Вы находитесь на странице: 1из 14

Article

pubs.acs.org/IECR

Comparative Techno-Economic and Environmental Analysis of


Ethylene and Propylene Manufacturing from Wet Shale Gas and
Naphtha
Minbo Yang† and Fengqi You*,‡

Department of Chemical and Biological Engineering, Northwestern University, Evanston, Illinois 60208, United States

Robert Frederick Smith School of Chemical and Biomolecular Engineering, Cornell University, Ithaca, New York 14853, United
States
*
S Supporting Information

ABSTRACT: In this work, we perform a comparative techno-economic and environmental


analysis for manufacturing ethylene and propylene from naphtha and from shale gas with rich
natural gas liquids (NGLs). We first propose two novel process designs for producing
ethylene and propylene from NGLs-rich shale gas. These two designs employ steam co-
cracking of an ethane−propane mixture and an integration of ethane steam cracking and
propane dehydrogenation, respectively. For benchmarking, we also consider a conventional
process design in which ethylene and propylene are produced via steam cracking of naphtha.
Detailed process models are developed for all the three designs to obtain the mass and
energy balances of each unit operation. On this basis, techno-economic analysis and life cycle
analysis are conducted for each of the three designs in order to systematically compare the production costs and life cycle
environmental impacts of ethylene and propylene manufactured from shale gas and naphtha based on the same conditions. The
economic analysis indicates that manufacturing ethylene and propylene from NGLs-rich shale gas is more attractive than from
naphtha. The environmental impact analysis shows that manufacturing ethylene and propylene from NGLs-rich shale gas results
in lower life cycle water consumption but higher life cycle greenhouse gas emissions.

1. INTRODUCTION ethylene, He and You developed three process designs by


Ethylene and propylene are important building blocks for integrating shale gas processing with ethane steam cracking,
manufacturing plastics, fibers, and other chemicals. In the which could significantly increase the overall profitability.15
United States, ethylene is primarily manufactured via steam Later, an integrated process design combining the conversion
cracking of hydrocarbons, including natural gas liquids (NGLs, of methane, ethane, and propane and the dehydration of
ethane, propane, butanes, etc.), naphtha, and gas oil.1 In bioethanol was proposed to improve the economics and
addition to ethylene, propylene is obtained as a byproduct in environmental impacts for ethylene production.16 Salkuyeh and
steam crackers, which accounts for roughly half of the total Adams proposed a polygeneration process design to produce
propylene production.2,3 In recent years, the successful ethylene and electricity from shale gas via methane oxidative
application of advanced extraction technologies resulted in a coupling, which showed the ethylene production costs could be
boom of shale gas production in the United States and provides close to those of steam cracking methods.17 However, these
extra NGLs at low costs for the chemical manufacturing publications only focus on the production of ethylene from
industry.4−8 As a result, manufacturing ethylene and propylene shale gas and do not explore the economics and environmental
from shale gas-based feedstocks (e.g., ethane and propane), implications of manufacturing propylene from shale gas. In
instead of from naphtha, is of growing interest.9 With different addition, some contributions were made to addressing the
feedstocks, ethylene and propylene may require different water management in shale gas production,18optimal design of
production costs and result in different environmental impacts. shale gas supply chain,19 and uncertainty in shale gas supply
It is of great significance to investigate the economic and chain design and optimization.20
environmental implications of ethylene and propylene Environmental impacts of manufacturing chemicals from
manufacturing from different feedstocks.10,11 Therefore, the shale gas are of particular interest. A recent concern associated
objective of this study is to compare the production of ethylene with shale gas production is the fugitive methane emissions,21
and propylene from NGL-rich shale gas and naphtha in terms which may result in high greenhouse gas (GHG) emissions.22
of economics and life cycle environmental impacts.
There are some recent publications regarding the process Received: January 25, 2017
design and synthesis for shale gas processing and upgrading.12 Revised: March 19, 2017
Shale gas is considered as the feedstock to produce syngas,13 Accepted: March 22, 2017
methanol,14 liquid fuels, and light olefins. For the production of Published: March 22, 2017

© XXXX American Chemical Society A DOI: 10.1021/acs.iecr.7b00354


Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

Figure 1. Block flow diagrams for the manufacturing of ethylene and propylene from shale gas and naphtha.

Several studies have examined the life cycle GHG emissions of consumption of ethylene and propylene manufactured from
shale gas. Laurenzi and Jersey examined the life cycle GHG shale gas and naphtha are investigated and compared.
emissions of Marcellus shale gas for power generation.23 Dale et The key contributions of this work are summarized as
al. performed a process-based life cycle assessment (LCA) for follows: (1) two novel process designs for manufacturing
Marcellus shale gas.24 The upstream carbon footprints of shale ethylene and propylene from NGLs-rich shale gas and (2)
gas and conventional natural gas were also compared.25 systematic comparison of manufacturing ethylene and
Another concern of shale gas production is the considerable propylene from NGLs-rich shale gas and naphtha based on
water usage associated with horizontal drilling and hydraulic the same conditions from economics and environmental
fracturing.26 Shale gas production is reported to consume perspectives.
notably more water than conventional natural gas.27 For olefins The rest of the paper is organized as follows. Section 2
production, the steam cracking of hydrocarbons (e.g., ethane, presents the description of each process design. Results of
propane, naphtha, etc.) is highly energy intensive and thus process simulation, economic analysis, and environmental
produces significant amounts of GHGs.28 Besides, waste heat analysis are provided and compared in Section 3. The
conclusion of this work is given in Section 4.
needs to be discharged to the ambient environment, and this is
commonly performed by cooling water systems that consume a
2. PROCESS DESCRIPTION
considerable amount of water.1 To the best of our knowledge,
there are very limited studies focusing on environmental In this study, shale gas from the Marcellus region in the United
impacts of manufacturing olefins from shale gas, except a recent States is analyzed for the production of ethylene and propylene
study on ethylene production from shale gas.29 However, it because there is an increasing attention for producing olefins in
remains inconclusive whether shale gas can be regarded as a this region from the locally produced shale gas.30 Table S1 in
low-carbon feedstock for the production of ethylene and the Supporting Information gives the composition of the raw
shale gas considered in this study. Depending on the raw shale
propylene compared with naphtha.
gas composition, two process designs for manufacturing
In this work, we first develop two novel process designs for
ethylene and propylene from shale gas are proposed as
producing ethylene and propylene from NGLs-rich shale gas. In
depicted in Figure 1(a) and (b), respectively. Each design
the first design, the ethane−propane mixture is sourced from consists of two stages: shale gas processing and olefins
shale gas at first and then stream co-cracked to produce production.
ethylene and propylene. In the second design, ethane and In the co-cracking design shown in Figure 1(a), raw shale gas
propane are derived from shale gas separately. Ethylene is is first fed into an acid gas removal unit to remove the acid
mainly produced via the steam cracking of ethane, and component. Then, the sweet gas is introduced into a
propylene is mostly manufactured via the dehydrogenation of dehydration unit to reduce the water content. NGLs are
propane. For the purpose of a systematic comparison, we also further recovered from the dry gas by a cryogenic separation
model a conventional naphtha cracking process that produces unit. The resulting methane-rich gas is pressurized and sent out
ethylene and propylene from naphtha. We develop detailed as pipeline gas, and the mixture of ethane and propane is sent
process models for all three designs to obtain the mass and to produce ethylene and propylene. At the olefins production
energy balances. Next, we conduct a techno-economic analysis stage, the ethane−propane mixture derived from shale gas and
and life cycle analysis for each of the three designs. Finally, the unreacted ethane and propane are steam co-cracked in the
production costs, life cycle GHG emissions, and life cycle water steam cracking unit. After being cooled, pressurized, and
B DOI: 10.1021/acs.iecr.7b00354
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

Figure 2. Process flowsheets for the processing of Marcellus shale gas.

purified, the cracking gas is then sent to an olefin separation rich MEA solution is depressurized via another valve (V-102)
unit and spilt into ethylene, propylene, hydrogen, methane, and preheated in a heat exchanger (E-103), and then, it is fed to
ethane, propane, and other products. a stripper (T-102). In stripper T-102, acid gas in the rich MEA
In the technology-integrated design shown in Figure 1(b), solution is stripped off, and the MEA solution is regenerated.
the shale gas processing stage is similar to that in the co- The lean MEA solution from stripper T-102 bottom is first
cracking design. The difference between these two shale gas cooled in a heat exchanger (E-103) and then pumped to the
processing stages is the NGLs fractionation unit for separating top of absorber T-101 along with the makeup MEA and water.
the ethane−propane mixture into ethane and propane. The The gas effluent from the top of stripper T-102 enters a flash
olefins production stage considers the integration of ethane drum (V-102) after being cooled. The liquid from flash drum
steam cracking and propane dehydrogenation technologies. V-102 is pumped to the top of stripper T-102, and the gas from
Ethane is fed into the ethane steam cracking unit, while flash drum V-102 is acid gas.
propane is taken as the feedstock for the propane dehydrogen- Since acid gas is removed in an amine-based unit, the sweet
ation unit. The cracking gas and the effluent from the propane gas leaving absorber T-101 is water saturated. Water must be
dehydrogenation unit are sent to an olefins separation unit, reduced to prevent hydrate formation in the cryogenic
where they are separated into ethylene, propylene, and other separation process.31 Triethylene glycol (TEG) is the typical
products. choice in most instances for water removal from natural gas.31
In addition to the two novel process designs that we propose, A TEG-based dehydration unit is depicted in Figure 2(b). The
a conventional process design for manufacturing ethylene and sweet gas enters an absorber (T-201) from the bottom and
propylene from naphtha is modeled and analyzed for contacts with a lean TEG stream. The rich TEG stream from
comparing economic and environmental performances of the the bottom of absorber T-201 is depressurized via a valve
products. As shown in Figure 1(c), naphtha is first steam (VLV-201) and flashed in a flash tank (V-201) to remove most
cracked in a stream cracking unit. Similarly, the cracking gas is hydrocarbons. Then, the rich TEG stream is fed into a
cooled, pressurized, and purified. After that, the cracking gas is distillation tower (T-202) for TEG regeneration after being
separated into ethylene, propylene, and other products in an preheated in a heat exchanger (E-203). Water and hydro-
olefins separation unit. Details of each unit are given in the carbons from the overhead condenser of tower T-202 are
following subsections. purged, and the lean TEG stream from the tower bottom is
2.1. Shale Gas Processing. Figure 2 shows the flowsheets sent to a stripper (T-203) to further reduce the water content
for the processing of Marcellus shale gas with the composition in the TEG stream. A small portion of dry gas from the top of
given in Table S1. It comprises five units, namely, acid gas absorber T-201 is used as strpping gas in stripper T-203. The
removal, dehydration, NGLs recovery, compression, and NGLs regenerated TEG stream mixed with the makeup TEG is
fractionation. pumped to the top of absorber T-201 after being cooled in a
Since CO2 in shale gas could lead to solids formation in the cooler (E-302).
cryogenic separation process for NGLs recovery, the CO2 After dehydration, shale gas is sent to a cryogenic separation
concentration should be reduced to less than 50 ppmv.31 As process for NGLs recovery. As shown in Figure 2(c), the dry
show in Figure 2(a), a monoethanolamine (MEA)-based gas is first cooled in a heat exchanger (E-301), and then, it
absorption unit is employed for acid gas removal. Raw shale enters a two-phase separator (V-301). The gas product from
gas is fed into the bottom of an absorber (T-101) and contacts separator V-301 flows through an expander (EX-301) to further
with a lean MEA solution from the top of absorber T-101. The reduce the temperature before it is fed into another separator
rich MEA solution from absorber T-101 bottom passes through (V-302). Liquid products from separators V-301 and V-302 are
a pressure relief valve (VLV-101) followed by a flash tank (V- introduced into a demethanizer (T-301). The gas product from
101) to remove dissolved light hydrocarbons. After that, the separator V-302 and that from the top of demethanizer T-301
C DOI: 10.1021/acs.iecr.7b00354
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

Figure 3. Process flowsheets of olefins production by steam co-cracking of ethane−propane mixture: (a) ethane−propane steam co-cracking and (b)
olefins separation.

are methane rich and mixed. The composition of the mixed reactor (PFR) model.29 In this work, a mass-based hydro-
product has met the corresponding specifications of pipeline carbons-to-steam ratio of 1:0.4 is applied.32 The reaction
gas, as shown from Table S1. After heat recovery in heat scheme for the steam co-cracking of an ethane−propane
exchanger E-301, a small portion of methane-rich gas is mixture is given in Table S2 in the Supporting Information.
consumed for steam generation, and the rest is introduced into The cracking gas from the cracking furnace flows through a
a gas compression unit as shown in Figure 2(d). The methane- heat exchanger (E-501) to generate high pressure (HP) steam
rich gas is pressurized to 6 MPa and sent out as pipeline gas. at 11 MPa, which is subsequently superheated in the cracking
Power generated by Expander EX-301 provides part of the furnace.1,33 The cracking gas is further cooled by another heat
power for gas compression; additional power is provided by exchanger (E-502) to preheat dilution steam and ultimately
steam turbines driven by the steam generated via the cooled to 40 °C in a quench tower by contacting with cooling
combustion of methane-rich gas. In the co-cracking design, water. After that, a five-stage compressor is employed to
the NGLs product (i.e., the ethane−propane mixture as shown pressurize the cracking gas to 3.7 MPa. The compression power
in Table S1) from the bottom of demethanizer T-301 is sent to is provided by steam turbines driven by the superheated HP
the olefins production stage. In the technology-integrated steam. During compression, the temperature of the cracking gas
design, the NGLs product is introduced into a deethanizer (T- is controlled below 100 °C by intermediate coolers to prevent
401), where the mixture is split into ethane and propane, as olefin polymerization and subsequent equipment fouling.1 Acid
illustrated in Figure 2(e). Then, ethane and propane are sent to components in the cracking gas are removed in the caustic
the olefins production stage. tower after the third stage compression, according to reactions
2.2. Co-Cracking of Ethane−Propane. In the co-cracking 1 and 2. The compressed cracking gas is introduced into the
design, the mixture of ethane and propane coming from the olefin separation unit after the water content is reduced in a
shale gas processing stage is steam co-cracked. Figure 3 shows molecular sieve dryer.
the process flowsheets of the olefins production stage in the co- CO2 + 2NaOH → 2Na 2CO3 + H 2O (1)
cracking design. Typically, a commercial steam cracking furnace
consists of convection and radiant sections.1 In the convection H 2S + 2NaOH → 2Na 2S + 2H 2O (2)
section, hydrocarbons, dilution steam, and boiler feedwater are
heated. In the radiant section, hydrocarbons are cracked into In the olefins separation unit shown in Figure 3(b), the
small molecules, such as ethylene, propylene, hydrogen, cracking gas is first cooled to about −37 °C through two heat
methane, etc. For the sake of convenience, the steam cracking exchangers (E-510 and E-504) and a cooler (E-505). Then, the
furnace is simulated using a heater combined with the plug flow cracking gas enters a separator (V-501). The gas product from
D DOI: 10.1021/acs.iecr.7b00354
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

Figure 4. Process flowsheets of olefins production by the integration of ethane steam cracking and propane dehydrogenation: (a) ethane steam
cracking, (b) propane dehydrogenation, and (c) olefins separation. The steam cracking unit is similar to that in Figure 3, and the rest of the olefins
separation section is similar to that in Figure 3 excluding a debutanizer.

separator V-501 is introduced into a series of cold boxes. Liquid in heat exchangers E-504 and E-511. A liquid stream is
products from separators V-501 to V-504 are fed into a extracted from the 99th tray of C2 splitter T-509 as a coolant in
demethanizer (T-503) where methane and hydrogen in the heat exchanger E-510 and returned to the 101st tray. The liquid
liquid phase are removed. The gas phase from separator V-504 product from the bottom of deethanizer T-504 is fed into a
is ultimately cooled to −165 °C in a heat exchanger (E-515) depropanizer (T-505). The overhead product from depropan-
and fed into separator V-505. The gas product from separator izer T-505 contains propane and propylene that are separated
V-505 is hydrogen rich, and the liquid product is methane rich. in a C3 splitter comprising two distillation columns (T-506 and
These two streams serve as coolants for heat exchanger E-515 T-507). Polymer-grade propylene with a purity of 99.5% is
and for heat exchangers E-511−E514 along with the gas obtained from the top of column T-506, and propane from the
product from demethanizer T-503. The bottom product of bottom of column T-507 is fed into cracking furnaces. Last, a
demethanizer T-503 is nearly methane free and is introduced debutanizer (T-508) is used to separate the bottom product
into a deethanizer (T-504). After being preheated and mixed from depropanizer T-505 into a C4 mixture and C5+ mixture.
with some hydrogen-rich gas, the gas product from deethanizer 2.3. Integration of Ethane Steam Cracking and
T-504 is sent to a hydrogenator (R-501), where acetylene is Propane Dehydrogenation. The technology-integrated
reacted with hydrogen and converted into ethylene or ethane, design considers the integration of ethane steam cracking and
according to reactions 3 and 4. propane dehydrogenation. Figure 4 presents the process
C2H 2 + H 2 → C2H4 flowsheets of the olefins production stage in the technology-
(3)
integrated design. Ethane sourced from shale gas along with the
C2H 2 + 2H 2 → C2H6 (4)
recycled ethane is sent to the ethane steam cracking unit, which
is similar to the one described in Figure 3. The reaction scheme
The separation of ethylene and ethane often requires large used for ethane steam cracking is given in Table S3 in the
distillation columns with 120−180 trays.34 The gas product Supporting Information. Propane derived from shale gas and
from hydrogenator R-501 is fed into a C2 splitter (T-509) with the recycled propane are merged as the feedstock for the
120 trays, after being cooled in a heat exchanger (E-506). The propane dehydrogenation unit, as shown in Figure 4(b).
ethylene product with a purity of 99.9% is drawn from the ninth Meanwhile, hydrogen is injected into the propane feed to
tray of C2 splitter T-509. The gas product from the top of C2 reduce the coke formation on the catalyst with a mole-based
splitter T-509 is ethylene rich and is further cooled in a cooler hydrogen-to-hydrocarbons ratio of 0.6:1.35 The reaction section
(E-509) to reduce the amount of gas recycled to be consists of four reactors (R-601−R-604) with fired heaters
recompressed. The liquid product from separator V-506 is (FH-601−FH-604).36 The mixture of propane and hydrogen is
routed to the first tray of C2 splitter T-509. Ethane from the introduced into the reactors after being heated to over 630
bottom of C2 splitter T-509 is depressurized and finally °C.35 In this work, reactors are modeled using the PFR model,
recycled as a feedstock of cracking furnaces, after being heated and the kinetic model for propane dehydrogenation is given in
E DOI: 10.1021/acs.iecr.7b00354
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

Figure 5. Process flowsheets of olefins production by the steam cracking of naphtha: (a) naphtha steam cracking and (b) olefins separation. The
olefins separation unit is similar to that in Figure 3.

Table S4 in the Supporting Information.35 The effluent from to fractionator T-701, and fuel oil is obtained from the bottom.
reactor R-604 is cooled to 40 °C and compressed to 1.2 MPa, Parallel to naphtha cracking furnaces, the recycled ethane and
and then, it is sent to an olefins separation unit, as shown in propane are co-cracked in a separate furnace.37 The cracking
Figure 4(c). In the separation unit, the product from the gas is cooled in heat exchangers E-707 and E-706. After that,
propane dehydrogenation unit is further cooled and presepa- these two cracking gas streams merge into one and enter a
rated in separators (V-601−V-603). Liquid products from quench water tower (T-703), where the cracking gas is
separators V-601−V-603 are introduced into a deethanizer (T- ultimately quenched to 40 °C by contacting with cooling
601) to remove C2 and lighter components. The gas product of water. The effluent from the bottom of tower T-703 consists of
the deethanizer T-601 is ethane rich and treated as a feed for heavy hydrocarbons and water, and it is separated by a three-
steam cracking. The liquid product containing C3 and heavier phase separator (V-701). The water stream along with the
components is introduced into a depropanizer. Note that the makeup water is split and recycled to the quench water tower
rest of the olefins separation unit is similar to that described in (T-703) as cooling water and to generate dilution steam. The
Figure 3 excluding a debutanizer. The gas product from hydrocarbon stream is spilt and introduced into fractionator T-
separator V-603 is hydrogen rich and is purified in a pressure 701 and another stripper (T-704). In stripper T-704, the
swing adsorption (PSA) unit. As aforementioned, a portion of hydrocarbon stream is boiled to strip off the dissolved gas. Gas
the purified hydrogen is injected into the propane feed. products from quench water tower T-703, stripper T-704, and
2.4. Naphtha Steam Cracking. For the purpose of a fair separator V-701 are mixed and pressurized to 3.7 MPa via a
comparison, a conventional naphtha cracking design is five-stage compressor. The condensation of water and hydro-
investigated in this work as well. Figure 5 presents the process carbons from the former three stages is routed into separator V-
flowsheets of the naphtha cracking design for the production of 701, while the condensate from the last two stages is fed into a
ethylene and propylene. Similar to the co-cracking of the condensate stripper (T-706). The gas product from stripper T-
ethane−propane mixture, naphtha is preheated and steam 706 is fed to the fourth stage of the compressor for
cracked in cracking furnaces. A steam-to-naphtha ratio of 0.5:1 recompression, and the bottom effluent is introduced into a
is applied in this work.1 The reaction scheme used for naphtha depropanizer in the olefins separation unit for further
steam cracking is given in Table S5 in the Supporting fractionation. After being dried by a molecular sieve, the
Information. After heat recovery in a heat exchanger (E-701), cracking gas is sent to the olefins separation unit, which is
the cracking gas is further cooled by contacting with cooling oil similar to that described in Figure 3.
in a quench oil tower. The resulting mixture is introduced into 2.5. Power and Refrigeration Generation. Power
a gasoline fractionator (T-701), where the cracking gas is generation via steam turbines is operated on the Rankine
further cooled and the heavy fraction is condensed. Most of the cycle.38 As shown in Figure 6, water is first pressurized and
condensed fraction is split and recycled to the quench oil tower heated to generate HP steam. At shale gas processing stages,
and gasoline fractionator T-701 after being cooled, while the HP steam is generated via the combustion of methane-rich gas;
rest is stripped by low pressure (LP) steam in a stripper (T- at olefins production stages, HP steam is generated by heat
702). The gas product from the top of stripper T-702 is routed recovery from the cracking furnaces and fired heaters. The HP
F DOI: 10.1021/acs.iecr.7b00354
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

modeled in detail to analyze and compare the economic and


environmental profiles of the products manufactured through
different feedstocks and routes.

3. RESULTS AND DISCUSSION


Following the Aspen HYSYS Customization Guide,41 all the
three process designs discussed in the previous section are
modeled and simulated in Aspen HYSYS V7.2. Detailed
operating parameters of important columns including dis-
tillation towers, strippers, and absorbers are listed in Table S6
Figure 6. Process flowsheet of power generation.38 in the Supporting Information. Heat integration of each process
design is extracted and investigated using Aspen Energy
Analyzer V7.2, according to the Aspen Energy Analyzer User
steam is then expanded in steam turbines, where steam energy Guide.42
is converted to mechanical power. Next, LP steam exiting steam 3.1. Mass and Energy Balances. In this study, we
turbines is used to satisfy on-site thermal requirements or consider a production scale of 1 Mt/yr ethylene for the three
condensed by cooling water. Finally, the condensed water is designs following a recent report.30 The operating time is 8000
recycled to the pump. h/yr.29 The mass and energy information on each unit
Refrigeration is used for NGLs recovery, NGLs fractionation, operation in the three designs is determined by HYSYS
and olefins separation. Figure 7 shows refrigeration cycles of simulation, and the results are summarized in Table 1.
propylene, ethylene, and methane. In the propylene cycle, To meet the given ethylene production rate, the co-cracking
propylene is first pressurized to 1.6 MPa via a multistage design consumes 969.5 million standard cubic feet per day
compressor (K-1) driven by steam turbines.39 Propylene is then (MMSCFD) of raw shale gas. At the stage of shale gas
condensed by cooling water and/or process streams. Depend- processing, pipeline gas is the main product. About 77.0% of
ing on the process requirements, the condensed propylene is the energy content of raw shale gas remains in the pipeline gas
depressurized to provide cold utilities with temperatures of −25 product, and 20.0% of that is taken by the mixture of ethane
and −40 °C. Lastly, propylene gases are gathered and and propane. Also, the methane-rich gas used for utility
recirculated to compressor K-1 for recompression. The generation occupies 2.7% of the energy content of raw shale
ethylene cycle is similar to the propylene cycle. Ethylene is gas. At the olefins production stage, the steam cracking of the
compressed to 2.0 MPa via compressor K-2 and then ethane−propane mixture results in an ethylene yield of 63.4 wt
condensed using cooling water and propylene refrigerant.39 % of the feed and a propylene yield of 14.4 wt % of the feed. It
Depending on the process requirements, ethylene is used to means that propylene and ethylene are produced with a ratio of
provide cold utilities with temperatures of −75 and −102 °C. 0.23. Typically, process energy use is expressed in terms of
In addition, methane is used to provide the cold utility with a specific energy consumption.34 The specific energy consump-
temperature of −135 °C.1 In the methane cycle, methane is tion of the olefins production stage in the co-cracking design is
pressurized to about 2.8 MPa and condensed by the ethylene determined as 23.9 GJ/t ethylene, which is slightly higher than
refrigerant. the specific energy consumption for ethane steam cracking
In this section, we present two novel process designs for reported in the existing literature.34
manufacturing ethylene and propylene from NGLs-rich shale The technology-integrated design consumes 1242.0
gas. The major difference between the two designs is the MMSCFD of raw shale gas to satisfy the given ethylene
technologies employed for converting ethane and propane into production rate. Similarly, at the shale gas processing stage,
ethylene and propylene. Also, a conventional process design for pipeline gas is the main product, which takes 76.6% of the
the steam cracking of naphtha is introduced for the purpose of energy content of raw shale gas. Ethane and propane totally
a fair comparison in systems analysis.40 Therefore, three process occupy 20.0% of the energy content of raw shale gas, and the
designs for the production of ethylene and propylene are utility generation consumes 3.1% of the energy content of raw

Figure 7. Process flowsheets of refrigeration generation: (a) propylene cycle, (b) ethylene cycle, and (c) methane cycle.

G DOI: 10.1021/acs.iecr.7b00354
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

Table 1. Mass and Energy Balances for Manufacturing Ethylene and Propylene from Shale Gas and Naphtha
co-cracking design technology-integrated design naphtha cracking design
shale gas processing stage
input
shale gas (MMSCFD) 969.5 1242.0 −
shale gas (MMBTU/h)a 47628 61012 −
MEA (kg/h) 26.0 33.4 −
TEG (kg/h) 26.2 33.6 −
output
pipeline gas (MMSCFD) 820.2 1045.7 −
pipeline gas (MMBTU/h)a 36673 46754 −
ethane−propane (t/h) 197.3 − −
ethane (t/h) − 150.9 −
propane (t/h) − 101.8 −
direct emissions
CO2 (t/h) 78.0 111.8 −
utilities
power (MW)b 97.0 139.7 −
LP steam (125 °C) (GJ/h)b 58.7 165.2 −
MP steam (175 °C) (GJ/h)b 13.3 17.1 −
HP steam (250 °C) (GJ/h)b 13.8 17.6 −
recycling cooling water (kt/h)c 59.6 89.4 −
make up water (kt/h) 1.2 1.8 −
olefins production stage
input
ethane−propane (t/h) 197.3 − −
ethane (t/h) − 150.9 −
propane (t/h) − 101.8 −
naphtha (t/h) − − 368.0
NaOH (kg/h) 60.6 56.0 84.5
output
ethylene (t/h) 125.0 125.0 125.0
propylene (t/h) 28.4 87.1 51.3
C4s (t/h) 8.4 13.1 7.2
C5+ (t/h) 1.6 0 134.2
hydrogen (t/h) 10.8 17.7 8.3
utilities
power (MW)b 88.9 142.8 93.0
LP steam (125 °C) (GJ/h)b 414 1509 447
MP steam (175 °C) (GJ/h)b 0 0 5.9
total fuel consumption (GJ/h) 2992 3977 3341
external fuel demand (GJ/h) 1643 3397 −351
recycling cooling water (kt/h)c 54.5 122.9 90.6
make up water (kt/h) 1.1 2.5 1.8
direct emissions
CO2 (t/h) 148.3 169.8 164.0
a b c
High heating value (HHV) basis. Power and steam are generated on-site. MP steam: medium pressure steam. In cooling water systems,
evaporation loss rate = 0.00153 × cooling range; blow down loss = 1.5 × evaporation loss.43 Cooling range is equal to 5 °C in this analysis.

shale gas. At the olefins production stage, ethane and propane energy consumption is identified as 26.7 GJ/t ethylene, which
are consumed with a total mass flow rate of 252.7 t/h. In terms meets the value reported in literature.34
of mass flow rate, ethylene and propylene account for 49.5 and Comparing shale gas processing stages in the two proposed
34.5 wt % of the feed, respectively. It is found that the designs, the main difference is the NGLs fractionation unit for
technology-integrated design results in a propylene-to-ethylene the separation of the ethane−propane mixture. To separate the
ratio of 0.70. The corresponding specific energy consumption is ethane−propane mixture into ethane and propane, more
determined as 31.8 GJ/t ethylene. refrigeration and LP steam are needed. As a result, in the
In the naphtha cracking design, naphtha is consumed with a technology-integrated design, slightly more methane-rich gas is
mass flow rate of 368.0 t/h to meet the given ethylene combusted and less pipeline gas is produced based on per unit
production rate. The steam cracking of naphtha shows an of raw shale gas.
ethylene yield of 34.0 wt % of the feed and a propylene yield of The three olefins production stages rely on different
13.9 wt % of the feed. It means that propylene and ethylene are technologies and feedstocks. From Table 1, we can see that
produced with a ratio of 0.41. The corresponding specific more naphtha is consumed in the naphtha cracking design than
H DOI: 10.1021/acs.iecr.7b00354
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

Table 2. Economic Evaluation Results of Shale Gas Processing Stages


designs co-cracking design technology-integrated design
raw shale gas cost (M$/yr) 1050.3 1345.5
other feeds cost (M$/yr) 3.4 3.9
plant overhead (M$/yr) 7.2 9.1
labor cost (M$/yr) 11.5 14.6
maintenance (M$/yr) 3.0 3.5
total capital cost (M$) 369.4 440.5
capital depreciation (M$/yr) 18.5 22.0
property taxes and insurance (M$/yr) 7.4 8.8
general expenses (M$/yr) 108.9 139.2
total production cost (M$/yr) 1210.2 1546.6

Table 3. Economic Evaluation Results of Olefins Production Stages


designs co-cracking design technology-integrated design naphtha cracking design
hydrocarbon feed cost (M$/yr) 350.1 451.2 1265.9
other feeds cost (M$/yr) 38.3 89.8 3.2
plant overhead (M$/yr) 3.9 5.2 4.3
labor cost (M$/yr) 2.9 3.6 3.2
maintenance (M$/yr) 5.0 6.8 5.5
total capital cost (M$) 912.6 1167.0 1042.8
capital depreciation (M$/yr) 45.6 58.4 52.1
property taxes and insurance (M$/yr) 18.3 23.3 20.9
general expenses (M$/yr) 45.9 63.1 134.0
total production cost (M$/yr) 510.0 701.4 1489.1

ethane and propane consumed in the other two designs in consumption quantity multiplied by the corresponding price.
terms of mass flows. This is because ethane and propane can Economic parameters for estimating costs of feedstocks and
achieve higher ethylene yields compared with naphtha in steam utilities are listed in Table S7 in the Supporting Information.
cracking. Besides, the technology-integrated design requires Aspen Process Economic Analyzer V7.2 is used to estimate the
more ethane and propane than the co-cracking design. The costs of labor, maintenance, and plant overhead.45 Capital
reason is that ethylene is mainly produced via the steam depreciation is estimated based on the total capital cost using a
cracking of ethane in the technology-integrated design. From straight-line method over the plant life, as given by eq 5.46 The
the viewpoint of the production of ethylene and propylene, the total capital cost of a process is also estimated using Aspen
three designs could coproduce ethylene and propylene. Process Economic Analyzer V7.2 and based on existing sources
However, the co-cracking design results in a propylene-to- as given in Table S8 in the Supporting Information. The cost of
ethylene ratio lower than the naphtha cracking design, meaning process equipment, units, or plants can be converted to current
that there exists a propylene gap if shifting feedstocks from value using the Chemical Engineering Plant Cost Index
naphtha to an ethane−propane mixture for steam crackers. The according to eq 6.47 The property taxes and insurance are
technology-integrated design could result in a propylene-to- estimated as 2% of the total project cost.46 General expenses
ethylene ratio higher than the naphtha cracking design. This including selling expenses, research and development costs, and
indicates that the propylene gap can be eliminated by the administrative expenses are estimated as 9% of the total
integration of ethane steam cracking and propane dehydrogen- production cost.48
ation. In terms of energy consumption, the olefins production TCC
stage in the technology-integrated design consumes more DP =
ls (5)
energy than those in the other two designs. The first reason is
that the propane dehydrogenation reaction is energy intensive ⎛ CEPCI 2016 ⎞⎛ S ⎞ β
and requires fired heat. The second reason is that the C2016 = C base⎜ ⎟⎜ ⎟
dehydrogenation product not only requires power for ⎝ CEPCIbase ⎠⎝ S base ⎠ (6)
compression but also needs refrigeration and steam for where DP is the capital depreciation; TCC is the total capital
separation. Note that the naphtha cracking design produces cost; ls is the life span; C2016 is the cost in 2016; Cbase is the cost
excess fuel, as the stream cracking of naphtha results in high in the base case; CEPCI2016 is the cost index in 2016; CEPCIbase
methane yields. is the cost index in the base case; S is the actual capacity; Sbase is
3.2. Economic Analysis. The total annual production cost the capacity in the base case; β is the cost scale factor.
can be estimated as the sum of direct manufacturing costs To evaluate production costs of manufacturing ethylene and
(feedstocks, utilities, labor-related operations, and mainte- propylene from the raw shale gas given in Table S1, costs for
nance), plant overhead, fixed costs (property taxes, insurance, producing ethane and propane are evaluated at first. The results
and capital depreciation), and general expenses.44 On the basis of the economic evaluation of shale gas processing stages are
of the obtained mass and energy balances, the total cost of given in Table 2. We find that raw shale gas cost dominates the
feedstocks and utilities can be determined by summing the cost total production cost. The other notable contributions come
of each feedstock and utility, which is equal to the annual from the fixed costs and general expenses. At shale gas
I DOI: 10.1021/acs.iecr.7b00354
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

processing stages, multiple products including pipeline gas, associated with the co-cracking design and 2.1 times of those
ethane, and propane (or ethane−propane mixture) are associated with the technology-integration design. Such results
produced. Thus, cost allocation is employed to estimate costs indicate that manufacturing ethylene and propylene from the
for producing ethane and propane from raw shale gas. In considered shale gas is more cost effective than from naphtha.
general, methods for cost allocation include the physical 3.3. Environmental Analysis. The environmental impacts
measure method, economic value method, and zero cost of manufacturing ethylene and propylene from NGLs-rich shale
method.49 The physical measure method is reasonable for cases gas and naphtha are systematically analyzed following the LCA
that the product market values are about the same, and the zero approach.50,51 The primary goal of this LCA study is to assess
cost method assigns all costs to the main product.49 Because and compare the life cycle environmental impacts of ethylene
economic values of pipeline gas, ethane, and propane differ manufactured from NGLs-rich shale gas and naphtha. The
greatly and pipeline gas is the main product at shale gas functional unit is defined as 1 kg of ethylene produced at the
processing stages, the physical measure method and the zero gate of plant. Because the use and end-of-life phases of the
cost method are not suitable. Therefore, the economic value ethylene products could vary significantly, a cradle-to-gate LCA
method is selected in this study. In the co-cracking design, is considered in this study.52 This LCA encompasses the
allocation factors for the pipeline gas and ethane−propane environmental impacts during the acquisition of feedstocks and
mixture are determined as 71.0% and 29.0%, respectively. In the the processing of shale gas (for manufacturing ethylene and
technology-integration design, allocation factors for pipeline propylene from shale gas), as well as the manufacturing of
gas, ethane, and propane are determined as 70.9%, 11.3%, and ethylene and propylene. The investigated life cycle impact
17.8%, respectively. As a result, the production cost for the categories include GHG emissions and water consumption.
ethane−propane mixture is estimated as $221.8/t, and The life cycle inventory is calculated based on the mass and
production costs for ethane and propane are estimated as energy balances provided in the previous subsections. Data
$145.8/t and $337.3/t, respectively. used to model GHG emissions and water consumption during
Subsequently, costs for manufacturing ethylene and feedstocks acquisition are collected from publications and the
propylene via the three designs are systematically evaluated. Ecoinvent database,53 as listed in Table S9 in the Supporting
Table 3 shows the economic evaluation results of olefins Information. Because shale gas processing and olefins
production stages in the three designs. The olefins production production stages produce multiple products, the correspond-
stage in the technology-integrated design requires the highest ing mass and energy flows as well as the associated
capital cost because of investments in the propane dehydrogen- environmental burdens must be allocated to each of the
ation unit and the expanded olefins separation unit. The steam products to accurately reflect their individual contributions to
co-cracking of the ethane−propane mixture results in a lower the environmental impacts.54 In this study, coproduct allocation
total capital cost than the steam cracking of naphtha. As shown is applied based on the mass allocation method and the
from Table 3, the naphtha cracking design requires the highest economic value allocation method.54 On the basis of the
total production cost. This is because the consumption quantity coproduct allocation, the environmental impacts for the
and price of naphtha are higher than those of ethane and production of propylene can be identified as well.
propane. Similarly, the economic value allocation method is 3.3.1. Life Cycle GHG Emissions. Figure 9 shows the
employed to evaluate the production costs of ethylene and breakdown of life cycle GHG emissions of ethylene produced
propylene. For the co-cracking design, the technology- from NGLs-rich shale gas and naphtha. For manufacturing
integrated design, and the naphtha cracking design, allocation ethylene from shale gas, shale gas production and olefins
factors of ethylene are determined as 62.3%, 41.9%, and 41.3%, production stages are major contributors for GHG emissions,
respectively, and those of propylene are determined as 18.5%, while shale gas processing stages contribute less than 10% of
38.0%, and 22.1%, respectively. On this basis, the production the total value. These observations are consistent with the
costs of ethylene and propylene are determined and compared recent studies on shale gas environmental impact analyses.55−57
as shown in Figure 8. It is shown that the production costs of For manufacturing ethylene from naphtha, the production of
ethylene and propylene are dominated by the direct naphtha causes GHG emissions close to the downstream
manufacturing costs for all the three designs. The technology- olefins production stage. From Figure 9, it is shown that
integrated design shows the best economic performance for ethylene manufactured by the steam cracking of naphtha shows
manufacturing ethylene and propylene. Manufacturing ethylene the least life cycle GHG emissions. Compared with ethylene
and propylene via the naphtha steam cracking design results in manufactured from naphtha, ethylene produced through the
the highest production costs, which are about 1.9 times of those technology-integrated design results in higher life cycle GHG
emissions by 19% using the economic value allocation method
or by 25% using the mass allocation method. Ethylene
manufactured via the co-cracking design causes the highest
life cycle GHG emissions, which are nearly 1.40 times of the life
cycle GHG emissions associated with the steam cracking of
naphtha. Such results imply that manufacturing ethylene from
the considered shale gas is not competitive in terms of GHG
emissions.
By employing coproduction allocation, we can also identify
the life cycle GHG emissions for manufacturing propylene from
NGLs-rich shale gas and naphtha, as shown in Figure 10. On
the basis of the economic value allocation, propylene produced
Figure 8. Breakdown of production costs for the manufacturing of via the naphtha steam cracking design has the lowest GHG
ethylene and propylene via the three designs (2016 dollar value). emissions, which are 71.4% and 83.8% of the GHG emissions
J DOI: 10.1021/acs.iecr.7b00354
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

Figure 9. Breakdown of life cycle GHG emissions of ethylene produced via the three designs.

life cycle water consumption. Compared with ethylene


manufactured via the co-cracking design, ethylene manufac-
tured via the technology-integrated design causes higher life
cycle water consumption by 29% based on the economic value
allocation or by 42% based on the mass allocation. The steam
cracking of naphtha results in notably higher life cycle water
consumption than the steam co-cracking of the ethane−
propane mixture by over 200%. Considering life cycle water
consumption, manufacturing ethylene from the considered
shale gas via the co-cracking design is more attractive than from
naphtha.
Figure 10. Breakdown of life cycle GHG emissions of propylene Similarly, the life cycle water consumption of propylene
produced via the three designs. manufactured via the three designs can be determined based on
coproduction allocation, as given in Figure 12. If coproduction
of propylene produced via the co-cracking design and the
technology-integrated design, respectively. On the basis of the
mass allocation, propylene produced via the naphtha steam
cracking design shows lower GHG emissions than propylene
produced via the co-cracking design and the technology-
integrated design by 27.3% and 20.1%, respectively. Regarding
life cycle GHG emissions, naphtha is more attractive than the
considered shale gas for the manufacturing of propylene.
3.3.2. Life Cycle Water Consumption. Life cycle water
consumptions of ethylene manufactured from NGLs-rich shale
gas and naphtha are compared in Figure 11. For manufacturing
ethylene from shale gas, water consumed at the olefins
production stage dominates life cycle water consumption of Figure 12. Breakdown of life cycle water consumption of propylene
ethylene. However, for manufacturing ethylene from naphtha, produced via the three designs.
water consumption associated with the production of naphtha
is more important. As shown from Figure 11, ethylene allocation is based on economic values, propylene produced via
manufactured via the co-cracking design leads to the lowest the naphtha steam cracking design leads to significantly higher
water consumption than propylene produced via the co-
cracking design and the technology-integrated design by
207.5% and 137.6%, respectively. If coproduction allocation is
performed based on mass flows, propylene produced via the
naphtha steam cracking design still shows the highest water
consumption, which is 302.6% and 213.0% of the water
consumption of propylene produced via the co-cracking design
and the technology-integrated design, respectively. In terms of
life cycle water consumption, manufacturing propylene from
the considered shale gas is more competitive than from
naphtha.
3.4. Systematic Comparison. Figure 13 provides a
comprehensive comparison of production costs, life cycle
Figure 11. Breakdown of life cycle water consumption of ethylene GHG emissions, and life cycle water consumption for
produced via the three designs. manufacturing ethylene and propylene from NGLs-rich shale
K DOI: 10.1021/acs.iecr.7b00354
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

ethane and propane do not affect the ethylene production cost


because ethane and propane are neither feedstocks nor final
products. Also, we find that a high propylene price could result
in a low ethylene production cost, which is different from the
other prices analyzed in Figure 14. In the technology-integrated
design, the ethylene production cost is more sensitive to the
propylene price than those in the other two designs. This is
because the technology-integrated design produces more
Figure 13. Relative cost and environmental impacts of ethylene and propylene than the other two designs. It is noteworthy that
propylene produced from shale gas and naphtha: (a) environmental the naphtha cracking design still results in higher ethylene
impacts based on economic value allocation and (b) environmental production cost than the other two designs even the naphtha
impacts based on mass allocation.
price decreases by 40%. Such results indicate that NGLs-rich
shale gas is more attractive than naphtha for the production of
gas via the two proposed designs and from naphtha via the ethylene from the standpoint of economics.
conventional naphtha cracking design. For each category, the As for environmental performance, economic parameters
largest value is normalized to 1. Thus, each point represents the could affect allocation factors when economic value allocation is
relative value. Although the economic value allocation method considered. Herein, we present effects of economic parameters
results in different life cycle environmental impacts for ethylene on life cycle GHG emissions of ethylene for illustration. From
and propylene, relative values considering ethylene and Figure 15, we see that the ethylene price is most important for
propylene are the same. In each figure, a point closer to the life cycle GHG emissions of ethylene. The propylene price is
center indicates the better corresponding performance. We find the second important factor after the ethylene price. Note that
that the economic value allocation method and the mass the increase in propylene price decreases life cycle GHG
allocation method show similar results: (1) Ethylene and emissions of ethylene because both ethylene and propylene are
propylene produced via the technology-integrated design have the products to which the GHG emissions are allocated. Ethane
the lowest production costs. (2) Ethylene and propylene and propane prices also affect life cycle GHG emissions of
produced through the co-cracking design result in the least ethylene in the co-cracking design and the technology-
water consumption. (3) Ethylene and propylene produced via integrated design, but they have little effect on the naphtha
the naphtha cracking design lead to the lowest GHG emissions. cracking design. Additionally, we find that life cycle GHG
From Figure 13, we also find that manufacturing ethylene and emissions of ethylene produced via the naphtha cracking design
propylene from NGLs-rich shale gas is more attractive than are always lower than those of ethylene produced via the other
from naphtha considering production costs and water two designs under various changes.
consumption. However, manufacturing ethylene and propylene
from naphtha is more competitive than from NGLs-rich shale
4. CONCLUSION
gas in terms of GHG emissions.
3.5. Sensitivity Analysis. In order to better understand the This work addressed the comparison of manufacturing ethylene
impacts of altering economic parameters on the economic and and propylene from NGLs-rich shale gas and naphtha regarding
environmental performance of manufacturing ethylene and economic and environmental performance. For manufacturing
propylene from NGLs-rich shale gas and naphtha, a sensitivity ethylene and propylene from shale gas, we proposed two novel
analysis is conducted. For illustration, Figure 14 presents the process designs. We also investigated a conventional design
sensitivity analysis result of ethylene production cost based on considering the steam cracking of naphtha for the production of
economic value changes of raw shale gas, ethane, propane, ethylene and propylene. On the basis of process modeling and
naphtha, ethylene, and propylene. It is shown that the feedstock HYSYS simulation, we conducted techno-economic analysis
price and the ethylene price have great impacts on the ethylene and life cycle analysis for the three designs. The economic
production cost. In the co-cracking design and the technology- analysis showed that manufacturing ethylene and propylene
integrated design, changes in ethane and propane prices also from NGLs-rich shale gas via the co-cracking design and the
lead to notable effects on the ethylene production cost. technology-integrated design resulted in lower production costs
However, in the naphtha cracking design, price changes of than from naphtha by 48.2% and 52.3%, respectively. As for

Figure 14. Sensitivity analysis of ethylene production cost: (a) co-cracking design, (b) technology-integrated design, and (c) maphtha cracking
design.

L DOI: 10.1021/acs.iecr.7b00354
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

Figure 15. Sensitivity analysis of life cycle GHG emissions of ethylene: (a) co-cracking design, (b) technology-integrated design, and (c) naphtha
cracking design.

environmental impacts, we assessed the life cycle GHG (3) Allen, K. The Changing Dynamics and New Challenges Facing
emissions and life cycle water consumption for the manufactur- the North American Petrochemical industry. http://www.slideshare.
ing of ethylene and propylene. On the basis of the economic net/14Balmoral/the-changing-dynamics-and-new-challenges-facing-
value allocation, ethylene and propylene manufactured from the-north-american-petrochemical-industry (accessed October 1,
NGLs-rich shale gas via the co-cracking design and the 2016).
technology-integrated design led to higher GHG emissions (4) DeRosa, S. E.; Allen, D. T. Impact of natural gas and natural gas
than those produced from naphtha by 40.0% and 19.4%, liquids supplies on the United States chemical manufacturing industry:
production cost effects and identification of bottleneck intermediates.
respectively. This indicated that naphtha could be regarded as a
ACS Sustainable Chem. Eng. 2015, 3, 451−459.
low-carbon feedstock for the production of ethylene and (5) Siirola, J. J. The impact of shale gas in the chemical industry.
propylene compared with NGLs-rich shale gas. However, AIChE J. 2014, 60, 810−819.
manufacturing ethylene and propylene from naphtha caused (6) Gao, J.; You, F. Design and optimization of shale gas energy
significantly higher water consumption than from NGLs-rich systems: Overview, research challenges, and future directions. Comput.
shale gas via the co-cracking design and the technology- Chem. Eng. 2017, DOI: 10.1016/j.compchemeng.2017.01.032.
integrated design by 207.5% and 137.6%, respectively. We also (7) Hughes, J. D. A reality check on the shale revolution. Nature
obtained similar results by investigating the environmental 2013, 494, 307−308.
impacts based on the mass allocation method. (8) Kerr, R. A. Natural Gas From Shale Bursts Onto the Scene.


*
ASSOCIATED CONTENT
S Supporting Information
Science 2010, 328, 1624−1626.
(9) Jenkins, S. Propylene production via propane dehydrogenation.
Chem. Eng. 2014, 121, 27−28.
The Supporting Information is available free of charge on the (10) Foster, J. Can Shale Gale Save the Naphtha Crackers? https://
www.platts.com/IM.Platts.Content/InsightAnalysis/
ACS Publications website at DOI: 10.1021/acs.iecr.7b00354.
IndustrySolutionPapers/ShaleGasReport13.pdf (accessed October 28,
Input data, reaction kinetics, and model parameters. 2016).
(PDF) (11) Gong, J.; You, F. Sustainable design and synthesis of energy


systems. Curr. Opin. Chem. Eng. 2015, 10, 77−86.
AUTHOR INFORMATION (12) Gong, J.; Yang, M.; You, F. A systematic simulation-based
process intensification method for shale gas processing and NGLs
Corresponding Author
recovery process systems under uncertain feedstock compositions.
*Phone: (607) 255-1162. Fax: (607) 255-9166. E-mail: fengqi. Comp ut. Chem. Eng. 201 6, DOI: 10.1016/j.compche-
you@cornell.edu. meng.2016.11.010.
ORCID (13) Martinez-Gomez, J.; Nápoles-Rivera, F.; Ponce-Ortega, J. M.; El-
Fengqi You: 0000-0001-9609-4299 Halwagi, M. M. Optimization of the production of syngas from shale
Notes gas with economic and safety considerations. Appl. Therm. Eng. 2017,
The authors declare no competing financial interest. 110, 678−685.


(14) Julián-Durán, L. M.; Ortiz-Espinoza, A. P.; El-Halwagi, M. M.;
Jiménez-Gutiérrez, A. Techno-economic assessment and environ-
ACKNOWLEDGMENTS
mental impact of shale gas alternatives to methanol. ACS Sustainable
The authors acknowledge financial support from National Chem. Eng. 2014, 2, 2338−2344.
Science Foundation (NSF) CAREER Award (CBET-1643244). (15) He, C.; You, F. Shale Gas Processing Integrated with Ethylene
This invited contribution is part of the I&EC Research special Production: Novel Process Designs, Exergy Analysis, and Techno-
issue for the 2017 Class of Influential Researchers. Economic Analysis. Ind. Eng. Chem. Res. 2014, 53, 11442−11459.

■ REFERENCES
(1) Zimmermann, H.; Walzl, R. Ethylene. In Ullmann’s Encyclopedia
(16) He, C.; You, F. Toward more cost-effective and greener
chemicals production from shale gas by integrating with bioethanol
dehydration: Novel process design and simulation-based optimization.
of Industrial Chemistry; Wiley-VCH Verlag GmbH & Co.: Weinheim, AIChE J. 2015, 61, 1209−1232.
Germany, 2009. (17) Khojasteh Salkuyeh, Y.; Adams, T. A., II A novel polygeneration
(2) Zinger, S. The High Stakes Gamble In the Olefins Industry- process to co-produce ethylene and electricity from shale gas with zero
Market Risks and Rewards for New Petrochemical Construction. CO2 emissions via methane oxidative coupling. Energy Convers.
http://lca.org/assets/doc/Wood_Mackenzie_Zinger_Risks_and_ Manage. 2015, 92, 406−420.
Rewards_for_Petrochemical_Construction.pdf (accessed October 20, (18) Gao, J.; You, F. Optimal design and operations of supply chain
2016). networks for water management in shale gas production: MILFP

M DOI: 10.1021/acs.iecr.7b00354
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

model and algorithms for the water-energy nexus. AIChE J. 2015, 61, energy and environmental comparative analysis. Sol. Energy 2014, 105,
1184−1208. 669−678.
(19) Gao, J.; You, F. Game Theory Approach to Optimal Design of (41) Aspen HYSYS Customization Guide; Aspen Technology, Inc.:
Shale Gas Supply Chains with Consideration of Economics and Life Burlington, MA, 2010.
Cycle Greenhouse Gas Emissions. AIChE J. 2017, DOI: 10.1002/ (42) Aspen Energy Analyzer User Guide; Aspen Technology, Inc.:
aic.15605. Burlington, MA, 2009.
(20) Gao, J.; You, F. Deciphering and handling uncertainty in shale (43) Morvay, Z. K.; Gvozdenac, D. D. Cooling Tower. In Applied
gas supply chain design and optimization: Novel modeling framework Industrial Energy and Environmental Management; Wiley-IEEE Press,
and computationally efficient solution algorithm. AIChE J. 2015, 61, 2008.
3739−3755. (44) Seider, W. D.; Seader, J. D.; Lewin, D. R. Product and Process
(21) Alvarez, R. A.; Pacala, S. W.; Winebrake, J. J.; Chameides, W. L.; Design Principles: Synthesis, Analysis and Evaluation, 3rd ed.; John Wiley
Hamburg, S. P. Greater focus needed on methane leakage from natural & Sons, 2009.
gas infrastructure. Proc. Natl. Acad. Sci. U. S. A. 2012, 109, 6435−6440. (45) Getting Started with Relative Economics in Aspen HYSYS; Aspen
(22) Howarth, R. W.; Santoro, R.; Ingraffea, A. Methane and the Technology, Inc.: Burlington, MA, 2011.
greenhouse-gas footprint of natural gas from shale formations. Clim. (46) Anderson, J. Determining Manufacturing Costs. CEP Magazine,
Change 2011, 106, 679−690. January 2009; pp 27−31.
(23) Laurenzi, I. J.; Jersey, G. R. Life cycle greenhouse gas emissions (47) Lozowski, D. Chemical engineering plant cost index (CEPCI).
and freshwater consumption of Marcellus shale gas. Environ. Sci. Chem. Eng. 2016, 123, 96.
(48) Peters, M. S.; Timmerhaus, K. D. Cost Estimation. In Plant
Technol. 2013, 47, 4896−4903.
Design and Economics for Chemical Engineers; McGraw-Hill: New York,
(24) Dale, A. T.; Khanna, V.; Vidic, R. D.; Bilec, M. M. Process based
1991.
life-cycle assessment of natural gas from the Marcellus Shale. Environ.
(49) Tayyari, F.; Parsaei, H. R. Joint Cost Allocation to Multiple
Sci. Technol. 2013, 47, 5459−5466.
Products: Cost Accounting v. Engineering Techniques. In Economics of
(25) Weber, C. L.; Clavin, C. Life cycle carbon footprint of shale gas:
Advanced Manufacturing Systems; Springer, 1992; pp 189−199.
Review of evidence and implications. Environ. Sci. Technol. 2012, 46, (50) Environmental Management-Life Cycle Assessment: Principles and
5688−5695. Framework; International Standards Organization: Geneva, 2006.
(26) Yang, L.; Grossmann, I. E.; Manno, J. Optimization models for (51) Environmental Management-Life Cycle Assessment: Requirements
shale gas water management. AIChE J. 2014, 60, 3490−3501. and Guidelines; International Standards Organisation: Geneva, 2006.
(27) Clark, C. E.; Horner, R. M.; Harto, C. B. Life cycle water (52) Lorenz, E. Life-cycle assessment in US codes and standards. PCI
consumption for shale gas and conventional natural gas. Environ. Sci. J. 2014, 59.
Technol. 2013, 47, 11829−36. (53) Ecoinvent Database, Ecoinvent: Zurich, Switzerland, 2016.
(28) Ghanta, M.; Fahey, D.; Subramaniam, B. Environmental impacts (54) Ayer, N. W.; Tyedmers, P. H.; Pelletier, N. L.; Sonesson, U.;
of ethylene production from diverse feedstocks and energy sources. Scholz, A. Co-product allocation in life cycle assessments of seafood
Appl. Petrochem. Res. 2014, 4, 167−179. production systems: review of problems and strategies. Int. J. Life Cycle
(29) He, C.; You, F. Deciphering the true life cycle environmental Assess. 2007, 12, 480−487.
impacts and costs of the mega-scale shale gas-to-olefins projects in the (55) Gao, J.; You, F. Economic and Environmental Life Cycle
United States. Energy Environ. Sci. 2016, 9, 820−840. Optimization of Noncooperative Supply Chains and Product Systems:
(30) Gonzalez, B.; Mehta, P. Ethylene Projects in the Ethane-Heavy Modeling Framework, Mixed-Integer Bilevel Fractional Programming
Northeast US. http://blogs.platts.com/2016/03/01/ethylene-projects- Algorithm, and Shale Gas Application. ACS Sustainable Chem. Eng.
ethane-northeast-us/ (accessed October 20, 2016). 2017, DOI: 10.1021/acssuschemeng.7b00002.
(31) Kidnay, A. J.; Parrish, W. R.; McCartney, D. G. Fundamentals of (56) Gao, J.; You, F. Shale Gas Supply Chain Design and Operations
Natural Gas Processing, 2nd ed.; CRC Press, 2011. toward Better Economic and Life Cycle Environmental Performance:
(32) Sundaram, K. M.; Froment, G. F. Modeling of thermal cracking MINLP Model and Global Optimization Algorithm. ACS Sustainable
kineticsI: Thermal cracking of ethane, propane and their mixtures. Chem. Eng. 2015, 3, 1282−1291.
Chem. Eng. Sci. 1977, 32, 601−608. (57) Allen, D. T. Methane emissions from natural gas production and
(33) Alizadeh, M.; Sadrameli, S. Modeling of Thermal Cracking use: reconciling bottom-up and top-down measurements. Curr. Opin.
Furnaces Via Exergy Analysis Using Hybrid Artificial Neural Chem. Eng. 2014, 5, 78−83.
Network−Genetic Algorithm. J. Heat Transfer 2016, 138, 042801.
(34) Ren, T.; Patel, M.; Blok, K. Olefins from conventional and heavy
feedstocks: Energy use in steam cracking and alternative processes.
Energy 2006, 31, 425−451.
(35) Chin, S.; Radzi, S.; Maharon, I.; Shafawi, M. Kinetic model and
simulation analysis for propane dehydrogenation in an industrial
moving bed reactor. International Journal of Chemical, Molecular,
Nuclear, Materials and Metallurgical Engineering 2011, 5, 260−266.
(36) Gregor, J.; Wei, D. UOP Oleflex Process for Light Olefin
Production. In Handbook of Petrochemicals Production Processes;
Meyers, R. A., Ed.; McGraw-Hill Education: New York, 2004.
(37) Aly, S. Ethylene from Naphtha by Millisecond Cracking with
Front-End Demethanization. http://www.diquima.upm.es/old_
diquima/docencia/tqindustrial/docs/etileno.pdf (accessed October
15, 2016).
(38) Combined Heat and Power Technology Fact Sheet Series: Steam
Turbine; United States Department of Energy, 2016.
(39) Mafi, M.; Naeynian, S. M.; Amidpour, M. Exergy analysis of
multistage cascade low temperature refrigeration systems used in olefin
plants. Int. J. Refrig. 2009, 32, 279−294.
(40) Yue, D.; You, F.; Darling, S. B. Domestic and overseas
manufacturing scenarios of silicon-based photovoltaics: Life cycle

N DOI: 10.1021/acs.iecr.7b00354
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX

Вам также может понравиться