39
2.65
2.95
0.27
0.02
0.50
6.99
2.05
2.81
0.20
15.84
77.47
14.99
5.34%

3.27
2.66
3.24
0.29
0.03
0.50
6.33
1.86
2.73
0.14
15.30
74.42
14.79
3.33%

3.27
2.65
3.21
0.33
0.03
0.50
6.52
1.91
3.53
0.14
14.76
71.31
13.97
5.33%

implies that there is a degree of synergy that can be achieved


through the use of both technologies. The resulting level of synergy
is likely to be tied to the capacity of the plant and the composition
of liquid fuels that will be produced. The CBGTL case studies using
superstructures 2 (FT with ZSM-5 upgrading) have a lower cost
ultimately due to a decrease in the complexity of the FT synthesis and upgrading section of the plant. In some case studies (i.e.,
GDK-L2, GD-L2, and GD-M2), the investment cost of the plant with
ZSM-5 upgrading was higher than that for the corresponding case
study without ZSM-5 upgrading. The increase in investment is due
to a higher overall ow rate of syngas through the renery due
to (i) increased recycle ow of the unreacted syngas to decrease
feedstock costs or (2) increased ow of the feedstocks to produce
additional byproduct electricity.
3.3. Parametric analysis
Table 6 indicates that the largest contribution to the overall
fuels cost is associated with the capital investment (i.e., capital
charges and operation/maintenance). A reduction in total plant
cost may be achieved through innovation of novel technologies
rather than relying on economies of scale for more mature processes (Adams & Barton, 2011). However, the coal, biomass, and
natural gas may have a wide variability in the overall cost of liquid fuel production. Depending on the demand for these materials

and the plant location throughout the country, the feedstock costs
may be higher or lower than the national average. Given the delivered feedstock costs in Table 4 and the feedstock lower heating
values in Supplementary Table S1, the cost per unit energy is calculated for coal ($3.0/GJ), biomass ($8.0/GJ), and natural gas ($5.5/GJ).
These cost parameters represent conservative estimates (Energy
Information Administration, 2011; Kreutz et al., 2008; Larson et al.,
2009; National Academy of Sciences, 2009) for the total delivered
cost of a particular feedstock, and it is important to investigate how
the BEOP will be affected if these cost parameters are reduced. As
an illustrative example, the BEOP for case study GDK-M4 is calculated assuming either low, nominal, or high cost values for each of
the three feedstocks. These respective values are (i) $2/GJ, $2.5/GJ,
and $3/GJ for coal, (ii) $5/GJ, $6.5/GJ, and $8/GJ for biomass, and (iii)
$4/GJ, $4.75/GJ, and $5.5/GJ for natural gas. The BEOP was calculated
for each of the 27 parameter combinations, and the histogram of
results is shown in Fig. 6.
Each cost bin in Fig. 6 represents a $2/barrel window for the
BEOP. That is, the rst bin represents all of the parameter combinations that had a BEOP between $60/bbl and $62/bbl, the second
bin is between $62/bbl and $64/bbl, and so on. The histogram
shows a Gaussian-like distribution with two major peaks in the
$68/bbl$72/bbl range with a total of 13 counts. The shape of the
histogram can be inferred from Table 6 since the contribution of
each feedstock to the overall cost is relatively similar. The singular

42

R.C. Baliban et al. / Computers and Chemical Engineering 47 (2012) 2956

yield lower total investment costs than their GDK counterparts


due to the less complicated rening that is needed to produce
kerosene.

3.5. Material and energy balances

Counts

5
4
3
2
1
0
60

64

68
72
BEOP Bin ($/bbl)

76

Fig. 6. Parametric analysis of feedstock cost. The histogram shows the number of
counts (out of 27) for break-even oil price (BEOP) when low, nominal, and high
values are used for the costs of coal, biomass, and natural gas.

peak in the leftmost bin corresponds to a BEOP of $62.7/bbl and is


obtained if the low parameters are used for each feed. The highest BEOP is equal to $80.0/bbl, and is obtained if all of the high
parameter values are used.
3.4. Investment costs
The plant investment cost is further decomposed into cost contributions from different sections of the plant in Table 7, namely
the syngas generation, syngas cleaning, hydrocarbon production,
hydrocarbon upgrading, hydrogen/oxygen production, heat and
power integration, and wastewater treatment sections. The syngas
generation section is consistently the highest contributing factor in
the investment cost due to the capital intensive coal and biomass
gasier units. The next highest contributing factors are the syngas
cleaning, hydrogen/oxygen production, and heat and power integration sections, followed by the hydrocarbon production section,
and nally the hydrocarbon upgrading and wastewater treatment
sections.
The total investment cost ranges from $1166 to $1296
MM for small plants producing gasoline, diesel, and kerosene,
$4359$4823 MM for medium plants, and $15,446$17,309 MM
for large plants. The normalized investment costs, however, reveal
the economies of scale obtained in larger sized plants, ranging from
$116k to $130k/bpd for small plants, $87k$96k/bpd for medium
plants, and $78k$87k/bpd for large plants. Among the small plant
case studies, the case with the most general superstructure (i.e.,
GDK-S4) is able to achieve the lowest investment cost. For larger
sized plants, however, GDK-M3 and GDK-L3 case studies have the
lowest investment costs for medium and large plants case studies,
respectively. Conversely, the case studies using superstructure 1
from all capacity levels have the highest total investment cost.
Comparisons between the GDK and GD case studies reveal interesting trade-offs in investment costs. For the small plants case
studies, plant topologies that produce only gasoline and diesel
result in higher investment costs than the ones that produce gasoline, diesel, and kerosene. The increased cost of the small GD case
studies is due to a higher ow rate of syngas throughout the process units due to a slightly higher level of recycle than the GDK
small case studies. The increased investment costs for the small GD
studies do lead into smaller levels of feedstock usage than the small
GDK studies, and therefore have a lower overall cost of liquid fuels
production (see Table 6). For the medium and large GD case studies,
the topologies that produce gasoline and diesel fuels consistently

The overall material and energy balances for the 24 case studies are shown in Tables 8 and 9, respectively. The biomass and
coal ow rates are based of dry tons (dt) while the natural gas is
shown in million standard cubic feet (mscf). From Tables 8 and 9,
it can be seen that coal provides the most energy input to the
plant, followed generally by natural gas, and then biomass. For
example, the most small capacity plant with the most general
superstructure (GDK-S4) requires 69.56 dt/h coal, 51.08 dt/h for
biomass, and 1.83 mscf/h natural gas. These values correspond
to 596 MW energy input from coal, 224 MW from biomass, and
497 MW from natural gas. This distribution remains relatively consistent when the plant size increases. For the medium sized plant
(case study GDK-M4), 377.39 dt/h is needed for coal, 251.95 dt/h
for biomass, and 7.77 mscf/h, corresponding to 3234 MW energy
input for coal, 1106 MW for biomass, and 2114 MW for natural gas.
Case study GDK-L4 requires 1607.23 dt/h coal, 997.60 dt/h biomass,
and 26.64 mscf/h natural gas, corresponding to 13,775 MW energy
input from coal, 4377 MW from biomass, and 7250 MW from natural gas. The smaller contribution of biomass relative to the other
two feedstocks is due to the higher $/GJ costs associated with
biomass. The highest driving force for the use of biomass is the
lifecycle GHG reduction potential, but the use of CO2 sequestration from the 24 case studies (see Table 8) will reduce the biomass
requirement for the plant. A restriction on the amount of CO2 that
is captured for sequestration (e.g., no nearby available locations
for CO2 storage) will ultimately increase the biomass feedstock
requirement, and the biomass could become the largest energy
contributor to the renery. The authors note that the biomass
requirement for the large case studies (i.e., 200,000/bpd) is necessary to achieve a life-cycle GHG emissions that is 50% lower
than petroleum-based processes. Though the biomass-based plant
designs by the National Renewable Energy Laboratory use approximately 2000 dry tons/day (National Renewable Energy Laboratory,
2011; Spath et al., 2005), the availability of biomass may be substantially higher in several counties (e.g., Midwestern United States)
after land-use change or an increase in crop yields (Department of
Energy, 2005).
Almost all of the case studies do not vent CO2 from the process,
and utilize CO2 sequestration to reduce the lifecycle GHG emissions
of the plant. The GDK-M1 and GDK-M2 studies vent a small amount
of CO2 , though the CO2 is only 12% of the total CO2 produced by
the plant. The balance of the CO2 is captured for sequestration. The
high utilization of CO2 sequestration allows for an increased use of
the cheaper fossil fuels coal and natural gas, which can be anywhere
from $3/GJ to $6/GJ less expensive than biomass. The biomass does
provide negative emission values from CO2 intake from the atmosphere during cultivation and additional soil storage from land use
change, so a level of biomass input on a mass/energy basis that is
roughly equivalent to that of coal or natural gas is still required.
The electricity production ranges from 179 to 221 MW for small
plants, 478631 MW for medium plants, and 18502661 MW for
large plants. In all case studies, a high amount of electricity is produced to help lower the overall cost of fuels for the plant. The
electricity output also improves the efciency of the topologies,
with GD-S2, GDK-S1, and GD-S1 achieving the highest energy efciencies (i.e., 67.1%, 65.8%, and 65.7%, respectively) compared to
other case studies in their subcategories (see Table 9). The energy
efciency values are calculated by dividing the total energy output (i.e., fuel products, propane, or electricity) by the total energy
input (i.e., via coal, biomass, natural gas, butane, or electricity). If

R.C. Baliban et al. / Computers and Chemical Engineering 47 (2012) 2956

43

Table 7
Breakdown of the investment costs for the 24 case studies. The major sections of the plant include the syngas generation section, syngas cleaning, hydrocarbon production,
hydrocarbon upgrading, hydrogen/oxygen production, heat and power integration, and wastewater treatment blocks. The values are reported in MM$ and normalized with
the amount of fuels produced ($/bpd). The results for the complete superstructure and medium sized capacity (M4) are shown in boldface.
Contribution to cost (MM$)

Case study
GDK-S1

Syngas generation
Syngas cleaning
Hydrocarbon production
Hydrocarbon upgrading
Hydrogen/oxygen production
Heat and power integration
Wastewater treatment
Total (MM $)
Total ($/bpd)

494
240
218
24
145
146
29
1296
129,647

Contribution to cost (MM$)

Case study

GDK-S2

GD-S1
Syngas generation
Syngas cleaning
Hydrocarbon production
Hydrocarbon upgrading
Hydrogen/oxygen production
Heat and power integration
Wastewater treatment
Total (MM $)
Total ($/bpd)

GDK-S3

GDK-S4

GDK-M1

GDK-M2

GDK-M3

GDK-M4

GDK-L1

GDK-L2

GDK-L3

GDK-L4

5362
3153
2682
343
2424
2521
377
16,862
87,211

5702
3186
2775
326
2495
2412
412
17,309
86,547

5063
2932
1985
207
2588
2364
307
15,446
79,335

5187
2870
2148
206
2451
2405
305
15,572
77,858

492
234
208
22
138
137
26
1258
125,754

476
222
170
16
139
138
28
1188
118,809

478
225
166
16
126
129
26
1166
116,609

1422
813
738
165
789
781
115
4823
96,451

1443
786
731
147
770
742
127
4745
94,897

1314
769
566
96
768
747
99
4359
87,177

1369
758
603
100
754
767
98
4450
88,993

GD-S2

GD-S3

GD-S4

GD-M1

GD-M2

GD-M3

GD-M4

1373
785
735
154
784
767
120
4717
94,335

1480
805
785
150
756
773
123
4872
97,434

1270
779
554
90
742
733
94
4261
85,226

1318
764
594
86
734
759
94
4348
86,958

500
244
218
25
148
146
30
1311
131,118

511
242
209
22
139
140
27
1290
128,975

486
219
161
16
134
143
30
1189
118,893

483
223
166
15
131
126
27
1171
117,083

electricity is output from the system, the value is listed as negative in Table 8 and the magnitude of the energy value in Table 9 is
added to the total output. If the value is positive in Table 8, then this
energy is added to the total input to the system. The overall energy
efciency of the CBGTL topologies producing gasoline, diesel, and
kerosene ranges between 58.5 and 67.1% for all plant sizes.

GD-L1
5068
3071
2399
338
2365
2326
367
15,935
79,677

GD-L2

GD-L3

5183
3035
2602
327
2413
2304
401
16,265
81,326

4875
2770
1867
190
2471
2237
314
14,723
73,615

GD-L4
4868
2919
2190
205
2292
2381
298
15,153
75,764

3.6. Carbon and greenhouse gas balances


The overall carbon balance for the CBGTL processes is shown
in Table 10 and highlights the eight major points where carbon is
either input or output from the system. The results for the complete
superstructure and medium sized capacity (M4) case studies are

Table 8
Overall material balance for the 24 case studies. The inputs to the CBGTL process are biomass, coal, natural gas, butane, and water, while the outputs include gasoline, diesel,
kerosene, LPG, sequestered and vented CO2 , and electricity. Biomass and coal are input in dry metric tons per hour (dt/h), natural gas in million standard cubic feet per hour
(mscf/h), liquids in thousand barrels per day (kBD), and CO2 in metric tons per hour (tonne/h). The results for the complete superstructure and medium sized capacity (M4)
are shown in boldface.
Material balances

Case study
GDK-S1

GDK-S2

GDK-S3

GDK-S4

GDK-M1

GDK-M2

GDK-M3

GDK-M4

GDK-L1

GDK-L2

GDK-L3

GDK-14

Biomass (dt/h)
Coal (dt/h)
Natural gas (mscf/h)
Butane (kBD)
Water (kBD)
Gasoline (kBD)
Diesel (kBD)
Kerosene (kBD)
LPG (kBD)
Seq. CO2 (tonne/h)
Vented CO2 (tonne/h)
Electricity (MW)

50.56
75.16
1.58
0.21
18.18
6.72
2.15
1.13
0.14
240.04
0.00
193.14

50.28
75.95
1.53
0.23
14.34
6.72
2.15
1.13
0.11
239.65
0.00
184.16

50.42
76.31
1.51
0.30
18.85
6.72
2.15
1.13
0.15
240.14
0.00
178.57

51.08
69.56
1.83
0.21
16.01
6.72
2.15
1.13
0.11
239.61
0.00
202.29

256.34
396.13
6.78
1.07
80.05
33.60
10.77
5.63
0.63
1183.81
15.23
484.88

262.08
395.36
6.66
0.92
77.55
33.60
10.77
5.63
0.51
1167.69
29.69
477.54

254.39
364.37
8.38
1.26
75.88
33.60
10.77
5.63
0.74
1200.70
0.00
567.06

251.95
377.39
7.77
1.21
68.38
33.60
10.77
5.63
0.62
1198.94
0.00
631.42

1005.02
1532.36
29.98
5.28
296.85
134.39
43.10
22.51
2.53
4796.55
0.00
2171.37

1018.08
1468.83
33.28
3.71
333.54
134.39
43.10
22.51
3.14
4805.53
0.00
2661.25

1002.86
1588.67
27.65
5.00
306.86
134.39
43.10
22.51
3.25
4807.09
0.00
2045.52

997.60
1607.23
26.64
5.38
313.53
134.39
43.10
22.51
2.85
4801.23
0.00
2357.99

Material balances

Case study
GD-S1

GD-S2

GD-S3

GD-S4

GD-M1

GD-M2

GD-M3

GD-M4

GD-L1

GD-L2

GD-L3

GD-L4

51.33
64.64
2.04
0.19
20.51
7.57
2.43
0.12
238.72
0.00
221.38

49.47
80.78
1.31
0.20
19.18
7.57
2.43
0.15
239.15
0.00
193.96

51.21
65.60
1.99
0.23
12.51
7.57
2.43
0.12
238.76
0.00
200.46

51.32
64.81
2.02
0.24
14.84
7.57
2.43
0.12
238.80
0.00
218.06

249.75
387.91
7.27
1.11
79.22
37.86
12.14
0.79
1196.62
0.00
509.61

250.28
373.28
7.84
1.42
92.56
37.86
12.14
0.61
1194.09
0.00
619.46

248.04
398.69
6.75
1.24
68.38
37.86
12.14
0.75
1196.04
0.00
481.53

248.07
385.57
7.27
1.35
79.22
37.86
12.14
0.51
1192.62
0.00
602.94

998.77
1523.11
30.17
4.90
286.85
151.44
48.56
2.61
4778.68
0.00
1916.43

988.91
1619.62
26.03
4.05
350.22
151.44
48.56
2.90
4782.98
0.00
1908.04

991.27
1558.90
28.55
4.33
366.90
151.44
48.56
2.13
4771.66
0.00
1849.71

990.18
1562.09
28.28
4.95
360.23
151.44
48.56
2.08
4770.88
0.00
2393.96

Biomass (dt/h)
Coal (dt/h)
Natural gas (mscf/h)
Butane (kBD)
Water (kBD)
Gasoline (kBD)
Diesel (kBD)
LPG (kBD)
Seq. CO2 (tonne/h)
Vented CO2 (tonne/h)
Electricity (MW)

44

R.C. Baliban et al. / Computers and Chemical Engineering 47 (2012) 2956

Table 9
Overall energy balance for the 24 case studies. The energy inputs to the CBGTL process come from biomass, coal, natural gas, and butane, and the energy outputs are gasoline,
diesel, kerosene, LPG, and electricity. The energy efciency of the process is calculated by dividing the total energy output with the total energy inputs to the process.
Energy balances (MW)

Biomass
Coal
Natural gas
Butane
Gasoline
Diesel
Kerosene
LPG
Electricity
Efciency (%)

Energy balances (MW)

Biomass
Coal
Natural gas
Butane
Gasoline
Diesel
LPG
Electricity
Efciency (%)

Case study
GDK-S1

GDK-S2

GDK-S3

GDK-S4

222
644
429
13
428
153
78
9
193
65.8%

221
651
417
14
428
153
78
7
184
65.3%

221
654
411
18
428
153
78
9
179
64.9%

224
596
497
13
428
153
78
7
202
65.3%

GDK-M1
1125
3395
1845
65
2141
766
390
38
485
59.4%

GDK-M2
1150
3388
1812
56
2141
766
390
31
478
59.4%

GDK-M3
1116
3123
2279
77
2141
766
390
45
567
59.3

GDK-M4

GDK-L1

GDK-L2

GDK-L3

GDK-L4

1106
3234
2114
74
2141
766
390
38
631
60.7%

4410
13,133
8157
321
8563
3065
1558
154
2171
59.6%

4467
12,589
9057
226
8563
3065
1558
191
2661
60.9%

4401
13,616
7522
304
8563
3065
1558
197
2046
59.7%

4377
13,775
7250
327
8563
3065
1558
173
2358
61.1%

Case study
GD-S1

GD-S2

GD-S3

GD-S4

GD-M1

GD-M2

225
554
554
12
482
173
7
221
65.7%

217
692
357
12
482
173
9
194
67.1%

225
562
541
14
482
173
7
200
64.3%

225
555
550
14
482
173
7
218
65.5%

1096
3325
1979
67
2412
863
48
510
59.3%

1098
3199
2133
86
2412
863
37
619
60.3%

highlighted in the table using boldface. Carbon that is input to the


system via air is neglected due to the low ow rate relative to the
other eight points. Over 99% of the input carbon is supplied from the
coal, biomass, and natural gas while the balance is supplied by the
butane input to the isomerization and alkylation units. The trends
seen in feedstock use from 8 are consistently displayed in the input
carbon ow rates in Table 10. That is, for all of the case studies,
a majority of the carbon is input from coal and CO2 sequestration
is highly utilized to reduce the GHG emissions. The biomass and
natural gas provide roughly equivalent amounts of input carbon
to the reneries, which combined represent approximately 40% of

GD-M3
1088
3417
1837
75
2412
863
45
482
59.3

GD-M4

GD-L1

GD-L2

GD-L3

GD-L4

1089
3305
1979
82
2412
863
31
603
60.6%

4383
13,054
8210
298
9649
3454
159
1916
58.5%

4339
13,881
7082
246
9649
3454
176
1908
59.4%

4350
13,361
7768
263
9649
3454
129
1850
58.6%

4345
13,388
7694
301
9649
3454
126
2394
60.7%

the input carbon. The output amount of carbon in the total product
is constant for each plant capacity, which is consistent with the
constant production capacity that is required for each feedstockconversion rate. The amount of carbon leaving as LPG is around 1%
of that leaving as gasoline, kerosene, and diesel. For all of the case
studies, most of the CO2 generated from the process is captured
and sequestered, with little or no CO2 venting.
For each of the case studies, the carbon conversion rate was set
as a lower bound (i.e., 40%) for the mathematical model. Thus, the
conversion of carbon in the four feedstocks to any of the four liquid
products must be at least as large as the set conversion rate. All of

Table 10
Carbon balances (in kg/s) for the optimal solutions for the 24 case studies. Carbon is input to the process via coal, biomass, natural gas, or butanes and exits the process as
liquid product, LPG byproduct, vented CO2 , or sequestered (Seq.) CO2 . The small amount of CO2 input to the system in the puried oxygen stream (<0.01%) is neglected. The
results for the complete superstructure and medium sized capacity (M4) are shown in boldface.
Case study

GDK-S1

GDK-S2

GDK-S3

GDK-S4

GDK-M1

GDK-M2

GDK-M3

GDK-M4

GDK-L1

GDK-L2

GDK-L3

GDK-L4

Biomass
Coal
Natural gas
Butane
Gasoline
Diesel
Kerosene
LPG
Vented CO2
Seq. CO2
% Conversion

5.90
16.91
7.32
0.19
7.78
2.85
1.40
0.10
0.00
18.20
40.0%

5.87
17.09
7.12
0.21
7.78
2.85
1.40
0.08
0.00
18.17
40.0%

5.88
17.17
7.02
0.27
7.78
2.85
1.40
0.11
0.00
18.20
40.0%

5.96
15.65
8.47
0.19
7.78
2.85
1.40
0.08
0.00
18.16
40.0%

29.91
89.13
31.47
0.98
38.91
14.24
6.98
0.46
1.15
89.74
40.0%

30.58
88.96
30.91
0.84
38.91
14.24
6.98
0.38
2.25
88.51
40.0%

29.68
81.98
38.88
1.15
38.91
14.24
6.98
0.55
0.00
91.02
40.0

29.39
84.91
36.06
1.11
38.91
14.24
6.98
0.46
0.00
90.88
40.0%

117.25
344.78
139.13
4.83
155.64
56.95
27.93
1.87
0.00
363.60
40.0%

118.78
330.49
154.47
3.39
155.64
56.95
27.93
2.33
0.00
364.28
40.0%

117.00
357.45
128.30
4.57
155.64
56.95
27.93
2.41
0.00
364.39
40.0%

116.39
361.63
123.65
4.92
155.64
56.95
27.93
2.11
0.00
363.95
40.0%

Case study

GD-S1

GD-S2

GD-S3

GD-S4

GD-M1

GD-M2

GD-M3

GD-M4

GD-L1

GD-L2

GD-L3

GD-L4

Biomass
Coal
Natural gas
Butane
Gasoline
Diesel
LPG
Vented CO2
Seq. CO2
% Conversion

5.99
14.54
9.45
0.18
8.77
3.21
0.09
0.00
18.10
40.0%

5.77
18.18
6.09
0.18
8.77
3.21
0.11
0.00
18.13
40.0%

5.97
14.76
9.22
0.21
8.77
3.21
0.09
0.00
18.10
40.0%

5.99
14.58
9.38
0.22
8.77
3.21
0.09
0.00
18.10
40.0%

29.14
87.28
33.75
1.01
43.85
16.04
0.58
0.00
90.71
40.0%

29.20
83.99
36.38
1.29
43.85
16.04
0.45
0.00
90.52
40.0%

28.94
89.70
31.33
1.13
43.85
16.04
0.55
0.00
90.66
40.0

28.94
86.75
33.75
1.23
43.85
16.04
0.38
0.00
90.40
40.0%

116.52
342.70
140.03
4.48
175.39
64.18
1.93
0.00
362.24
40.0%

115.37
364.41
120.79
3.70
175.39
64.18
2.15
0.00
362.57
40.0%

115.65
350.75
132.49
3.96
175.39
64.18
1.58
0.00
361.71
40.0%

115.52
351.47
131.23
4.52
175.39
64.18
1.54
0.00
361.65
40.0%

R.C. Baliban et al. / Computers and Chemical Engineering 47 (2012) 2956

45

Table 11
Greenhouse gas (GHG) balances for the optimal solutions for the 24 case studies. The total GHG emissions (in CO2 equivalents kg CO2 eq/s) for feedstock acquisition and
transportation, product transportation and use, CO2 sequestration, and process venting are shown for each study. Process feedstocks include biomass, coal, natural gas, and
butane while products include gasoline, diesel, kerosene, and LPG. The results for the complete superstructure and medium sized capacity (M4) are shown in boldface.
Case study

GDK-S1

GDK-S2

GDK-S3

GDK-S4

GDK-M1

GDK-M2

GDK-M3

GDK-M4

GDK-L1

GDK-L2

GDK-L3

GDK-L4

Biomass
Coal
Natural gas
Butane
Gasoline
Diesel
Kerosene
LPG
Vented CO2
Seq. CO2
Total GHG (kg/bbl)

27.76
2.11
3.42
0.02
30.73
11.17
5.49
0.43
0.00
3.33
250.00

27.60
2.13
3.32
0.03
30.73
11.17
5.49
0.35
0.00
3.33
250.00

27.68
2.14
3.27
0.03
30.73
11.17
5.49
0.45
0.00
3.34
250.00

28.04
1.95
3.95
0.02
30.73
11.17
5.49
0.34
0.00
3.33
250.00

140.73
11.10
14.67
0.12
153.64
55.86
27.46
1.89
4.23
16.44
250.00

143.88
11.08
14.41
0.10
153.64
55.86
27.46
1.55
8.25
16.22
250.00

139.66
10.21
18.13
0.14
153.64
55.86
27.46
2.23
0.00
16.68
250.00

138.32
10.58
16.81
0.13
153.64
55.86
27.46
1.87
0.00
16.65
250.00

551.76
42.94
64.87
0.58
614.54
223.45
109.83
7.63
0.00
66.62
250.00

558.93
41.16
72.02
0.41
614.54
223.45
109.83
9.48
0.00
66.74
250.00

550.58
44.52
59.82
0.55
614.54
223.45
109.83
9.80
0.00
66.77
250.00

547.68
45.04
57.65
0.60
614.54
223.45
109.83
8.60
0.00
66.68
250.00

Case study

GD-S1

GD-S2

GD-S3

GD-S4

GD-M1

GD-M2

GD-M3

GD-M4

GD-L1

GD-L2

GD-L3

GD-L4

Biomass
Coal
Natural gas
Butane
Gasoline
Diesel
LPG
Vented CO2
Seq. CO2
Total GHG (kg/bbl)

28.18
1.81
4.41
0.02
34.62
12.59
0.35
0.00
3.32
250.00

27.16
2.26
2.84
0.02
34.62
12.59
0.44
0.00
3.32
250.00

28.11
1.84
4.30
0.03
34.62
12.59
0.36
0.00
3.32
250.00

28.18
1.82
4.37
0.03
34.62
12.59
0.37
0.00
3.32
250.00

137.11
10.87
15.74
0.12
173.12
62.95
2.37
0.00
16.62
250.00

137.40
10.46
16.96
0.16
173.12
62.95
1.85
0.00
16.58
250.00

136.17
11.17
14.61
0.14
173.12
62.95
2.25
0.00
16.61
250.00

136.19
10.80
15.74
0.15
173.12
62.95
1.55
0.00
16.56
250.00

548.33
42.68
65.29
0.54
692.49
251.79
7.87
0.00
66.37
250.00

542.91
45.38
56.32
0.45
692.49
251.79
8.76
0.00
66.43
250.00

547.68
45.04
57.65
0.55
614.54
223.45
109.83
6.27
0.00
66.68

543.61
43.77
61.19
0.55
692.49
251.79
6.27
0.00
66.26
250.00

the 24 case studies reached this bound, implying that this constraint
was active in the optimal solution. Note that this constraint can be
relaxed if a smaller conversion rate of liquid fuels is desired. Ultimately, this will have the effect of decreasing the overall fuels cost
by potentially generating additional byproduct electricity. However, recent studies have suggested that the CBGTL process designs
will tend to convert between 34% and 37% of the feedstock carbon when a lower conversion threshold of 25% is set (Baliban, Elia,
Misener, et al., 2012). Therefore, the minimum threshold of 40%
will serve to provide a baseline measure of comparison between
the case studies while not dramatically impacting the nal overall
cost.
The greenhouse gas (GHG) emission balances for the case studies are shown in Table 11. For each of the studies, the total GHG
emission target was set to be equal to 50% of the emissions from
a standard petroleum based process. For a typical emission level
of 500 kg of CO2 equivalent per barrel, this implies that the total
well-to-wheel GHG emissions for the CBGTL renery must be less
than 250 kg CO2 eq/bbl. The GHG emission rates (in kg CO2 eq/s) for
the ten major point sources in the renery are listed in Table 11
and include (a) acquisition and transportation of the biomass, coal,
natural gas, and butane feeds, (b) transportation and use of the
gasoline, diesel, kerosene, and LPG, (c) transportation and sequestration of any CO2 , and (d) venting of any process emissions. The
GHG emissions for feedstock acquisition and transportation in (a),
product transportation in (b), and CO2 transportation in (c) are
calculated from the GREET model for well-to-wheel emissions
(Argonne National Laboratory. GREET 1.8b, 2007) and assuming transportation distances for feedstocks (50 miles), products
(100 miles), and CO2 (50 miles). The GHG emissions from product use in (b) are calculated assuming that each product will be
completly combustion to generate CO2 that is simply vented to the
atmosphere.
From Table 11, it is clear that a major component of the lifecycle
emissions are attributed to the liquid fuels. In fact, over 80% of the
liquid fuel emissions result from combustion of these fuels in light
and heavy duty vehicles. The total emissions from transportation
of the feedstocks, products, and CO2 represents the balanced of the
lifecycle emissions for the process. To balance the GHG lifecycle, the

CO2 removed from the atmosphere due to storage in the biomass


or storage in the soil is included in the total emissions for biomass.
Note that while the net emissions for biomass is negative, there
will still be a positive component to the emissions for biomass harvesting and transportation. It is important to observe that though
the coal was the highest energy input to the renery, the emissions
contribution from natural gas is higher from coal or biomass.
4. Conclusions
This study has detailed the development of a framework for
the process synthesis of a thermochemical hybrid coal, biomass,
and natural gas to liquids plant that incorporates multiple possibilities for hydrocarbon production and hydrocarbon upgrading.
The framework also included a simultaneous heat, power, and
water integration to compare the costs of utility generation and
wastewater treatment in the overall cost of liquid fuels. This
work expands previous studies on the CBGTL process (Baliban,
Elia, & Floudas, 2012; Baliban et al., 2011) by directly quantifying the economic and environmental benets that are associated
with (i) FischerTropsch synthesis and subsequent hydrocarbon
upgrading and (ii) methanol synthesis, conversion to hydrocarbons, and subsequent upgrading. The proposed optimization model
was tested using 24 distinct case studies that are derived from
two combinations of products, three plant capacities, and four
superstructure possibilities. The overall conversion of carbon from
feedstock to liquid products was selected to be 40% and the greenhouse gas reduction target was equal to 50% of current petroleum
based reneries. Each case study was globally optimized using a
branch-and-bound global optimization algorithm to theoretically
guarentee that the cost associated with the optimal design was
within 36% of the best value possible.
When producing gasoline, diesel, and kerosene in ratios commensurate with Untied States demands, the overall cost of
liquid fuels production ranges from $86/bbl to $94/bbl for small
plants (10,000 barrels per day; kBD), $79/bbl$88/bbl for medium
plants (50 kBD), and $72/bbl$80/bbl for large plants (200 kBD).
When only gasoline and diesel are produced in a ratio consistent with national demand, the cost decreases for each of the

46

R.C. Baliban et al. / Computers and Chemical Engineering 47 (2012) 2956

capacities to a range of $85/bbl$93/bbl for small, $78/bbl$86/bbl


for medium, and $71/bbl$78/bbl for large plants. This decrease
in cost is generally due to the reduction in investment needed
to fractionate and convert the distillate to diesel only opposed
to both diesel and kerosene. For the four different superstructure
possibilities investigated in this study, it is evident that FT synthesis followed by fractionation (superstructure 1) and upgrading
is more expensive than FT synthesis followed by catalytic ZSM5 conversion to gasoline-range hydrocarbons (superstructure 2).
Additionally, methanol synthesis, conversion to hydrocarbons, and
subsequent upgrading (superstructure 3) is consistently cheaper
than FT synthesis for all capacity levels. This is due to the decrease
in investment cost associated with hydrocarbon production and
upgrading when compared to FT synthesis. These ndings indicate
that the methanol route is preferential to the FT route when following an either or logic. However, investigation of a combination
superstructure that considered all of the topologies (superstructure 4) in superstructures 13 indicates that a combination of FT
synthesis and methanol synthesis will provide the lowest overall
cost. In this case, the MTG route provides a majority of the gasoline
while a majority of the distillate (diesel and kerosene) is generated through fractionation and rening of the FT efuent. Though
over 80% of the nal hydrocarbons were produced via the methanol
synthesis route, the nal process topologies show that the ability
to consume CO2 in iron-based FT reactors helps to reduce feedstock
costs and therefore provide an economic advantage over a topology
that utilizes only methanol synthesis.

Indices/sets
The indices are used throughout the mathematical model are
listed below.
u : Process unit index
s : Species index
a : Atom index
p : Proximate analysis index
r : Reaction index
i : General counting index
The set, U, is dened as the complete set of process units. Several
subsets of units are then dened for specic areas of the CBGTL
process as presented below.
uBGS
uCGS
uRGS
uATR

= {u : u = BGSn }
= {u : u = CGSn }
= {u : u = RGSn }
= {u : u = ATRn }

The set of all atoms, A, includes C, H, O, N, S, Cl, Ar, and a generic


Ash atom. Typically, the biomass and coal ash will consist of multiple metal oxides, but the ash is assumed to be inert in the CBGTL
process, so the treatment of the ash as an atomic element is justied.
a A = {C, H, O, N, S, Cl, Ar, Ash}
The list of all unit connections, UC, is derived below.

Acknowledgements
UC = {(u, u ) : a connection between unit u and unit u in the superstructure}
The authors acknowledge partial nancial support from the
National Science Foundation (NSF EFRI-0937706).
Appendix A. Mathematical model for process synthesis
with simultaneous heat, power, and water integration

u , s) SUF is then constructed from all streams in UC along with the


set of all species s that exist within a given unit u (SU ).

The nomenclature for all terms in the mathematical model for


process synthesis with simultaneous heat, power, and water integration is shown below. All constraints included in the model are
listed subsequently with a corresponding description of how that
particular equation governs proper operation of the process design.
For a more extensive discussion of the mathematical model, the
reader is directed to previously published works (Baliban, Elia, &
Floudas, 2012; Baliban, Elia, Misener, et al., 2012; Baliban et al.,
2011).
A.1. Process units
The set of units, U, is presented in full detail in Table A1 and
dened formally in Eq. (A.1). Note that several units in Table A1
are listed as un . The n subscript represents the consideration of
multiple forms of the same process unit, each with a distinct set
of operating conditions (e.g., temperature and pressure). Though
these unit properties are generally given as continuous variables
in a process synthesis problem, they have been assumed to take
discrete choices and will be modeled using binary variables.
u U = {Complete set of process units listed in Table A1}

(A.1)

UC
S UF = {(u, u , s) : s Su,u
}

S U = {(s, u) : (u, u , s) S UF or (u , u, s) S UF }

Parameters
With the exception of all biomass and coal species, char, and the
pseudocomponents, the molecular formula is equal to the species
index dened in Table A2. The pseudocomponent hydrocarbons
and oxygenate formulas are given by Bechtel while the formulas for
biomass and coal compounds are derived from the ultimate analysis and normalized to one mole of carbon. Char has been assumed
to consist completely of carbon and ash has been assigned a generic
molecular weight of 1.0 g/mol. The atomic ratio (ARs,a ) of atom a in
species s is derived from the molecular formulas in Table A2.
ARs,a : Atomic ratio of atom a in species s
Using the appropriate atomic weight of atom a (AWa ), the molecular weight of all species s (MWs ) is dened using Eq. (A.3).
AW a : Atomic weight of atom a

Process species
The set of all species, S, is listed in Table A2 and dened formally
in Eq. (A.2).
s S = {Complete set of species listed in Table A2}

Using a priori knowledge about the operations of each unit in


the CBGTL process, the complete set of species that can possibly
UC . The set (u,
exist in a stream from unit u to unit u is dened as Su,u


(A.2)

MW s =

AW a ARs,a

(A.3)

The proximate analysis for the biomass and coal species s is


described by the total mass of moisture per unit mass of dry input

R.C. Baliban et al. / Computers and Chemical Engineering 47 (2012) 2956

47

Table A1
Process units present in the CBGTL synthesis problem. The subscript n corresponds to multiple forms of the same process unit, each with a distinct set of operating conditions
or ratios of feedstock. Distinct process units are used in lieu of continuous variables representing the process operating conditions. This will prevent the use of bilinear terms
when specifying feedstock ratios or highly non-linear equations when specifying equilibrium constants or species enthalpies.
Unit name

Unit index

Unit name

Unit index

Process inlets
Inlet coal
Inlet biomass
Inlet water

INCOAL
INBIO
INH 2 O

Inlet Natural gas


Inlet air
Inlet butane

INNG
INAIR
INBUT

Process outlets
Outlet gasoline
Outlet kerosene
Outlet sulfur
Outlet vent
Outlet sequestered CO2

OUTGAS
OUTKER
OUTS
OUTV
OUTCO2

Outlet diesel
Outlet ash
Outlet scrubbed HCl
Outlet propane
Outlet Wastewater

OUTDIE
OUTASH
OUTSCR
OUTPRO
OUTWW

Syngas generation
Biomass dryer
Biomass lockhopper
First biomass vapor cyclone
Tar cracker
Tar cracker cooler
Coal dryer air heater
Coal gasier
Second coal vapor cyclone
Second coal cyclone cooler

BDR
BLK
BC1
TCK
XTCK
XCDR
CGSn
CC2
XTCK

Biomass dryer air heater


Biomass Gasier
Second biomass vapor cyclone
Tar cracker splitter
Coal dryer
Coal lockhopper
First coal vapor cyclone
Second coal cyclone splitter

XBDR
BGSn
BC2
SPTCK
CDR
CLK
CC1
SPCC2

Syngas cleaning
Reverse water gas shift unit
COSHCN hydrolyzer
Acid gas ash vapor cooler
Acid gas ash unit
Acid gas removal unit
CO2 recycle compressor
Acid gas compressor

RGSn
CHH
XAGF
AGF
AGR
CO2 RC
AGC

RGS efuent cooler


HCl scrubber
Acid gas ash 2-phase cooler
Acid gas thermal analyzer
First CO2 compressor
CO2 sequestration compressor

XRGS
HSC
XAGFn
XAGR
CO2 C
CO2 SC

Claus sulfur recovery


Acid gas splitter
Claus combustor
First sulfur separator
Second sulfur converter
Third sulfur converter heater
Third sulfur separator
Tail gas hydrolyzer
Tail gas ash 2-phase cooler
Tail gas compressor

SPAG
CC
SS1
SC2
XSC3
SS3
TGH
XTGFn
TGC

Acid gas preheater


First sulfur converter
Second sulfur converter heater
Second sulfur separator
Third sulfur converter
Sulfur pit
Tail gas ash vapor cooler
Tail gas ash unit

XAG
SC1
XSC2
SS2
SC3
SPT
XTGF
TGF

Hydrocarbon production
MTFTWGS-N
FT-ZSM5
MEOHS
MEDEG
MTO
OGD
FischerTropsch compressor
Low-temperature preheater
Low-temperature iron-based FT
High-temperature preheater
High-temperature iron-based FT
Low-temperature efuent cooler
Water-soluble oxygenates separator
Primary vaporliquidwater separator

Iron MT fWGS nominal wax FT


ZSM-5 hydrocarbon conversion unit
Methanol synthesis unit
Methanol degasser
Methanol to olens ZSM-5 reactor
Olens to gasoline/distillate
FTC
XLTFT
LTFT
XHTFT
HTFT
XLTFTC
WSOS
VLWS

MTFTWGS-M
ZSM5F
MEOH-F
MTG
MTO-F
MTODF
FischerTropsch splitter
Low-temperature splitter
Low-temperature cobalt-based FT
High-temperature splitter
High-temperature cobalt-based FT
High-temperature efuent cooler
Vapor-phase oxygenates separator

Iron MT fWGS minimal wax FT


ZSM-5 product fractionation
Methanol ash unit
Methanol to gasoline ZSM-5 reactor
MTO fractionation
OGD fractionation
SPFT
SPLTFT
LTFTRGS
SPHTFT
HTFTRGS
XHTFTC
VPOS

Hydrocarbon recovery
Hydrocarbon recovery column
Distillate hydrotreater
Naphtha hydrotreater
C4 Isomerizer
C3 C4 C5 Alkylation unit
Diesel blender
HCKO1
DEETH
CO2 SEP
ALK-UN
SP-COL

HRC
DHT
NHT
C4 I
C345 A
DBL
Mixed hydrocarbon knockout 1
De-ethanizer
1-stage Rectisol CO2 separation
HF alkylation unit
Splitter column

Wax Hydrocracker
Kerosene hydrotreater
Naphtha reformer
C5 C6 Isomerizer
Saturated gas plant
Gasoline blender
HCKO2
ABS-COL
STA-COL
LPG-ALK

WHC
KHT
NRF
C56 I
SGP
GBL
Mixed hydrocarbon knockout 2
Absorber column
Stabilizer column
LPG/Alkylate splitter

Recycle gas treatment


Light gas compressor
Auto-thermal reactor
Fuel combustor
Fuel combustor ash unit

LGC
ATRn
FCM
FCF

Light gas splitter


Auto-thermal reactor splitter
Fuel combuster efuent cooler
First gas turbine air compressor

SPLG
SPATRn
XFCM
GTAC1

48

R.C. Baliban et al. / Computers and Chemical Engineering 47 (2012) 2956

Table A1 (Continued)
Unit name

Unit index

Unit name

Unit index

Second gas turbine air compressor


First gas turbine
Gas turbine efuent cooler
Gas turbine efuent compressor
Water gas shift unit

GTAC2
GT1
XGT
GTEC
WGS

Gas turbine combustor


Second gas turbine
Gas turbine ash unit
CO2 recovery unit

GTC
GT2
GTF
CO2 R

Water treatment
Sour stripper
Biological digestor
Cooling tower
Heat &Power system
Deaerator
Process water boiler

SS
BD
CLTR
HEP
DEA
XWBL

Sour gas compressor


Reverse osmosis
Process cooling
Heat &Power utilities
Process water economizer

SGC
RO
COOL-P
HEAT-P
XWPR

Hydrogen/oxygen production
PSA efuent splitter
PSA hydrogen preheater
Electrolyzer
Electrolyzer oxygen splitter
Electrolyzer hydrogen splitter
Air separation unit
ASU oxygen preheater
OC oxygen preheater

SPPSA
XH2 P
EYZ
SPO2 E
SPH2 E
ASU
XO2 A
XO2 C

Pressure-swing Absorption unit


PSA hydrogen splitter
Electrolyzer oxygen preheater
Electrolyzer hydrogen preheater
Air Compressor
Oxygen compressor
OC Oxygen splitter

PSA
SPH2 P
XO2 E
XH2 E
AC
OC
SPO2 C

D
(PAM
s ) and the dry weight fractions (PAp,s ) of the ash, xed carbon,
and volatile matter components p.

Variables

Continuous variables are used in the mathematical model to


S
PAM
describe the species molar ow rates (Nu,u
s : Mass of water per unit mass of dry species s
 ,s ), the total molar ow
PAD
p,s : Dry mass fraction of proximate analysis component p in species s rates (N T  ), the extent of reaction in a process unit ( u ), the molar
r
u,u
In this study, switchgrass was chosen for the biomass feedstock
and low-volatile bituminous coal was chosen for the coal feedstock.

S
composition of a stream (xu,u
 ,s ), the split fraction of a stream
between two units (spu,u ), the total stream enthalpy ow rate

Table A2
Species present in the CBGTL synthesis problem. The molecular formula of the pseudocomponent hydrocarbons and oxygenates are given by Bechtel. The formula for the
biomass and coal species are derived from the ultimate analysis assuming that the atomic weight of ash is 1.0 g/mol.
Species name

Species index

Species name

Species index

Species name

Species index

Acid gases
Sulfur dioxide

SO2

Hydrogen sulfur

H2 S

Carbonyl sulde

COS

Hydrogen cyanide
Carbon dioxide

HCN
CO2

Ammonia

NH3

Hydrogen chloride

HCl

Light non-hydrocarbon gases


Oxygen
Nitric oxide
Carbon monoxide

O2
NO
CO

Nitrogen
Nitrous oxide
Hydrogen

N2
N2 O
H2

Argon
Water

Ar
H2 O

Hydrocarbons
Methane
Ethane
Isobutylene
n-Butane
n-Pentane
n-Hexane
1-Octene
n-Nonane
1-Undecene
n-Dodecane
1-Tetradecene
n-Pentadecane
1-Heptadecene
n-Octadecane
1-Eicosene
C22 Pseudocomponent
C25 Pseudocomponent
C28 Pseudocomponent
VP Oxygenate

CH4
C2 H6
iC4 H8
nC4 H10
nC5 H12
nC6 H14
C8 H16
C9 H20
C11 H22
C12 H26
C14 H28
C15 H32
C17 H34
C18 H38
C20 H40
C22 OP
C25 OP
C28 OP
OXVAP

Acetylene
Propylene
1-Butene
1-Pentene
1-Hexene
1-Heptene
n-Octane
1-Decene
n-Undecane
1-Tridecene
n-Tetradecane
1-Hexadecene
n-Heptadecane
1-Nonadecene
n-Eicosane
C23 Pseudocomponent
C26 Pseudocomponent
C29 Pseudocomponent
HP Oxygenate

C2 H2
C3 H6
nC4 H8
C5 H10
C6 H12
C7 H14
C8 H18
C10 H20
C11 H24
C13 H26
C14 H30
C16 H32
C17 H36
C19 H38
C20 H42
C23 OP
C26 OP
C29 OP
OXHC

Ethylene
Propane
Isobutane
2-Methylbutane
2-Methylpentane
n-Heptane
1-Nonene
n-Decane
1-Dodecene
n-Tridecane
1-Pentadecene
n-Hexadecane
1-Octadecene
n-Nonadecane
C21 Pseudocomponent
C24 Pseudocomponent
C27 Pseudocomponent
C30+ Pseudocomponent
AP Oxygenate

C2 H4
C3 H8
iC4 H10
iC5 H12
iC6 H14
C7 H16
C9 H18
C10 H22
C12 H24
C13 H28
C15 H30
C16 H34
C18 H36
C19 H40
C21 OP
C24 OP
C27 OP
C30 Wax
OXH2O

Products
Gasoline
Solid sulfur

GAS
S

Diesel

DIE

Kerosene

KER

Non-conventional components
Biomass
Feedstock ash

e.g. Perennial
Ash

Coal

e.g. LV-bituminous

Gasier char

Char

R.C. Baliban et al. / Computers and Chemical Engineering 47 (2012) 2956


T ), the heat lost from a unit (Q L ), the heat transferred to or
(Hu,u

u

absorbed from a unit (Qu ), the delivered cost of feedstock (Cost Fs ),


the cost of CO2 sequestration (CostSeq ), the cost of electricity (CostEl ),
and the levelized unit investment cost (Cost U
u ). Note that the subscripts u and u are both used to denote an element of the set U and
can be used interchangeably in the stream ow indices.

49

Flash units
Upper bound of liquid phase split fraction
S
min{1,
xu,u
L ,s

1
VLE
Ku,s

} 0 (u, uL , s) S UF , u UFl

(A.10)

Upper bound of vapor phase split fraction


S
VLE
xu,u
min{1, Ku,s
} 0 (u, uV , s) S UF , u UFl
V ,s

S

Nu,u
 ,s : Molar ow of species s from unit u to unit u
T

Nu,u : Total molar ow from unit u to unit u
ru : Extent of reaction r in unit u
S

xu,u
 ,s : Molar composition of species s from unit u to unit u
spu,u : Split fraction of stream going from unit u to unit u
T

Hu,u
 : Total enthalpy ow from unit u to unit u
L
Qu : Heat lost from unit u
Qu : Heat transferred to or absorbed from unit u
Cost Fs : Total delivered cost of feedstock s
Cost Seq : Total sequestration cost of CO2
Cost El : Total cost of electricity
Cost U
u : Total levelized cost of unit u

(A.11)

Set liquid phase split fraction


S
T
S
xu,u
Nu,u
Nu,u
= 0 u UFl
L ,s
L
L ,s

(A.12)

Set vapor phase split fraction


S
T
S
xu,u
Nu,u
Nu,u
= 0 u UFl
V ,s
V
V ,s

(A.13)

Set phase equilibrium


S
xu,u
V ,s

VLE
S
Ku,s
xu,u
= 0 u UFl
L ,s

(A.14)

Heat balances
Conservation of energy

Binary variables (yu ) are introduced to represent the logical use


of a process unit u. These binary variables are only needed for specic process units since many of the units in the CBGTL process will
always be required. The units that require binary variables include
the biomass and coal gasiers, the reverse water gas shift unit, the
FischerTropsch units, the autothermal reactor, and the gas turbine.

T
Hu,u


(u,u )UC

HuT ,u Qu QuL Wu = 0 u

(u ,u)UC

U
(A.15)
UAgg

Total heat balance


T
Hu,u


S
Hu,u
(u, u ) UC
 ,s = 0

(A.16)

(u,u ,s)S UF

yu : Logical existence of process unit u (i.e., it takes the value


of one if unit u is selected and zero otherwise)

Logical unit existence


Bound on molar ows

Ex
NuT ,u UBN
u yu 0 u U

General constraints

(A.17)

(u ,u)UC

Mass balances
Species balances

NuS ,u,s

(u ,u)UC

Upper bound on inlet enthalpy ow

(u,r,s )RU

r,s u

r,s r

= 0 s

SuU , u


Ex
HuT ,u UBH
u ,u yu 0 (u , u) UC, u U
S
Nu,u
 ,s

Lower bound on inlet enthalpy ow

(u,u )UC

Bal
USp

LBH
u ,u
(A.4)

Atom balances
ARs,a NuS ,u,s

(A.19)


Ex
UBH
u ,u yu 0 (u, u ) UC, u U

(A.20)

Lower bound on outlet enthalpy ow

NuS ,u,s = 0 (u, r, s) RU

(A.5)

(u ,u,s)S UF

yu HuT ,u 0 (u , u) UC, u U Ex

Upper bound on outlet enthalpy ow


T
Hu,u


Extent of reaction
ru fc r

(A.18)

(u ,u,s)S UF

LBH
u,u

T
yu Hu,u
(u, u ) UC, u U Ex
 0

(A.21)

Process inlets
S
ARs,a Nu,u
 ,s

Feedstock moisture content


Set biomass moisture content from proximate analysis

(u,u ,s)S UF
Bal
= 0 a AU
u , u UAt

(A.6)

S
Mu,u
 ,H

2O

S

PAM
s Mu,u ,s = 0 (u, u ) = (INBIO , BDR)

(A.22)

sSBio

Total mole balance


NuT ,u

NuS ,u,s = 0 (u, u ) UC

Set coal moisture content from proximate analysis


(A.7)

2O

(u,u ,s)S UF

(u,u ,s)S UF

S
xu,u
(u, u ) UC Comp
 ,s 1 = 0

S

PAM
s Mu,u ,s = 0 (u, u ) = (INCOAL , CDR)

(A.23)

Known stream compositions


Set stream compositions for inlet streams
(A.8)
S
K
T
Nu,u
(u, u , s) S UF , u = {INAIR , INNG , INBUT }
 ,s xu,s Nu,u = 0

Split fractions sum to 1

sSCoal

Process splitters
Set unit split fractions
S
S
T
(u, u , s) S UF , u USp
Nu,u
 ,s xu ,u,s Nu,u = 0
I

S
Mu,u
 ,H

(A.9)

(A.24)

50

R.C. Baliban et al. / Computers and Chemical Engineering 47 (2012) 2956

Coal to natural gas ratio


Set coal to natural gas inlet ratio based on lower heating value
ratios

LHV s LHV Rat


CG

S
NIN
COAL ,CDR,s

UB (1 yu ) 0 (u, u ) = (INCOAL , CDR)

T
S
MT Bio Mu,u
 Mu,u ,H

2O

(A.36)

LHV NG

sSCoal

Lower bound for coal drier activation

T
NIN
NG ,u

=0
Upper bound for biomass drier moisture evaporation

(INNG ,u)UC

(A.25)
UB (1 yu ) 0 (u, u ) = (BDR, BLK)

T
S
MT Bio Mu,u
 Mu,u ,H

2O

(A.37)

Greenhouse gas emissions reduction


Set reduction from petroleum based processes
GHGCBGTL GHGRed GHGPet = 0

(A.26)

S
Mu,u
 ,H

Sum emissions from CBGTL components

2O

GHGCBGTL GHGSeq GHGProc GHGFeed = 0

T
MT Bio Mu,u
(u, u ) = (BDR, BLK)
 UB (1 yu ) 0

(A.38)

(A.27)

Set emissions from feedstock acquisition


GHGFeed

Lower bound for biomass drier moisture evaporation

Upper bound for coal drier moisture evaporation

S
GHGTs Mu,u
 ,s = 0

(A.28)

T
S
MT Coal Mu,u
 Mu,u ,H

2O

uUIn (u,u ,s)S UF

UB (1 yu ) 0 (u, u ) = (CDR, CLK)


(A.39)

Set emissions from CO 2 sequestration


S
GHGSeq GHGTCO2 MW CO2 NCO

2 SC,OUTCO2 ,CO2

=0

(A.29)

Set emissions from CO2 venting


S
GHGProc MW CO2 NCO

2 R,OUTV ,CO2

Lower bound for coal drier moisture evaporation


S
Mu,u
 ,H O
2

=0

T
MT Coal Mu,u
(u, u ) = (CDR, CLK)
 UB (1 yu ) 0

(A.30)

(A.40)

Process outlet fuel ratios


Gasier lockhoppers
Set CO2 lockhopper ow rate

Set gasoline to diesel output ratio


S
MW GAS NGBL,OUT

GAS

S
,GAS Rat GD MW DIE NDBL,OUT

DIE ,DIE

=0

S
MCO

2 C2 ,BLK,CO2

mf u

S
MBDR,BLK,s
=0

(A.41)

sSBio

(A.31)

Biomass gasier
Watergas-shift equilibrium

Set diesel to kerosene output ratio

Nu,BC1,CO Nu,BC1,H2 O KuRGS Nu,BC1,CO2 Nu,BC1,H2 = 0 u UBGS


S
MW DIE NDBL,OUT
DIE ,DIE

S
Rat DK MW KER NKHT,OUT
KER ,KER

=0
(A.32)

(A.42)
Hydrocarbon conversion fraction
Mu,BC1,s

HC

cf u,s MsS,Calc = 0 s SHC , u UBGS

(A.43)

(u ,u,s)S UF

Syngas generation

Hydrocarbon generation from pyrolysis


Biomass/coal driers
Upper bound for biomass drier activation
S
Mu,u
 ,H

2O

MsS,Calc

S
Pyr HC
s,s Mu ,u,s

s SBio (u ,u,s )S UF

T
MT Bio Mu,u
(u, u ) = (INBIO , BDR)
 UB yu 0

MuS ,u,s =0 u UBGS

(u ,u)UC

(A.44)
(A.33)
Set ratio of NO to N2 O

Upper bound for coal drier activation


S
Mu,u
 ,H

T

O MT Coal Mu,u UB yu 0 (u, u ) = (INCOAL , CDR)

(A.34)

Nu,BC1,NO sr u,

NO
N2 O

Nu,BC1,N2 O = 0 u UBGS

(A.45)

Set ratio of HCN to NH3


Nu,BC1,HCN sr u, HCN Nu,BC1,NH3 = 0 u UBGS

(A.46)

NH3

Lower bound for biomass drier activation


T
S
MT Bio Mu,u
 Mu,u ,H

Set amount input nitrogen to NH3 and N2


O UB (1 yu ) 0 (u, u ) = (INBIO , BDR)

Nu,BC1,NH3 + 2 Nu,BC1,N2 nf u

(A.35)

S
Nu,BC1,s
ARs,N

(u,BC1,s)S UF

= 0 u UBGS

(A.47)

R.C. Baliban et al. / Computers and Chemical Engineering 47 (2012) 2956

51

Set ratio of HCN to NH3


Set ratio of NH3 to N2

Nu,CC1 ,HCN sr HCN Nu,CC1 ,NH3 = 0 u UCGS

= 0 u UBGS

Set amount input nitrogen to NH3 and N2

(A.48)

(A.49)

Amount of char production


S
(a1u,Char + a2u,Char Tu )
MW Char Nu,BC1,Char

= 0 u UBGS

(A.62)

Nu,CC1 ,NH3 (a1u,N + a2u,N Tu ) (Nu,CC1 ,NH3 + 2 Nu,CC1 ,N2 )


2

(A.50)

= 0 u UCGS

= 0 u UBGS

(A.51)

Nu,CC1 ,H2 S sr u, COS Nu,CC1 ,COS = 0 u UCGS


Amount of char production

S
MBLK,u,s
LHV s = 0 u UBGS

(A.52)

S
MW Char Nu,CC

1 ,Char

(a1u,Char + a2u,Char Tu )

sSBio

yu 1 = 0

Rate of ash removal


S
Nu,OUT

QuL
(A.55)

(A.56)

sSCoal

Watergas-shift equilibrium
Nu,CC1 ,CO2 Nu,CC1 ,H2 = 0 u UCGS

HC

cf u,s MsS,Calc = 0 s SHC , u UCGS

(A.58)

(u ,u,s)S UF

s SCoal (u ,u,s )S UF

yu 1 = 0

(A.68)

(A.69)

Removal of solids from second cyclone


T
T
NCC2,CGS
=0
rf CC2 NCC1,CC2

(A.70)

Reverse watergas-shift unit


Bypass of inert species

(u ,u,s)S UF

MuS ,u,s =0 u UCGS

NuS ,u,s

S
Nu,u
 ,s

(u,u ,s)S UF

= 0 s SuIn , u URGS

(A.71)

(u ,u)UC

(A.59)

Nu,CC1 ,N2 O = 0 u UCGS

Watergas-shift equilibrium
Nu ,u,CO2 Nu ,u,H2
Nu ,u,CO Nu ,u,H2 O KuRGS


Set ratio of NO to N2 O
NO
N2 O

(A.67)

Coal gasier solids


Removal of solids from rst cyclone

S
Pyr HC
s,s Mu ,u,s

S
MCLK,u,s
LHV s = 0 u UCGS

Logical use of one gasier temperature

Hydrocarbon generation from pyrolysis

(A.66)

Syngas cleaning

Hydrocarbon conversion fraction

T
T
NCC1,CGS
=0
rf CC1 NCGS,CC1

(A.57)

NuS ,u,Ash = 0 u UCGS

uUCGS

S
MCDR,CLK,s
=0

Nu,CC1 ,CO Nu,CC1 ,H2 O KuRGS

+ hlu

Coal gasier
Set CO2 lockhopper ow rate

sSCoal

T
T
NBC2,BGS
=0
rf BC2 NBC1,BC2

mf u

sf u,Ash

Gasier heat loss

(A.54)

Removal of solids from second cyclone

CO2 ,CLK,CO2

ASH ,Ash

(u ,u)UC

T
T
NBC1,BGS
=0
rf BC1 NBGS,BC1

S
MSP

(A.65)

(A.53)

Biomass gasier solids


Removal of solids from rst cyclone

Nu,CC1 ,NO sr

S
MW s NCLK,u,s

= 0 u UCGS

uUBGS

MsS,Calc


sSCoal

Logical use of one gasier temperature

Mu,CC1 ,s

(A.64)

H2 S

Gasier heat loss

(A.63)

Set ratio of COS to H2 S


NuS ,u,Ash

(u ,u)UC

ARs,N

Set ratio of NH3 to N2

S
MW s NBLK,u,s

sSBio

QuL + hlu

1 ,s

= 0 u UCGS

H2 S

sf u,Ash

S
Nu,CC

(u,CC1 ,s)S UF

Nu,BC1,COS sr u, COS Nu,BC1,H2 S = 0 u UBGS

S
Nu,OUT
ASH ,Ash

Nu,CC1 ,NH3 + 2 Nu,CC1 ,N2 nf u

Set ratio of COS to H2 S

Rate of ash removal

(A.61)

NH3

Nu,BC1,NH3 (a1u,N + a2u,N Tu ) (Nu,BC1,NH3 + 2 Nu,BC1,N2 )

(A.60)

= 0 u URGS , u = XRGS

(A.72)

52

R.C. Baliban et al. / Computers and Chemical Engineering 47 (2012) 2956

Logical use of unit with at most one temperature

yu 1 0

Adjust weight fraction of C4 species


(A.73)

uURGS


1
Wn )
(1
6

W4 =

(A.85)

n=5

COSHCN hydrolyzer
Bypass of inert species

NuS ,u,s

(u ,u,s)S UF

Set weight fraction of Cn species from AndersonSchultzFlory distribution


S
Nu,u
s SuIn , u UCHH
 ,s = 0

(A.74)

(u,u ,s)S UF

Wn = n(1 )2 n1

Nu ,u,COS Nu ,u,H2 O

WWax =

Nu ,u,CO2 Nu ,u,H2 S


(A.75)

n(1 )2 n1

(A.87)

Set carbon distribution from weight fractions

(A.76)

(A.88)

n Wn + nWax WWax
n=1

Nu ,u,CO Nu ,u,NH3

= 0 (u , u) = (CHH, HSC)

n Wn

29

cr n =

HCNNH3 equilibrium
KuHCN


n=30

= 0 (u , u) = (CHH, HSC)

Nu ,u,HCN Nu ,u,H2 O

(A.86)

Set weight fraction of wax

COSH2 S equilibrium
KuCOS


5 n 29

Set exactly one low-temperature unit


yLTFT + yLTFTRGS 1 = 0

(A.89)

Set exactly one high-temperature unit


Acid gas recovery
Set CO2 molar fraction in clean output
S
NAGR,SP

AGR ,CO2

T
rf AGR NAGR,SP

CG

yHTFT + yHTFTRGS 1 = 0

=0

(A.77)

Set CO2 output ow rates


T
NAGR,CO

2C

T
sf AGR (NAGR,CO

2C

T
+ NAGR,MX

CO2RC

)=0

(A.78)

Set inlet combustor oxygen level


S
Nu,CC,O
2

Aqueous phase oxygenates separator


Removal of aqueous phase oxygenates
S
= 0 s SAPO
NWSOS,VLWS,s

(A.91)

Vapor phase oxygenates separator


Removal of vapor phase oxygenates

Claus sulfur recovery

(A.90)

er CC

S
Nu,CC,s

S
= 0 s SVPO
NVPOS,HRC,s

sor s = 0

(A.79)

(u,CC,s)S UF

(u,CC)UC

(A.92)

Hydrocarbon upgrading

Hydrocarbon production

Hydrocarbon upgrading units


Set carbon distribution fractions of total input

FischerTropsch
Set ratio of H2 to CO in cobalt-based inlet

S
Nu,u
 ,s ARs,C cf u,u ,s

FTRu,CO

(u ,u,H2 )S UF

(A.80)

FTRu,CO

(u ,u,H2 )S UF

FTRu,CO2

(u ,u,CO)S UF

= 0 u UIrFT


1
(1
Wn )
2

S
NSGP,C

(u ,u,CO2 )S UF

(A.82)

n=5

S
Nu,FCM,O
er FCM

S
NSP

LG ,FCM,s

sor s = 0

(SPLG ,FCM,s)S UF

(A.95)

(A.83)
Auto-thermal reactor
Logical use of one temperature

Adjust weight fraction of C3 species

n=5

(A.94)

Fuel combustor
Set inlet combustor oxygen level

(u,FCM)UC

n=5


1
(1
Wn )
6

S
Nu,SGP,s
= 0 s SC4

(u,SGP,s)S UF

W3 =

rf s

Recycle gas treatment

Adjust weight fraction of C2 species


1
Wn )
(1
6

4 I,s

(A.81)

W2 =

(A.93)

Saturated gas plant


Set fractional recovery of light gases

Adjust weight fraction of C1 species


W1 =

= 0 u UUG , (u, u , s) sUF

(u ,u,CO)S UF

NuS ,u,s ARs ,C

(u ,u,s )S UF

= 0 u UCoFT

Set ratio of H2 to CO and CO2 in iron-based inlet

(A.84)
uUATR

yu 1 = 0

(A.96)

R.C. Baliban et al. / Computers and Chemical Engineering 47 (2012) 2956

Watergas-shift equilibrium

Energy balance using reboiler and condensor

S
S
RGS
S
S
Nu,u
Nu,u
 ,CO Nu,u ,H Ku
 ,CO Nu,u ,H

Reb
Cond
+ QSS
QSS = 0
QSS

2O

= 0 (u, u ) UC, u UATR

(A.97)

Set energy use for reboiler and condensor

2O

(A.110)

Biological digestor
Set biogas ratio of CH4 to CO2

SR
S
S
Ku,CH
xu,u
 ,CH xu,u ,H

(A.109)

Reb
Cond
+ QSS
HRSS QSS
=0

CH4 Steam reforming equilibrium


S
S
xu,u
 ,CO xu,u ,H

53

= 0 (u, u ) UC, u UATR

(A.98)

S
S
NBD,CC,CH
cr BD NBD,CC,CO
=0
4

C2 H2 steam reforming equilibrium


S
xu,u
 ,C

2 H2

S
xu,u
 ,H

2O

SR
Ku,CH

Reverse osmosis
Set removal fraction of solids

S
S
xu,u
 ,CH xu,u ,CO

SR
Ku,C

S
NRO,SP

2 H2

= 0 (u, u ) UC, u UATR

(A.99)

C2 H4 steam reforming equilibrium

S
xu,u
 ,C

2 H4

SR
Ku,C

2 H2

SR
Ku,C

S
xu,u
 ,C

2 H2

2 H4

S
xu,u
 ,C

2 H6

2 H4

S
xu,u
(u, u ) UC, u UATR
 ,H = 0
2

(u ,u,s)S UF

S
xu,u
(u, u ) UC, u UATR
 ,H = 0
2

S
In
Nu,u
u UATR , s SATR
(A.102)
 ,s = 0

S
lkGTAC1 NIN
AIR ,GTAC1 ,s

S
byGTAC1 NIN

AIR ,GTAC1 ,s

= 0 (GTAC1 , s) S

Evap

Drift

S
NCLTR + NCLTR NCLTR,OUT

Kn
xCLTR,SP
CLTR ,s

Kn
xHEP,MX
BLR ,s

V ,H2 O

=0

(A.116)

T
NCLTR,SP

CLTR

S
NCLTR,SP

CLTR ,s

= 0 s SSol

(A.117)

S
Nu,GTC,O
=0
2

NXT

PWB ,MXBLR

NXS

PWB ,MXBLR ,s

= 0 s SSol

(A.118)

(A.105)

T
NHEP,MX

BLR

S
NHEP,MX

BLR ,s

= 0 s SSol

(A.119)

WW ,OUTV ,s

Max
xMX

T
NMX

WW ,OUTV ,s

WW ,OUTV

0 s SWW (A.120)

Hydrogen/oxygen production

(u,GTC,s)S UF

L
T
QGTC
hlGTC (HSP

LG ,GTC

Pressure-swing absorption
Set recovery fraction of H2 from inlet

HXT

GTF ,GTF

)=0

(A.106)

S
NPSA,SP

H2 P ,H2

(A.121)

Set inlet mole fraction of H2

S
Nu,SS,s

=0

(A.107)

SS ,s

= 0 (SS, SPSS , s) S UF

(u,PSA)UC

T
Nu,PSA
=0

(A.122)

(u,PSA)UC

Air separation unit


Recovery fraction of O2

Set fraction of sour species in bottoms


T
NSS,SP

S
2
Nu,PSA,H
InH
PSA
2

(u,SS)UC

SS ,s

S
Nu,PSA,H
=0

(u,PSA)UC

Sour stripper
Set recovery fraction of H2 O in bottoms
rf SS,H2 O

2
RevH
PSA

Wastewater treatment

Kn
xSS,SP

(A.115)

Outlet wastewater
Upper bound on output wastewater concentrations
S
NMX

Set heat loss in combustor

SS ,s

=0

Set known heat engine boiler output solid concentrations

= 0 (GTAC1 , s) S U

S
sor s Nu,GTC,s

(u,GTC,s)S UF

S
NSS,SP

CLTR ,CLTR,H2 O

Sum total cooling tower losses

PWB ,MXBLR ,s

Set inlet oxygen ow rate in combustor

S
NSS,SP
SS ,H2 O

(A.114)

Cooling tower drift loss

Kn
xX

(A.104)

=0

Set air bypass from rst compressor

er GTC

2O

Steam cycle
Set known process steam boiler output solid concentrations

(A.103)

1 ,GT2 ,s

(A.113)

Set known cooling tower output solid concentrations

Gas turbine
Set air leakage from rst compressor

S
NGTAC

(A.112)

=0

S
0.00085 TCLTR NCLTR,COOLP,H

Drift

(u,u ,s)S UF

S
NGTAC
1 ,OUTV ,s

= 0 s SSol

Cooling tower evaporation loss


Evap
NCLTR

S
NCLTR 0.001 NMX

(A.101)

NuS ,u,s

RO ,RO,s

2O

Bypass of inert species

S
rf RO NMX

Cooling cycle
Cooling tower ow rate from energy requirement

(A.100)

S
xu,u
 ,C

RO ,s

S
QC hr COOLP NCLTR,COOLP,H

C2 H6 Steam reforming equilibrium


SR
Ku,C
2 H4
SR
Ku,C
2 H6

(A.111)

(A.108)

S
NASU,OUT

V ,s

S
U
(1 sf ASU ) NAC,ASU,s
= 0 s SASU

(A.123)

54

R.C. Baliban et al. / Computers and Chemical Engineering 47 (2012) 2956

Process hot/cold/power utility requirements

Objective function

Set electricity needed for process units


QPEl

Base

Su Elu

=0

Levelized cost of fuel production


(A.124)

MIN

Set cooling water needed for process units

Su CW Base
=0
u

Set heating fuel needed for process units

Su FuBase = 0

(A.126)

uUUtil

Cost U
u

(A.137)

uUInv

Pinch points
Set pinch points based on inlet temperatures

HP-in
Tpi = Tu,u
(u, u ) HP; Tpi = Tu u HPt HB ;


T = T (ut, pi) HPt PI Ut ; T = T PCin (b, c, t) HEP;

pi
b,c,t
pi ut
Tpi = T CPin
+ T


Set utilities needed for process units


HU
Base
Qu,ut
Su Uu,ut
= 0 ut, u UUtil

Simultaneous heat and power integration


(A.125)

uUUtil

QFCM

Cost Fs + Cost El + Cost Seq +

uUIn (u,s)S U

uUUtil

QPCW

 

(A.127)

u,u

SHin

Tpi = Tb,t
+ T

(u, u ) CP;

ECin
Tpi = Tb,c
+ T

(b, t) CP ;
SH

Tpi = Tut + T

Tpi = Tb + T

Tpi = Tc

(b, c) CP EC ;

(ut, pi) CPt PI Ut ;

(A.138)
Process costs
Feedstock costs
Levelized cost of biomass feedstock
Cost Fs =

S
MW s NIN

BIO ,BDR,s

CsF

Prod LHV Prod

Temperature differences
Process unit hot stream inlets

s SBio

(A.128)

S
MW s NIN

COAL ,CDR,s

CsF

Prod LHV Prod

s SCoal

HP-out
HP-out
Tu,u
Tpi }
 ,pi = max{0, Tu,u

(A.129)

Cost Fs =

(INNG ,u)UC

CsF

Prod LHV Prod

s SNG

(A.130)

Cost FH2 O =

S
MW H2 O NIN

H2 O ,SPWRI ,H2 O

Cost

(A.142)

Heat engine precooler inlets

F
CH

2O

Prod LHV Prod

(A.131)

PCin
PCin
Tb,c,t,pi
= max{0, Tb,c,t
Tpi }

(A.143)

Heat engine precooler outlets


PCout
PCout
= max{0, Tb,c,t
Tpi }
Tb,c,t,pi

Electricity costs
Levelized cost of electricity
El

(A.141)

Process unit cold stream outlets


CPout
CPout
Tu,u
(Tpi T )}
 ,pi = max{0, Tu,u

Levelized cost of freshwater feedstock

(A.140)

Process unit cold stream inlets


CPin
CPin
Tu,u
(Tpi T )}
 ,pi = max{0, Tu,u

Levelized cost of natural gas feedstock


S
MW s NIN
NG ,u,s

(A.139)

Process unit hot stream outlets

Levelized cost of coal feedstock


Cost Fs =

HP-in
HP-in
Tpi }
Tu,u
 ,pi = max{0, Tu,u

(A.144)

Heat engine economizer inlets

El C El F El C El
FIn
In
Out
Out

(A.132)

Prod LHV Prod

ECin
ECin
Tb,c,pi
= max{0, Tb,c
(Tpi T )}

(A.145)

Heat engine economizer outlets


CO2 sequestration costs
Levelized cost of CO2 sequestration
Cost Seq =

S
MW CO2 NCO

2 SC,OUTCO2 ,CO2

ECout
ECout
Tb,c,pi
= max{0, Tb,c
(Tpi T )}

C Seq

Heat engine superheater inlets


(A.133)

Prod LHV Prod

SHin
SHin
Tb,t,pi
= max{0, Tb,t
(Tpi T )}

SHout
SHout
Tb,t,pi
= max{0, Tb,t
(Tpi T )}

sf u

(A.134)

Variable capital costs of process units


(A.135)

CC u (1 + OM)
CAP Prod LHV Prod

En
En
Fb,c,t yb,c,t
Fb,c,t
(b, c, t) HEP

(A.149)

Bound on total amount of heat engines

Levelized cost of process units


Cost U
u =

(A.148)

Heat engine logical existence


Bound on heat engine ow rate
Up

CC u = LCCR IDCF TOC u

(A.147)

Heat engine superheater outlets

Levelized investment costs


Total overnight cost of process units
Su
TOC u = (1 + IC u ) (1 + BOP u ) Co,u
So,u

(A.146)


(A.136)

(b,c,t)HEP

En
yb,c,t
EnMax

(A.150)

R.C. Baliban et al. / Computers and Chemical Engineering 47 (2012) 2956

Heat balances
Heat engine electricity balance

55

Appendix B. Supplementary data

Tur
Pum
En
(wb,c,t
wb,c,t
) Fb,c,t
= FEl

(A.151)

(b,c,t)HEP

Supplementary data associated with this article can be


found, in the online version, at http://dx.doi.org/10.1016/
j.compchemeng.2012.06.032.

Upper heat balance for pinch points

 

H
Qpi
=

(u,u )HP

P
s
HP-in
HP-out
Nu,u
 ,s Cpu,u ,s (Tu,u ,pi Tu,u ,pi )

PCin
PCout
En
Fb,c,t
CpHEP (Tb,c,t,pi
Tb,c,t,pi
)

(b,c,t)HEP

(c,pi)HPtPI C

HU
Qu,ut
+

(ut,pi)HPtPI Ut (u,ut)HPt

References

Qu

(u,pi)HPtPI HB
C

En
Fb,c,t
dH c

(A.152)

Lower heat balance for pinch points

 

C
=
Qpi

(u,u )CP

P
CPout
CPin
s
Nu,u
 ,s Cpu,u ,s (Tu,u ,pi Tu,u ,pi )

s
ECout
ECin
En
Fb,c,t
CpHEE (Tb,c,pi
Tb,c,pi
)

(b,c,t)HEP

SHout
SHin
En
Fb,c,t
CpHES (Tb,t,pi
Tb,t,pi
)

(b,c,t)HEP

HU
Qu,ut
+

(ut,pi)CPtPI Ut (u,ut)CPt



(b,pi)CPtPI B

En
Fb,c,t
dH b

(A.153)
Pinch point heating decit
C
H
Qpi
zpi = Qpi

(A.154)

Negativity of pinch decits


zpi 0

(A.155)

Total heating decit


 Qc = 0

(A.156)

Total heat balance

 

=

(u,u )HP

P
s
HP-in
HP-out
Nu,u
Tu,u
)
 ,s Cpu,u ,s (Tu,u


s
PCin
PCout
En
Fb,c,t
CpHEP (Tb,c,t
Tb,c,t
)

(b,c,t)HEP

HU
Qu,ut
+

(u,u )CP

Qu +

uHPt HB

(u,ut)HPt

 

En
Fb,c,t
dH c

(b,c,t)HEP

P
CPout
CPin
s
Nu,u
Tu,u
)
 ,s Cpu,u ,s (Tu,u


ECout
ECin
En
Fb,c,t
CpHEE (Tb,c
Tb,c
)

(b,c,t)HEP

SHout
SHin
En
Fb,c,t
CpHES (Tb,t
Tb,t
)

(b,c,t)HEP

(u,ut)CPt

HU
Qu,ut


(b,c,t)HEP

En
Fb,c,t
dH b

(A.157)

Adams, T. A., II, & Barton, P. I. (2010). High-efciency power production from coal
with carbon capture. AIChE Journal, 56(12), 31203136.
Adams, T. A., II, & Barton, P. I. (2011). Combining coal gasication and natural
gas reforming for efcient polygeneration. Fuel Processing Technology, 92(3),
639655.
Agrawal, R., Singh, N. R., Ribeiro, F. H., & Delgass, W. N. (2007). Sustainable fuel for
the transportation sector. PNAS, 104(12), 48284833.
Ahmetovic, E., & Grossmann, I. E. (2010a). Optimization of energy and water consumption in corn-based ethanol plants. Industrial and Engineering Chemistry
Research, 49(17), 79727982.
Ahmetovic, E., & Grossmann, I. E. (2010b). Global superstructure optimization for
the design of integrated process water networks. AIChE Journal, 57(2), 434457.
Argonne National Laboratory. GREET 1.8b (2007). The greenhouse gases, regulated
emisssions, and energy use in transportation (GREET) model, release September
2008.
Baliban, R. C., Elia, J. A., & Floudas, C. A. (2010). Toward novel biomass, coal, and natural gas processes for satisfying current transportation fuel demands, 1: Process
alternatives, gasication modeling, process simulation, and economic analysis.
Industrial and Engineering Chemistry Research, 49(16), 73437370.
Baliban, R. C., Elia, J. A., & Floudas, C. A. (2011). Optimization framework for the simultaneous process synthesis, heat and power integration of a thermochemical
hybrid biomass, coal, and natural gas facility. Computers and Chemical Engineering, 35(9), 16471690.
Baliban, R. C., Elia, J. A., & Floudas, C. A. (2012). Simultaneous process synthesis,
heat, power, and water integration of thermochemical hybrid biomass, coal,
and natural gas facilities. Computers and Chemical Engineering, 37(10), 297327.
Baliban, R. C., Elia, J. A., Misener, R., & Floudas, C. A. (2012). Global optimization
of a MINLP process synthesis model for thermochemical based conversion of
hybrid coal, biomass, and natural gas to liquid fuels. Computers and Chemical
Engineering, 42, 6486.
Bechtel (1992). Baseline design/economics for advanced FischerTropsch technology. Contract No. DE-AC22-91PC90027.
Bechtel (1998). Aspen process owsheet simulation model of a battelle
biomass-based gasication, FischerTropsch liquefaction and combined-cycle
power plant. Contract number: DE-AC22-93PC91029. http://www.schertropsch.org/.
Bechtel Corp, Global Energy Inc., and Nexant Inc. (2003). Gasication plant cost
and performance optimization, task 2 topical report: Coke/coal gasication with
Liquids Coproduction. USDOE contract DE-AC26-99FT40342.
Cao, Y., Gao, Z., Jin, J., Zhou, H., Cohron, M., Zhao, H., et al. (2008). Synthesis gas
production with an adjustable H2 /CO ratio through the coal gasication process:
Effects of coal ranks and methane addition. Energy and Fuels, 22(3), 17201730.
Chen, Y., Adams, T. A., II, & Barton, P. I. (2011a). Optimal design and operation of static
energy polygeneration systems. Industrial and Engineering Chemistry Research,
50(9), 50995113.
Chen, Y., Adams, T. A., II, & Barton, P. I. (2011b). Optimal design and operation of
exible energy polygeneration systems. Industrial and Engineering Chemistry
Research, 50(8), 45534566.
Chiesa, P., Consonni, S., Kreutz, T., & Williams, R. (2005). Co-production of hydrogen, electricity and CO2 from coal with commercially ready technology. Part
A: Performance and emissions. International Journal of Hydrogen Energy, 30(7),
747767.
CPLEX ILOG CPLEX C++API 12.1 Referece Manual, 2009.
de Fraiture, C., Giordano, M., & Liao, Y. (2008). Biofuels and implications for agricultural water use: Blue impacts of green energy. Water Policy, 10, 6781.
de Klerk, A. (2011). FischerTropsch rening. Wiley-VCH Verlag & Co. KGaA.
Department of Energy (2005). Biomass as feedstock for a bioenergy
and bioproducts industry: The technical feasibility of a billionannual
supply.
Document
number:
DOE/GO-102005-2135.
ton
http://www1.eere.energy.gov/biomass/publications.html.
Duran, M. A., & Grossmann, I. E. (1986). Simultaneous optimization and heat integration of chemical processes. AIChE Journal, 32(1), 123138.
Elia, J. A., Baliban, R. C., & Floudas, C. A. (2010). Toward novel biomass, coal, and
natural gas processes for satisfying current transportation fuel demands, 2:
Simultaneous heat and power integration. Industrial and Engineering Chemistry
Research, 49(16), 73717388.
Elia, J. A., Baliban, R. C., Xiao, X., & Floudas, C. A. (2011). Optimal energy supply
network determination and life cycle analysis for hybrid coal, biomass, and
natural gas to liquid (CBGTL) plants using carbon-based hydrogen production.
Computers and Chemical Engineering, 35(8), 13991430.
Elia, J. A., Baliban, R. C., & Floudas, C. A. (2012). Nationwide energy supply chain analysis for hybrid feedstock processes with signicant CO2 emissions reduction.
AIChE Journal, 58(7), 21422154.
Energy Information Administration (2009). Ranking of U.S. Reneries.
http://www.eia.doe.gov/neic/rankings/reneries.htm.

56

R.C. Baliban et al. / Computers and Chemical Engineering 47 (2012) 2956

Energy Information Administration (2011). Annual Energy Outlook 2011


with Projections to 2035. Document number: DOE/EIA-0383(2011).
http://www.eta.doe.gov/oiaf/aeo/.
Floudas, C. A., Akrotirianakis, I. G., Caratzoulas, S., Meyer, C. A., & Kallrath, J. (2005).
Global optimization in the 21st century: Advances and challenges. Computers
and Chemical Engineering, 29(6), 11851202.
Floudas, C. A. (1995). Nonlinear and mixed-integer optimization. New York: Oxford
University Press.
Floudas, C. A. (2000). Deterministic global optimization: Theory, methods and applications. Dordrecht, Netherlands: Kluwer Academic Publishers.
Floudas, C. A., Ciric, A. R., & Grossmann, I. E. (1986). Automatic synthesis of optimum
heat exchanger network congurations. AIChE Journal, 32(2), 276290.
Floudas, C. A., Elia, J. A., & Baliban, R. C. (2012). Hybrid and Single Feedstock Energy
Processes for Liquid Transportation Fuels: A Critical Review. Computers and
Chemical Engineering, 41, 2451.
Floudas, C. A., & Gounaris, C. E. (2009). A review of recent advances in global optimization. Journal of Global Optimization, 45(1), 338.
Floudas, C. A., & Pardalos, P. M. (1995). State of the art in global optimization:
Computational methods and applications. Journal of Global Optimization, 7(2),
113.
Grossmann, I. E., & Martn, M. (2010). Energy and water optimization in biofuel
plants. Chinese Journal of Chemical Engineering, 18(6), 914922.
Karuppiah, R., & Grossmann, I. E. (2006). Global optimization for the synthesis of
integrated water systems in chemical processes. Computers and Chemical Engineering, 30(4), 650673.
Keil, F. J. (1999). Methanol-to-hydrocarbons: Process technology. Microporous and
Mesoporpus Materials, 29(1-2), 4966.
Kreutz, T., Williams, R., Consonni, S., & Chiesa, P. (2005). Co-production of hydrogen, electricity and CO2 from coal with commercially ready technology. Part b:
Economic analysis. International Journal of Hydrogen Energy, 30(7), 769784.
Kreutz, T. G., Larson, E. D., Liu, G., & Williams, R. H. (2008). FischerTropsch fuels from
coal and biomass. In Proceedings of the 25th international pittsburg coal conference
Larson, E. D., & Jin, H. (1999). Biomass conversion to FischerTropsch liquids: Preliminary energy balances. In Proceedings of the 4th biomass conference of the Americas
(pp. 843853).
Larson, E. D., Jin, H., & Celik, F. E. (2009). Large-scale gasication-based coproduction
of fuels and electricity from switchgrass. Biofuels, Bioproducts and Biorening, 3,
174194.
Liu, P., Pistikopoulos, E. N., & Li, Z. (2009). A mixed-integer optimization approach
for polygeneration energy systems design. Computers and Chemical Engineering,
33(3), 759768.
Liu, P., Pistikopoulos, E. N., & Li, Z. (2010a). Decomposition based stochastic programming approach for polygeneration energy systems design under uncertainty.
Industrial and Engineering Chemistry Research, 49(7), 32953305.
Liu, P., Pistikopoulos, E. N., & Li, Z. (2010b). A multi-objective optimization approach
to polygeneration energy systems design. AIChE Journal, 56(5), 12181234.
Lynd, L. R., Larson, E., Greene, N., Laser, M., Sheehan, J., Dale, B. E., et al. (2009).
The role of biomass in Americas energy future: Framing the analysis. Biofuels,
Bioproducts and Biorening, 3(2), 113123.
Martin, M., & Grossmann, I. E. (2011). Process Optimization of FT-Diesel Production
from Lignocellulosic Switchgrass. Industrial and Engineering Chemistry Research,
50(23), 1348513499.
Mobil Research and Development Corporation (1978). Research Guidance Studies to Assess gasoline from coal by methanol-to-gasoline and Sasol-type
FischerTropsch technologies. USDOE contract EF-77-C-O1-2447.

Mobil Research and Development Corporation (1986). Slurry FischerTropsch/mobil


two stage process of converting syngas to high octane gasoline. USDOE contract
DE-AC22-80PC30022.
Mobil Research and Development Corporation (1985). Two-stage process for conversion of synthesis gas to high quality transporation fuels. USDOE contract
DE-AC22-83PC60019.
Nashawi, I. S., Malallah, A., & Al-Bisharah, M. (2010). Forecasting world crude oil production using multicyclic Hubbert model. Energy and Fuels, 24(3), 17881800.
National Academy of Sciences. (2009). National Academy of Engineering, and National
Research Council. Liquid transportation fuels from coal and biomass: Technological
status, costs, and environmental issues. Washington, DC: EPA.
National Energy Technology Laboratory (2007). Cost and performance baseline for fossil energy plants. Volume 1: Bituminous coal and natural
gas to electricity nal report. Document number: DOE/NETL-2007/1281.
http://www.netl.doe.gov/energy-analyses/baseline studies.html.
National Renewable Energy Laboratory (2011). Gasoline from Wood via integrated
gasication, synthesis, and methanol-to-gasoline technologies. USDOE contract
DE-AC36-08GO28308.
National Research Council (2008). Water implications of biofuels production in the
United States. http://www.nap/edu/catalog/12039.html.
Spath, P., Aden, A., Eggeman, T., Ringer, M., Wallace, B., & Jechura, J. Biomass to hydrogen production detailed design and economics utilizing the battelle columbus
laboratory indirectly-heated gasier. Document number: NREL/TP-510-37408.
http://www.nrel.gov/docs/fy05osti/37408.pdf.
Sudiro, M., & Bertucco, A. (2007). Synthetic fuels by a limited CO2 emission process
which uses both fossil and solar energy. Energy and Fuels, 21, 36683675.
Sudiro, M., & Bertucco, A. (2009). Production of synthetic gasoline and diesel fuel
by alternative processes using natural gas and coal: Process simulation and
optimization. Energy, 34, 22062214.
Sudiro, M., Bertucco, A., Ruggeri, F., & Fontana, M. (2008). Improving process performances in coal gasication for power and synfuel production. Energy and Fuels,
22(6), 38943901.
Tabak, S. A., & Krambeck, F. J. (1985). Shaping process makes fuels. Hydrocarbon
Processing, 64(9), 7274.
Tabak, S. A., Krambeck, F. J., & Garwood, W. E. (1986). Conversion of propylene and
butylene over ZSM-5 catalyst. AIChE Journal, 32(9), 15261531.
Tabak, S. A., & Yurchak, S. (1990). Conversion of methanol over ZSM-5 to fuels and
chemicals. Catalysis Today, 6(3), 307327.
Toyir, J., Miloua, R., Elkadri, N. E., Nawdali, M., Touk, H., Miloua, F., et al. (2009).
Sustainable process for the production of methanol from CO2 and H2 using
Cu/ZnO-based multicomponent catalyst. Physics Procedia, 2(3), 10751079.
Government
Printing
Ofce
(2009).
Economic
indicators.
US
http://www.gpoaccess.giv/indicators/index.html.
Vliet, O., Faaij, A., & Turkenburg, W. (2009). FischerTropsch diesel production in a
well-wheel perspective: A carbon, energy ow and cost analysis. Energy Conversion and Management, 50(4), 855876.
Water Science and Technology Board (2008). Water Implications of Biofuels Production in the United States.
Weekman, V. W., Jr. (2010). Gazing into an energy crystal ball. Chemical Engineering
Progress, 6, 2327.
Zittel, W., Schindler, J. (2007). Crude oil: The supply outlook. Document number:
EWG-series No. 3/2007. Report to the Energy Watch Group, October 2007.

Vous aimerez peut-être aussi