Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

pubs.acs.

org/IECR Article

Renewable Methanol Synthesis through Single Step Bi-reforming of


Biogas
Nazanin Entesari, Alain Goeppert, and G. K. Surya Prakash*
Cite This: https://dx.doi.org/10.1021/acs.iecr.0c00755 Read Online

ACCESS Metrics & More Article Recommendations *


sı Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: Biogas is considered a renewable source of carbon for methanol


production. For this, biogas, containing mainly CH4 and CO2, is first reformed into
syngas (a CO/H2 mixture) followed by conversion to methanol. Conventional
reformers, however, require additional upgrading steps to adjust the H2:CO ratio
in syngas to 2:1; ideal for methanol synthesis. We formerly introduced the concept
Downloaded via UPPSALA UNIV on May 20, 2020 at 19:09:12 (UTC).

of bi-reforming that provides the ideal H2:CO ratio by combining dry and steam
reforming in one stage without the need for additional syngas ratio adjustments.
Based on these experimental bi-reforming findings, we have now developed a
thermodynamic model to determine the optimal conditions for the highest
possible carbon conversion and minimum coke formation. The proposed process
based on bi-reforming was found to be an efficient alternative, delivering the ideal H2:CO ratio of 2 for methanol synthesis with no
coke formation (a common challenge in conventional reformers) and complete carbon conversion at atmospheric pressure and
temperatures of around 900 °C.

1. INTRODUCTION supply is a core component of any green process employing


According to the “Global Energy and CO2 Emission Status renewable energies.12
Report”, world energy demand rose by 2.3% in 2018, the The Methanol Economy concept, as suggested by Olah et
highest rate in the past decade.1 Fossil fuels, including coal, al.,9,13,14 offers a feasible approach to store surplus renewable
petroleum oil, and natural gas, remain our primary sources of energy while at the same time closing the carbon cycle. The
energy, particularly for heat and electricity generation and in concept is based on the conversion of carbon-containing
the transportation sector.2 On the other hand, the same report sources to methanol and derived hydrocarbons such as
indicated that the higher energy consumption led to a 1.7% dimethyl ether (DME). Methanol is a safe and easy-to-handle
increase in energy-related CO2 emissions, to a total of 33.1 Gt liquid and is considered as one of the most promising green
per year, pointing out the enormity of the problem that we are energy carriers. Methanol has a high octane number of greater
facing. To address these issues, humanity clearly needs to than 100 and an energy density of 15.8 MJ·L−1. This soot-free
rapidly shift to renewable energy sources and sustainable combustion fuel with low NOx emissions is blendable with
processes with lower or, better yet, no carbon emissions.1 other hydrocarbons and has the potential for replacing gasoline
To address concerns of anthropogenic CO2 emissions and and diesel fuels in many applications. Methanol, produced on
their associated consequences, including global warming, ocean an annual 100 million metric ton scale, is already a major
acidification, and sea-level rise, many researchers have been chemical feedstock for a wide range of valuable products such
proposing possible solutions. A number of approaches have as formaldehyde, DME, olefins, and polypropylene.15−17
been explored such as carbon capture and storage (CCS) or Methanol is also an excellent hydrogen carrier and fuel for
CO2 utilization (CCU) to produce value-added products like direct methanol fuel cells (DMFCs). More importantly, a
methanol, dimethyl ether, plastics, and a variety of other mismatch between supply from renewables and demand can be
hydrocarbons.3−9 However, CCS and CCU by themselves will well managed by storing the energy in the form of methanol:
not be able to solve our fossil fuel addiction problem. For that, easy to store and transport liquid with high energy density. As
we need to rapidly wean ourselves off these forms of fossilized a result, considerable attention has been given to the
sunshine and replace them with alternative energy resources
such as wind, solar, geothermal, hydro, and biogas.10,11 One of
the issues with the most scalable renewable energy sources, Received: February 13, 2020
solar and wind, is their intermittence and fluctuation on a Revised: May 4, 2020
minute, hourly, daily, and seasonal basis. A steady supply of Accepted: May 5, 2020
energy over time is a crucial yet challenging matter when it
comes to renewable resources. Therefore, efficient storage and
transportation of surplus energy to address fluctuations in

© XXXX American Chemical Society https://dx.doi.org/10.1021/acs.iecr.0c00755


A Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

development of efficient reaction routes and processes with pressure swing adsorption (PSA), water gas shift reactors
low carbon footprint to convert carbon oxides into (WGSRs), or membrane separation units.
methanol.7,9 As an alternative to conventional reforming approaches, we
Methanol is synthesized through direct hydrogenation of have reported in a series of publications a bi-reforming process
CO and CO2, as shown in reactions (R1) and (R2).18,19 The (RRR7) that combines steam and dry reforming of CH4 and
hydrogenation is an exothermic and equilibrium-limited CO2 in a single step. Bi-reforming (BR) was carried out over
reaction. Therefore, higher pressures and lower temperatures NiO/MgO at high temperatures and pressures, and the
thermodynamically favor methanol production.20 Concurrent gaseous product with a H2:CO ratio of 2:1 was obtained,
to methanol synthesis, CO2 and H2 can also undergo the which we have called metgas, a mixture highly suitable for
reverse water gas shift reaction (R3) (RWGSR), a reversible methanol synthesis8,30−32
reaction that produces CO and H2O. In contrast to
hydrogenation, RWGSR is endothermic and has a weak SR: CH4 + H 2O F CO + 3H 2 , ΔH298K = + 206 kJ· mol−1
dependency on pressure. Yield of methanol synthesis is (R4)
strongly dependent on the extent of the competing RWGSR.
DR: CH4 + CO2 F 2CO + 2H 2 , ΔH298K = + 247 kJ ·mol−1
Currently, Cu/ZnO/Al2O3-based heterogeneous catalysts are
used for the commercial production of methanol.9 These (R5)
catalysts deliver a 25−70% conversion and more than 99% 1
selectivity per pass at a pressure of 50−100 bar and elevated PO: CH4 + O2 F CO + 2H 2 , ΔH298K = − 36 kJ ·mol−1
2
temperatures of 220−300 °C.21,22 Designing a more effective (R6)
catalyst, either heterogeneous or homogeneous, for methanol
synthesis that would require lower temperature, pressure, or BR: 3CH4 + 2H 2O + CO2 F 4CO + 8H 2 , ΔH298K = +659 kJ·mol−1
have higher yields is still the focus of many research (R7)
programs.20,23 Driven by sustainability-oriented goals, in recent years more
attention has been given to the development of green
CO + 2H 2 F CH3OH, ΔH298K = − 91 kJ·mol−1
processes for renewable energy generation and storage, several
(R1) of them centered around the synthesis of methanol from
CO2 + 3H 2 F CH3OH + H 2O, ΔH298K = − 50 kJ· mol−1
biogas.33−36 Biogas is the product of anaerobic digestion of
organic wastes that can originate from agricultural waste,
(R2)
industrial waste, or landfills, each resulting in a biogas with a
CO2 + H 2 F CO + H 2O, ΔH298K = + 41 kJ·mol−1 different composition. Generally, CH4 (50−80%) and CO2
(25−50%) are the main constituents of biogas streams, with
(R3)
potential impurities present such as sulfides, nitrates, hydrogen,
Synthesis gas (syngas) is a mixture of CO and H2 and serves and siloxanes.37 Biogas with its high content in both CH4 and
as the main feed stream for methanol reactors.24 Natural gas, CO2, if released into the atmosphere, can significantly
coal, and petroleum are the main resources for syngas contribute to the greenhouse effect. In the framework of the
production.25 Biogas, a renewable energy source with high Methanol Economy concept, however, biogas can be converted
carbon content (mainly CH4 and CO2), has received into value-added methanol and contribute to closing the
considerable attention recently as an alternative green carbon cycle. In this case, after initial purification to separate
feedstock for syngas generation. Conventionally, these sulfides, siloxanes, and other impurities,38−40 the carbon-rich
carbon-containing streams go through partial oxidation or gaseous mixture of CO2 and CH4 can then be directed to a
reforming to obtain syngas. reforming unit to first produce syngas, followed by the
Depending on the source of carbon and the intended production of methanol.41
application, there are three common reforming methods for The high CO2 content in biogas favors dry reforming for
syngas production: steam reforming (SR), dry reforming syngas production. DR has been studied mainly over Ni-based
(DR), and partial oxidation (PO). Steam reforming is a widely catalysts.42−44 In a methanol synthesis process studied by
implemented approach to produce syngas by reaction with Santos et al.,45 biogas from four different sources was
water at high temperature. SR commonly uses natural gas converted into syngas in a dry reformer and later to methanol
(mostly methane) as the carbon source (R4).26,27 Dry in a tubular reactor. Among the selected biogas compositions,
reforming, on the other hand, is the oxidation of methane biogas from palm oil with the highest CH4:CO2 ratio of 1.4
with CO2 (R5).28 Partial oxidation (PO) is an alternative path resulted in the highest methanol yield. Hernández et al.46,47
to convert CH4 and added O2 to CO and H2 (R6).29 focused on the optimization of a biogas-to-methanol process in
Practically, the syngas composition or, in other words, the a dry reformer with two objectives: the reduction of the cost of
H2:CO ratio defines its area of application from the production methanol production and the reduction of the carbon footprint
of hydrocarbons to a wide range of chemical feedstocks, of the overall process. Economic optimization in this work
including methanol. SR delivers a H2:CO ratio of 3:1, DR a favored a DR followed by a water gas shift unit, which adjusts
ratio of 1:1, and PO a ratio of 2:1. If syngas is intended to be the ratio of H2:CO in syngas before feeding it into the
converted into methanol in an economic and energy-efficient methanol reactor.
manner, a H2:CO ratio of 2:1 is desired. Among the standard CH4 decomposition and CO disproportionation are
reforming processes, syngas from PO has the optimal ratio for common side reactions of the dry reforming process.48 The
methanol synthesis. However, the risk of thermal runaway and generated carbon formed in these reactions may deposit on the
local hotspots is high in the case of PO. In most instances, the catalyst and reduce its activity through coking. Addition of
H2:CO ratio needs to be adjusted for the intended application steam during reforming was shown to prevent coking to a large
after the reformer module with intermediate units such as extent.14 Furthermore, based on Le Chatelier’s principle, the
B https://dx.doi.org/10.1021/acs.iecr.0c00755
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

presence of steam shifts the equilibrium in the WGS reaction Ultimately, the process is optimized to improve the overall
in favor of more H2 production. This is particularly beneficial yield of methanol production.
in the case of a syngas intended for methanol synthesis, since
the generated H2 can then be used for the hydrogenation of 2. METHODS
carbon oxides. Vita et al.49,50 compared thermodynamic A schematic of the proposed process based on bi-reforming of
aspects and performance of the three main reforming processes biogas to metgas and subsequent conversion to methanol is
(DR, SR, and PO) in an attempt to use the syngas produced in shown in Figure 1. The initial purification steps are not
the reformer directly for methanol synthesis without any
intermediate adjustments in the syngas composition. Their
feasibility studies showed that steam reforming of biogas with a
high CH4 content (CH4:CO2 of 2.3) resulted in the highest
methane conversion (over 80%). In another study, thermody-
namics of large-scale steam reforming of biogas with a
CH4:CO2 of 1.4 to methanol with a capacity of 1.8 × 106
m3 per year was studied by Ghosh et al.51 In their proposed
process, CO2 was first separated from CH4, and the obtained
methane-rich gas was directed to a SR unit. Instead of venting
it, the separated CO2 was mixed with the syngas coming out of
Figure 1. Process schematic for the conversion of biogas to green
the reformer. In this work by Ghosh, three scenarios were
methanol with a single stage bi-reformer.
studied: (i) the RWGS reaction and hydrogenation of CO to
methanol, (ii) the direct CO2 hydrogenation to methanol in a
single reactor, and (iii) the direct CO2 hydrogenation to considered in this work, and it is assumed that the biogas
methanol with intermediate steps to remove the water stream only contains CH4 and CO2. Prior to the reformer,
generated by the RWGS during hydrogenation. Among these biogas is mixed with steam and is brought to the temperature
scenarios, steam reforming of biogas followed by direct and pressure of the BR. The amount of CO2 in the BR is
hydrogenation of CO2 in two consecutive reactors with regulated with a makeup line connected to a reservoir tank of
water removal in between resulted in the highest methanol captured CO2, potentially separated from the air.55−58 Biogas is
yield. Alternatively, the endothermic steam and dry reforming converted to metgas in BR and is directed to the next section
of methane were combined with the exothermic partial of the process for methanol synthesis. Excess amount of steam
oxidation to produce syngas via tri-reforming. Hernandez et in the feed gas of the reactor negatively impacts the overall
al.52 proposed a process converting biogas to syngas in such a methanol yield. Therefore, the current process considers a
manner in a tri-reformer, followed by a WGS reactor to adjust liquid trap after the reformer to control the water content in
the amount of CO in the syngas composition. An optimal the metgas, if necessary. The stream leaving BR unit goes
H2:CO ratio for methanol synthesis was achieved using a through a membrane unit that separates the unreacted CH4
biogas with a high CH4 content and by adding steam to the and recycles it back to the reformer. The CH4-free metgas,
WGSR unit. composed mainly of CO and H2, is then fed into a packed bed
The thermodynamic studies discussed above have confirmed tubular reactor for methanol synthesis. Temperature and
that biogas is indeed a viable source of carbon for green pressure of the metgas entering the reactor are controlled with
methanol synthesis. Depending on the origin of the biowaste, a heat exchanger and compressor units placed before the reactor.
variety of biogas compositions were considered and higher A flash separation is used to separate the gas and liquid phases
CH4:CO ratios in biogas were shown to improve the overall after the reactor. Unreacted gases after the reaction are
methanol yield of the process. Additionally, a cost-effective and recycled back into the reactor to increase the overall yield of
efficient methanol production required a H2:CO ratio of 2 in the system. The methanol in the collected liquid can be
the syngas. Commonly, the CO and H2 contents of the syngas separated from water by simple distillation.
were adjusted with WGS reactors, pressure swing adsorptions, A flowsheet of the proposed process for renewable methanol
or membranes placed after the reformer unit.53,54 To eliminate synthesis from biogas was developed and simulated in Aspen
these often costly and energy-intensive intermediate steps, we Plus, as shown in Figure 2. Soave Redlich Kwong (SRK)
propose here a process for renewable methanol synthesis equations of state for nonideal gases were chosen in the
through bi-reforming of biogas in a single stage bi-reformer. simulation to describe the thermodynamics of the overall
The metgas produced after the BR has the ideal H2:CO ratio process.
of 2 and can be used directly for methanol synthesis without 2.1. Biogas. The composition of the biogas is defined by its
any further adjustments in its composition. source, being from a landfill, a waste water treatment plant, or
Our extensive experimental studies on bi-reforming have from other organic waste streams. When the chemical
shown a promising performance over NiO/MgO at pressures composition in biogas from various sources is compared, on
of 1−42 bar and temperatures of 830−910 °C. The NiO/MgO average CH4 comprises 50−75% of the biogas, while CO2 has a
catalyst is shown to be stable for bi-reforming under these share of 25−50%.40,59 In the case of a single stage bi-reformer
conditions over a time period of 320 h.30−32 These and based on the stoichiometries in reaction (R7), a high
experimental findings served as a basis for the more detailed methane content in biogas is desired to assure an efficient
thermodynamic analyses presented in this work to study the performance. Therefore, a biogas stream with a CH4:CO2 ratio
overall process schematic for the conversion of biogas to of 3 was selected for metgas synthesis in this study. If required,
methanol. Here, we evaluate the impact of key parameters on the biogas composition was adjusted with captured CO2 from
the overall performance of the process implementing BR and sources such as the exhaust of power plants or by direct CO2
compare the results with the conventional approaches. capture from the air.56−58
C https://dx.doi.org/10.1021/acs.iecr.0c00755
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

Figure 2. Process flowsheet for renewable methanol synthesis from biogas.

Impurities such as nitrates, sulfides, and siloxanes are lower temperatures. The amount of solid carbon formed in the
removed from biogas in pretreatment steps before the reformer was monitored to assure long-term stability of the BR
reformer. Particular attention is given to H2S and other operation.
sulfur-containing compounds that tend to poison the Ni-based In a system with N components, the total Gibbs free energy
catalysts used in this work for bi-reforming, as well as the (GT) is the sum of the chemical potential of all of the
downstream methanol catalyst. There are several approaches to components involved, as shown in eq 1. In this equation, i
reduce the concentration of such compounds to the ppm range indicates each component, n is the number of moles, and μ is
by passing the biogas through absorption units (with NaOH), the chemical potential. Chemical potential is defined in eq 2, in
adsorption units (on iron oxides or activated carbon), or its which G0f i is the standard Gibbs free energy of formation for
biological transformation to hydrates. The remaining sulfur- component i, R is the constant of gases, T is the temperature, f i
containing compounds can potentially still deactivate the is the fugacity at an operating condition, and f 0i is the fugacity
catalyst depending on the source of biogas, the type of the of component i at a standard state.
sulfur-containing compound, and the operating condition in N
the reformer. Studies on steam reforming of biogas on Ni- GT = ∑ niμi
based catalysts have shown that most of the catalyst poisoning (1)

ij f yz
i=1
occurs at temperatures lower than 700 °C and that the catalyst

μi = G 0fi + RT lnjjjj i0 zzzz


jf z
can effectively be reactivated by increasing the temperature in

k i{
the reformer up to 900 °C.60,61 It was also shown in the
literature that the presence of small amounts of sulfur in a (2)
CO2-rich feed gas might be beneficial and prevent carbon For a mixture at equilibrium, eqs 3 and 4 are valid to
formation on Ni-based reforming catalysts in a process known reformulate the total Gibbs free energy as in eq 5. Here, y is the
as sulfur-passivated reforming process (SPARG).62 mole fraction in gas phase, ⌀ is the fugacity coefficient, P is the
The pretreatment units for biogas before the BR were not pressure of the system, and P0 is the standard state pressure (1
modeled in this work. Additionally, the operating temperature bar).
of the BR was optimized to be higher than 700 °C; hence, the
impact of the sulfur-containing impurities on the catalyst fi = yi ⌀iP (3)
activity was considered to be minimal.
2.2. Bi-reforming (BR). In the proposed upgrading process G 0fi = ΔG 0fi and f i0 = P 0 (4)

i i yi ⌀iP yzyzz
∑ nijjjjjΔG0fi + RT lnjjjj
of biogas, impurities were not considered in the biogas
zzz
0 z zz
N
feedstock for the single stage bi-reformer. Experimental and

k k P {{
thermodynamic studies confirmed that reformers generally GT =
operate in an equilibrium state.63 It was hence assumed that i=1 (5)
the BR in this process was at equilibrium as well and was able The total Gibbs free energy at equilibrium in BR is then
to reach the highest possible conversion at any given condition. minimized by applying the Lagrange multiplier (eq 6). In the
To have a better understanding of the BR and to perform multiplier term, αik is the number of atoms of element k in
comparative thermodynamic analysis, a dry reformer was also component i, and λk is the total mass of element k.

ji i yi ⌀iP yz zy
∑ nijjjjjΔG0fi + RT lnjjjj z+ ∑ λkαik zzzzz = 0
modeled in this work under the same conditions as BR without

0 z z
N

kP {
the added steam to the biogas inlet.
k {
At equilibrium, Gibbs free energy of a system is at its
minimum. Therefore, minimization of Gibbs free energy64 was i=1 k (6)
applied to determine conversion and metgas composition. For 2.3. MeOH Reactor. Methanol synthesis is an exothermic
this purpose, the R-Gibbs reactor model in Aspen Plus is best and equilibrium-limited reaction; hence, the effective heat
suited to describe the bi-reforming unit. In addition to the management in the reactor has a significant impact on the
main reactants and products of bi-reforming and WGS performance of the reaction. Gas-phase reactions to produce
reactions listed in reactions R3 and R7, DME and light methanol are conventionally carried out in packed bed tubular
hydrocarbons were also considered in calculations to account reactors under isothermal or adiabatic conditions. There have
for the potential side reactions. Undesired coking and carbon also been efforts to shift the equilibrium either by in situ
deposition might occur during the reforming, particularly at removal of one or more products in a membrane reactor or by
D https://dx.doi.org/10.1021/acs.iecr.0c00755
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

altering the temperature and concentration profiles over the Both reforming of biogas and synthesis of methanol are
catalytic bed in a transient reverse flow reactor such as STAR limited by equilibrium. Unreacted gases after each of these
configuration introduced by Vanden Bussche and Fro- units were recycled back to the individual unit with a ratio of
ment.36,65 0.7 to improve the overall yield of each section. Under optimal
In the methanol synthesis section of this process, we have reaction conditions, the stream leaving the BR will mainly
focused on the technology commonly practiced industrially. consist of CO and H2 with minimal traces of other
Therefore, the metgas from the BR unit is directed into a components such as water and CH4. Presence of H2O and
packed bed isothermal reactor. The reactor is simulated with CH4 in trace amounts in the metgas will not have a significant
the R-Plug model in Aspen Plus that is equipped with a impact on the methanol reaction. In this case, the metgas
countercurrent cooling jacket to control the heat released leaving the BR can alternatively be sent to the methanol
during the exothermic hydrogenation. Methanol synthesis was reactor without going through the intermediate separation and
carried out over the CuO/ZnO/Al2O3 catalyst. This catalyst is recycle steps (SEP 1 and SEP 2 will be eliminated), to reduce
the well-established commercial catalyst for large-scale the overall cost of the process. CH4 residuals can then be
methanol production, with stability over the course of several separated from the liquid phase after the reactor to be either
years. Alternatively, there are also reports of other heteroge- recycled or burned to generate heat required for the process.
neous catalysts for methanol synthesis in the presence of both
CO and CO2. The commercial catalyst in this work could be 3. RESULTS
replaced by such catalysts while the composition of the inlet 3.1. Biogas to Metgas. The equilibrium state in the bi-
stream to the reactor is in the range of (H2 − CO2)/(CO + reforming unit was studied at temperatures in the range of
CO2) ∼ 2.66 If necessary, a makeup line with the captured CO2 300−1000 °C and a pressure of 1−15 bar. Key parameters for
from the air can be used to balance the composition in this the thermodynamic analysis of BR are listed in Table 1. Both
stream.
Kinetics of CO/CO2 hydrogenations, as well as WGSR over Table 1. Process Parameters in BR
the CuO/ZnO/Al2O3 catalysts, is well studied in the literature.
Vanden Bussche and Froment proposed a comprehensive temperature [°C] 300−1000
study of the kinetics for these reactions based on Langmuir− pressure [bar] 1−15
Hinshelwood−Hougen−Watson (LHHW)-type rate equa- CH4/CO2 [−] 3/1
tions.67 Predefined models in Aspen Plus use LHHW-type H2O/CO2 [−] 2/1
kinetics in the form of eq 7. For the original kinetics found in inlet flow rate [mL·min−1] 100
the literature to be compatible with the predefined form of catalyst NiO/MgO
LHHW in the software, equations and parameters were
transformed by Luyben68 (eqs 8 and 9) and reaction rates
were redefined in kmol·kgcat−1·s−1. In this set of equations, R experimental and numerical preliminary studies on BR with the
stands for the rate of reaction, P is the partial pressure, KE is model developed here showed no trace of DME, ethane,
the equilibrium constant, and k is the reaction rate constant. propane, butane, and higher hydrocarbons as potential
byproducts. Therefore, this work limited the components to
(driving force term)
R = (kinetic term) the ones involved in BR and WGS reactions. Biogas conversion
(adsorption term) (7) and metgas composition, as well as the amount of solid coke
ÄÅ ÉÑ
ÅÅ Ñ
ÅÅ1 − 1 ijj PMeOHPH3 2O yzzÑÑÑ
formed in BR, were determined. A biogas stream with the same

ÅÅ KE1 j PCO2PH2 zÑ
{ÑÑÖ
composition was sent to a DR unit, and the performance of the
ÅÇ k
= (k4PCO2PH 2) Ä ÉÑ3
BR was evaluated by comparing the values obtained for biogas
ÅÅ i PH O y ÑÑ
ÅÅ
ÅÅ1 + k 3jjj P 2 zzz + k1 PH2 + k 2PH2OÑÑÑ
RMeOH reforming in a dry reformer under the same conditions. Finally,

ÅÅÇ k H2 { ÑÑÖ
the energy input to the BR was evaluated.
In the first step, the model was validated based on our prior
experimental studies.30−32 The experimental studies on bi-

ÄÅ ÉÑ
(8)
reforming were carried out in a pressurized tubular flow reactor
ÅÅ Ñ
ÅÅ1 − 1 ijjj PCOPH2O yzzzÑÑÑ
at 7−42 bar and 830−910 °C. The catalyst was mixed with a
ÅÅ KE2 PCO2 PH2 Ñ
ÅÇ k {ÑÑÖ
RRWGS = (k5PCO2) ÄÅ ÉÑ
tabular alumina (typically 100 mg catalyst to 900 mg alumina)

ÅÅ Ñ
ÅÅ1 + k 3ijjj PH2O yzzz + k1 PH + k 2PH OÑÑÑ
and was placed inside an alumina tube in the center of the

ÅÅ 2 ÑÑÑÖ
ÅÇ k {
metallic tube. The reactive gases leaving the reactor were
PH2 2 directed to a gas chromatograph for analysis. The conditions in
(9)
the present model were set to mimic the ones in these
experimental studies. Figure 3 shows both the theoretical and
2.4. Compressors and Recycle Streams. Previous experimental conversions of CH4, as well as CO2, in BR in a
studies have shown that compressors are generally the most temperature range of 830−910 °C and a fixed pressure of 7
costly parts of the methanol production processes.68,69 The bar. It can be seen that the developed model and experiments
reforming section of this process was therefore carried out with NiO/MgO are in an acceptable range of agreement in
under atmospheric pressure to reduce the total cost of the terms of carbon conversion. The minimal deviation between
system. Pressure of the metgas was then increased to the the theoretical and experimental data is likely due to the
reaction pressure in a polytropic compressor positioned before accuracy of the trajectory fitted into the limited experimental
the methanol reactor. More detailed economic studies on the data in a given temperature range. Therefore, it is valid to
costs associated with the compressor will be addressed in consider this model as a reliable tool for more detailed
future studies. evaluation and optimization on BR.
E https://dx.doi.org/10.1021/acs.iecr.0c00755
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

bar. The observed trend shows no significant difference


between DR and BR. However, at lower temperatures,
competing WGS reaction becomes dominant and its impact
can be detected in the form of a dip in the amount of
converted CO2 for both BR and DR, as shown in Figure 5. In

Figure 3. Theoretical and experimental conversions of (a) CH4 and


(b) CO2 at 7 bar in BR.

Figure 4 shows the theoretical conversion of CH4 in BR Figure 5. Conversion of CO2 for a biogas stream with a CH4:CO2
compared to that in DR at different temperatures and ratio of 3:1 in (a) BR and (b) DR.
pressures, as determined with the equilibrium model. Higher
temperatures favor methane reforming, and consequently CH4 the case of BR, the generated CO is more likely to react with
conversion increases with increasing temperatures until the available steam to form CO2. A more substantial decrease
reaching full conversion at temperatures above 900 °C and 1 in CO2 conversion at temperatures lower than 700 °C is
therefore observed in the thermodynamic studies of bi-
reforming. Based on Le Chatelier’s principle, lower pressures
shift the equilibrium in reforming reactions toward metgas.
The decreasing trend is confirmed by CH 4 and CO 2
conversions at higher pressures in both bi-reforming and dry
reforming of biogas.
At lower temperatures, CO can potentially convert to CO2
and solid carbon in the so-called Boudouard reaction, resulting
in coking and deactivation of the catalyst.9 Thermodynamic
studies of dry reforming of biogas in Figure 6b indicate that up
to 50% of the inlet carbon content might transform into coke
during reforming, even at higher temperatures. However, it was
stated in previous studies49,70 that the addition of steam to the
reforming reaction suppresses undesired coke formation
particularly at higher temperatures. In the case of BR, Figure
6a shows that, as the temperature reaches 600 °C, steam
reforming becomes the thermodynamically preferred route,
and consequently the potential for solid carbon formation in
BR begins to fall. Based on these findings, BR can be assumed
to be coke-free at temperatures above 900 °C.
Among reforming approaches, BR has the advantage of
delivering metgas with the optimal composition for the
subsequent methanol synthesis. The thermodynamic studies
on BR in Figure 7 confirm that, at temperatures above 800 °C,
the H2:CO ratio in metgas reaches 2 with no significant
Figure 4. Conversion of CH4 for a biogas stream with a CH4:CO2 dependency on pressure. In the case of a DR, over-
ratio of 3:1 in (a) BR and (b) DR. stoichiometric CH4 in the biogas (CH4:CO2 ratio of 3:1
F https://dx.doi.org/10.1021/acs.iecr.0c00755
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

shown in Figure 8. From the sensitivity analysis, it can be


observed that the energy input increases at higher reforming

Figure 8. Heat duty of BR.

temperatures for this endothermic reaction. In contrast to


temperature, Figure 8 indicates that higher pressures favor the
reverse reaction and reduce the total energy intake of the BR.
Hence, operating conditions in BR need to be selected in a
manner that minimizes the energy input to the unit.
Additionally, internal heat integration strategies can potentially
be employed to provide the energy requirements of units such
Figure 6. Solid carbon formed per mole of inlet carbon for a biogas
stream with a CH4:CO2 ratio of 3:1 in (a) BR and (b) DR.
as BR through the heat generated within the process.
For an efficient transformation of biogas to metgas, we
propose here a process implementing a single stage bi-reformer
with no need for intermediate upgrading steps to adjust the
composition. The base temperature of BR is set to 900 °C and
pressure to 1 bar, under which nearly full conversion of biogas
is achieved with no sign of carbon deposition. The gaseous
mixture leaving the BR under this condition has a molar
composition of H2: 65%, CO: 32%, CO2: 0.4%, CH4: 0.5%,
and less than 1% of H2O. Kinetic studies on the hydrogenation
of carbon oxides have shown that the presence of CO2 in
syngas potentially improves the performance of the methanol
synthesis.46 As a result, metgas composed of H2 and CO with a
minor share of CO2 was sent directly to the methanol reactor
with no further upgrading.
3.2. Metgas to Methanol. The temperature and pressure
of the metgas were adjusted before the methanol reactor
through a series of heat exchange and compression steps.
Kinetics of hydrogenation requires elevated temperatures, yet
this equilibrium-limited reaction is hindered thermodynami-
cally at higher temperatures. Therefore, determination of the
optimal temperature for the methanol reaction is critical in
defining the overall performance of the system. The reaction
conditions in this process were adopted from the established
Lurgi methanol plant.45,71 The pressure was set to 80 bar and
temperature to 220 °C. Methanol synthesis was carried out in a
Figure 7. H2:CO ratio for a biogas stream with a CH4:CO2 ratio of tubular reactor over the CuO/ZnO/Al2O3 commercial catalyst.
3:1 in (a) metgas after BR and (b) syngas after DR. The packed bed reactor had a diameter of 0.03 m and a length
of 1 m to mimic the high-pressure pilot setup at the Loker
Hydrocarbon Research Institute for further experimental
instead of 1:1) results in a syngas leaving the reformer with a studies on this process.32 The main parameters used in this
H2:CO ratio of 3. While metgas from BR can directly be sent work to describe the methanol synthesis reactor are listed in
into the MeOH reactor, the composition of the syngas leaving Table 2. The yield of the process was improved by recycling
the DR requires further adjustments in a WGS reactor or a unreacted gases back to the reactor.
pressure swing adsorber before it is fed into the reactor. Under the selected conditions, studies show that the MeOH
One of the key factors impacting the economy of the process reactor delivers an overall conversion of 98% with a total heat
is the amount of heat input to each unit. The impact of duty of 5.1 kW for the reactor, with the geometry and flow
variations in pressure and temperature on the required amount rates given in Tables 1 and 2. The amount of heat generated by
of heat in the bi-reformer was determined in this work and is the exothermic hydrogenation of CO2/CO in the MeOH
G https://dx.doi.org/10.1021/acs.iecr.0c00755
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

Table 2. Modeling Parameters in MeOH Reactor detailed economic evaluations and optimizations can further
improve the overall efficiency and energy demand of this
diameter [m] 0.03
process.


length [m] 1
catalyst density [kg·m−3] 2000
ASSOCIATED CONTENT
catalyst bed void [%] 50
* Supporting Information

The Supporting Information is available free of charge at
reactor can be used in the bi-reforming section to either https://pubs.acs.org/doi/10.1021/acs.iecr.0c00755.
preheat the biogas or conduct the distillation of methanol after
Thermal properties of the main process streams for
the reaction. Therefore, the energy efficiency of the process can
energy analysis (Table S1); and heat exchanger network
be significantly improved by applying suitable heat integration
with integrated cooling water and heating streams
strategies to use this generated heat within the process.
(Figure S1) (PDF)
4. CONCLUSIONS
Biogas has gained increasing attention in recent years both as a
renewable source of energy and as a potential chemical
■ AUTHOR INFORMATION
Corresponding Author
feedstock for a variety of value-added products. In the G. K. Surya Prakash − Loker Hydrocarbon Research Institute
framework of the proposed methanol economy and in its and Department of Chemistry, University of Southern
goal to close the anthropogenic carbon cycle, biogas can be California, University Park, Los Angeles, California 90089-
used as a renewable and sustainable carbon source. CH4 and 1661, United States; orcid.org/0000-0002-6350-8325;
CO2 present in biogas are converted into methanol, a liquid Email: gprakash@usc.edu
fuel with high energy content and raw material for a wide range
of chemicals and products. In this concept, biogas is first Authors
reformed to a gaseous mixture of H2 and CO, which is then Nazanin Entesari − Loker Hydrocarbon Research Institute and
directed to a methanol synthesis reactor. Ideally, a H2:CO ratio Department of Chemistry, University of Southern California,
of 2 in the syngas is desired for an efficient methanol synthesis. University Park, Los Angeles, California 90089-1661, United
Conventional reforming methods, however, require additional States
separation or feeding steps to adjust the composition of syngas Alain Goeppert − Loker Hydrocarbon Research Institute and
exiting from the reformer. Department of Chemistry, University of Southern California,
In our prior publications,30−32 we introduced a bi-reforming University Park, Los Angeles, California 90089-1661, United
process that combines steam and dry reforming in a single step States; orcid.org/0000-0001-8667-8530
to produce metgas with an optimal H2:CO ratio of 2 for Complete contact information is available at:
methanol synthesis. As a result, intermediate upgrading units https://pubs.acs.org/10.1021/acs.iecr.0c00755
such as a water gas shift reactor or pressure swing adsorption
units were no longer required. This results in a potential Notes
improvement in the economics and energy efficiency of the The authors declare no competing financial interest.


process. Our experimental studies showed a promising
performance for bi-reforming of CH4 with CO2 and water ACKNOWLEDGMENTS
over NiO/MgO as the catalyst. Based on our previous The authors dedicate the work to their good friend, Professor
findings,8,31,55 we proposed here a process concept for Grzegorz Mlostoń, University of Lodz, Poland, on the occasion
renewable methanol synthesis through bi-reforming of biogas. of his 70th birthday. The authors gratefully acknowledge the
A thermodynamic model was developed, and the impact of key support of their work by the Loker Hydrocarbon Research
parameters on the methanol yield and energy demand was Institute at the University of Southern California.


evaluated. The results were compared with conventional biogas
reforming processes. REFERENCES
Studies on bi-reforming of biogas herein were aimed at
determining the optimal conditions in a reformer to produce a (1) Global Energy and CO2 Status Report, IEA, 2018.
(2) Short-Term Energy Outlook (STEO), US Energy Information
syngas for methanol synthesis with an optimized H2:CO ratio, Administration, 2019.
highest possible carbon conversion, and minimum coke (3) Rackley, S. A. Carbon Capture and Storage; Butterworth-
formation. The findings suggest that intermediate syngas Heinemann, 2009.
upgrading units are inevitable after the simple dry reforming of (4) Cheah, W. Y.; Ling, T. C.; Juan, J. C.; Lee, D. J.; Chang, S. J.;
biogas. Additionally, coking can be a major drawback in dry Show, P. L. Biorefineries of Carbon Dioxide: From Carbon Capture
reformers. Employing a bi-reformer, which operates at a high and Storage (CCS) to Bioenergies Production. Bioresour. Technol.
temperature of about 900 °C and atmospheric pressure, 2016, 215, 346−356.
assured a coke-free reaction with nearly complete conversion (5) Orr, F. M. Carbon Capture, Utilization, and Storage: An Update.
of the carbon species in the biogas feed. The optimal H2:CO SPE J. 2018, 23, 2444−2455.
ratio of 2 in the final metgas was confirmed under these (6) Bui, M.; Adjiman, C. S.; Bardow, A.; Anthony, E. J.; Boston, A.;
Brown, S.; Fennell, P. S.; Fuss, S.; Galindo, A.; Hackett, L. A.; Hallett,
conditions. Additionally, the energy required for the reforming J. P.; Herzog, H. J.; Jackson, G.; Kemper, J.; Krevor, S.; Maitland, G.
unit can be partly provided by the heat released during the C.; Matuszewski, M.; Metcalfe, I. S.; Petit, C.; Puxty, G.; Reimer, J.;
hydrogenation of carbon oxides. Thermodynamic analysis on Reiner, D. M.; Rubin, E. S.; Scott, S. A.; Shah, N.; Smit, B.; Trusler, J.
the single-stage bi-reforming has shown the process to be a P. M.; Webley, P.; Wilcox, J.; Mac Dowell, N. Carbon Capture and
feasible and effective alternative for the conversion of biogas to Storage (CCS): The Way Forward. Energy Environ. Sci. 2018, 11,
a value-added liquid energy carrier such as methanol. More 1062−1176.

H https://dx.doi.org/10.1021/acs.iecr.0c00755
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

(7) Olah, G. A.; Goeppert, A.; Prakash, G. K. S. Chemical Recycling (31) Olah, G. A.; Goeppert, A.; Czaun, M.; Prakash, G. K. S. Bi-
of Carbon Dioxide to Methanol and Dimethyl Ether: From Reforming of Methane from Any Source with Steam and Carbon
Greenhouse Gas to Renewable, Environmentally Carbon Neutral Dioxide Exclusively to Metgas (CO−2H2) for Methanol and
Fuels and Synthetic Hydrocarbons. J. Org. Chem. 2009, 74, 487−498. Hydrocarbon Synthesis. J. Am. Chem. Soc. 2013, 135, 648−650.
(8) Olah, G. A.; Prakash, G. K. S.; Goeppert, A. Anthropogenic (32) Olah, G. A.; Goeppert, A.; Czaun, M.; Mathew, T.; May, R. B.;
Chemical Carbon Cycle for a Sustainable Future. J. Am. Chem. Soc. Prakash, G. K. S. Single Step Bi-Reforming and Oxidative Bi-
2011, 133, 12881−12898. Reforming of Methane (Natural Gas) with Steam and Carbon
(9) Olah, G. A.; Goeppert, A.; Prakash, G. K. S. Beyond Oil and Gas: Dioxide to Metgas (CO-2H2) for Methanol Synthesis: Self-Sufficient
The Methanol Economy, 3rd ed.; Wiley-VCH, 2018. Effective and Exclusive Oxygenation of Methane to Methanol with
(10) Stolten, D.; Scherer, V. Transition to Renewable Energy Systems; Oxygen. J. Am. Chem. Soc. 2015, 137, 8720−8729.
Wiley-VCH, 2013. (33) Yang, L.; Ge, X.; Wan, C.; Yu, F.; Li, Y. Progress and
(11) Pimentel, D. Biofuels, Solar and Wind as Renewable Energy Perspectives in Converting Biogas to Transportation Fuels. Renewable
Systems, Benefits and Risks; Springer: The Netherlands, 2008. Sustainable Energy Rev. 2014, 40, 1133−1152.
(12) Weitemeyer, S.; Kleinhans, D.; Vogt, T.; Agert, C. Integration (34) Patel, S. K. S.; Kumar, V.; Mardina, P.; Li, J.; Lestari, R.; Kalia,
of Renewable Energy Sources in Future Power Systems: The Role of V. C.; Lee, J.-K. Methanol Production from Simulated Biogas
Storage. Renewable Energy 2015, 75, 14−20. Mixtures by Co-Immobilized Methylomonas Methanica and Methyl-
(13) Olah, G. A. After Oil and Gas: Methanol Economy. Catal. Lett. ocella Tundrae. Bioresour. Technol. 2018, 263, 25−32.
2004, 93, 1−2. (35) Riaz, A.; Zahedi, G.; Klemeš, J. J. A Review of Cleaner
(14) Olah, G. A. Beyond Oil and Gas: The Methanol Economy. Production Methods for the Manufacture of Methanol. J. Cleaner
Angew. Chem., Int. Ed. 2005, 44, 2636−2639. Prod. 2013, 57, 19−37.
(15) Lefevere, J.; Mullens, S.; Meynen, V.; Van Noyen, J. Structured (36) Bozzano, G.; Manenti, F. Efficient Methanol Synthesis:
Catalysts for Methanol-to-Olefins Conversion: A Review. Chem. Pap. Perspectives, Technologies and Optimization Strategies. Prog. Energy
2014, 68, 1143−1153. Combust. Sci. 2016, 56, 71−105.
(16) Chen, D.; Moljord, K.; Holmen, A. A Methanol to Olefins (37) Wellinger, A.; Murphy, J.; Baxter, D. The Biogas Handbook:
Review: Diffusion, Coke Formation and Deactivation on SAPO Type Science, Production and Applications; Woodhead Publishing, 2013.
Catalysts. Microporous Mesoporous Mater. 2012, 164, 239−250. (38) Divsalar, A.; Entesari, N.; Dods, M. N.; Prosser, R. W.;
(17) Á lvarez, A.; Bansode, A.; Urakawa, A.; Bavykina, A. V.; Egolfopoulos, F. N.; Tsotsis, T. T. A UV Photodecomposition
Wezendonk, T. A.; Makkee, M.; Gascon, J.; Kapteijn, F. Challenges in Reactor for Siloxane Removal from Biogas: Modeling Aspects. Chem.
the Greener Production of Formates/Formic Acid, Methanol, and Eng. Sci. 2018, 192, 359−370.
DME by Heterogeneously Catalyzed CO2 Hydrogenation Processes. (39) Lau, C. S.; Allen, D.; Tsolakis, A.; Golunski, S. E.; Wyszynski,
M. L. Biogas Upgrade to Syngas through Thermochemical Recovery
Chem. Rev. 2017, 117, 9804−9838.
Using Exhaust Gas Reforming. Biomass and Bioenergy 2012, 40, 86−
(18) Bertau, M.; Offermanns, H.; Plass, L.; Schmidt, F.; Wernicke,
95.
H. J. Methanol: The Basic Chemical and Energy Feedstock of the Future,
(40) Rasi, S.; Lehtinen, J.; Rintala, J. Determination of Organic
Asinger’s Vision Today; Springer: Berlin, 2014.
Silicon Compounds in Biogas from Wastewater Treatments Plants,
(19) Goeppert, A.; Czaun, M.; Jones, J. P.; Prakash, G. K. S.; Olah,
Landfills, and CO-Digestion Plants. Renewable Energy 2010, 35,
G. A. Recycling of Carbon Dioxide to Methanol and Derived
2666−2673.
Products-Closing the Loop. Chem. Soc. Rev. 2014, 43, 7995−8048.
(41) Patel, S. K. S.; Mardina, P.; Kim, D.; Kim, S. Y.; Kalia, V. C.;
(20) Jadhav, S. G.; Vaidya, P. D.; Bhanage, B. M.; Joshi, J. B.
Kim, I. W.; Lee, J. K. Improvement in Methanol Production by
Catalytic Carbon Dioxide Hydrogenation to Methanol: A Review of Regulating the Composition of Synthetic Gas Mixture and Raw
Recent Studies. Chem. Eng. Res. Des. 2014, 92, 2557−2567. Biogas. Bioresour. Technol. 2016, 218, 202−208.
(21) Behrens, M.; Studt, F.; Kasatkin, I.; Kuhl, S.; Havecker, M.; (42) Goula, M. A.; Charisiou, N. D.; Papageridis, K. N.; Delimitis,
Abild-Pedersen, F.; Zander, S.; Girgsdies, F.; Kurr, P.; Kniep, B.-L.; A.; Pachatouridou, E.; Iliopoulou, E. F. Nickel on Alumina Catalysts
Tovar, M.; Fischer, R. W.; Norskov, J. K.; Schlogl, R. The Active Site for the Production of Hydrogen Rich Mixtures via the Biogas Dry
of Methanol Synthesis over Cu/ZnO/Al2O3 Industrial Catalysts. Reforming Reaction: Influence of the Synthesis Method. Int. J.
Science 2012, 336, 893−897. Hydrogen Energy 2015, 40, 9183−9200.
(22) Centi, G.; Perathoner, S. Advances in Catalysts and Processes for (43) Goula, M. A.; Charisiou, N. D.; Siakavelas, G.; Tzounis, L.;
Methanol Synthesis from CO2; Springer, 2013. Tsiaoussis, I.; Panagiotopoulou, P.; Goula, G.; Yentekakis, I. V. Syngas
(23) Goeppert, A.; Olah, G. A.; Prakash, G. K. S. Toward a Production via the Biogas Dry Reforming Reaction over Ni Supported
Sustainable Carbon Cycle. Green Chem. 2018, 919−962. on Zirconia Modified with CeO2 or La2O3 Catalysts. Int. J. Hydrogen
(24) Weissermel, K.; Arpe, H. J. Industrial Organic Chemistry; WILY- Energy 2017, 42, 13724−13740.
VCH, 1997. (44) Barrai, F.; Jackson, T.; Whitmore, N.; Castaldi, M. J. The Role
(25) Basile, A.; Dalena, F. Methanol: Science and Engineering; of Carbon Deposition on Precious Metal Catalyst Activity during Dry
Elsevier, 2017. Reforming of Biogas. Catal. Today 2007, 129, 391−396.
(26) Van Hook, J. P. Methane-Steam Reforming. Catal. Rev. 1980, (45) Santos, R. O. D.; Santos, D. D. S.; Prata, D. M. Simulation and
21, 1−51. Optimization of a Methanol Synthesis Process from Different Biogas
(27) Ashrafi, M.; Pröll, T.; Pfeifer, C.; Hofbauer, H. Experimental Sources. J. Cleaner Prod. 2018, 186, 821−830.
Study of Model Biogas Catalytic Steam Reforming: 1. Thermody- (46) Hernández, B.; Martín, M. Optimal Process Operation for
namic Optimization. Energy Fuels 2008, 22, 4182−4189. Biogas Reforming to Methanol: Effects of Dry Reforming and Biogas
(28) Usman, M.; Wan Daud, W. M. A.; Abbas, H. F. Dry Reforming Composition. Ind. Eng. Chem. Res. 2016, 55, 6677−6685.
of Methane: Influence of Process Parameters - A Review. Renewable (47) Hernández, B.; Martín, M. Optimal Integrated Plant for
Sustainable Energy Rev. 2015, 45, 710−744. Production of Biodiesel from Waste. ACS Sustainable Chem. Eng.
(29) Choudhary, T. V.; Choudhary, V. R. Energy-Efficient Syngas 2017, 5, 6756−6767.
Production through Catalytic Oxy-Methane Reforming Reactions. (48) Pawar, V.; Ray, D.; Subrahmanyam, C.; Janardhanan, V. M.
Angew. Chem., Int. Ed. 2008, 47, 1828−1847. Study of Short-Term Catalyst Deactivation Due to Carbon
(30) Olah, G. A.; Prakash, G. K. S.; Goeppert, A.; Czaun, M.; Deposition during Biogas Dry Reforming on Supported Ni Catalyst.
Mathew, T. Self-Sufficient and Exclusive Oxygenation of Methane and Energy Fuels 2015, 29, 8047−8052.
Its Source Materials with Oxygen to Methanol via Metgas Using (49) Vita, A.; Italiano, C.; Previtali, D.; Fabiano, C.; Palella, A.;
Oxidative Bi-Reforming. J. Am. Chem. Soc. 2013, 135, 10030−10031. Freni, F.; Bozzano, G.; Pino, L.; Manenti, F. Methanol Synthesis from

I https://dx.doi.org/10.1021/acs.iecr.0c00755
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

Biogas: A Thermodynamic Analysis. Renewable Energy 2018, 118, (70) Effendi, A.; Zhang, Z.-G.; Hellgardt, K.; Honda, K.; Yoshida, T.
673−684. Steam Reforming of a Clean Model Biogas over Ni/Al2O3 in
(50) Bozzano, G.; Pirola, C.; Italiano, C.; Pelosato, R.; Vita, A.; Fluidized- and Fixed-Bed Reactors. Catal. Today 2002, 77, 181−189.
Manenti, F. Biogas: A Possible New Pathway to Methanol? Comput.- (71) Chen, L.; Jiang, Q.; Song, Z.; Posarac, D. Optimization of
Aided Chem. Eng. 2017, 40, 523−528. Methanol Yield from a Lurgi Reactor. Chem. Eng. Technol. 2011, 34,
(51) Ghosh, S.; Uday, V.; Giri, A.; Srinivas, S. Biogas to Methanol: A 817−822.
Comparison of Conversion Processes Involving Direct Carbon
Dioxide Hydrogenation and via Reverse Water Gas Shift Reaction.
J. Cleaner Prod. 2019, 217, 615−626.
(52) Hernandez, B.; Martin, M. Optimization for Biogas to
Chemicals via Tri-Reforming. Analysis of Fischer-Tropsch Fuels
from Biogas. Energy Convers. Manage. 2018, 174, 998−1013.
(53) Park, N.; Park, M.-J.; Ha, K.-S.; Lee, Y.-J.; Jun, K.-W. Modeling
and Analysis of a Methanol Synthesis Process Using a Mixed
Reforming Reactor: Perspective on Methanol Production and CO2
Utilization. Fuel 2014, 129, 163−172.
(54) Yoo, C.-J.; Lee, D.-W.; Kim, M.-S.; Moon, D. J.; Lee, K.-Y. The
Synthesis of Methanol from CO/CO2/H2 Gas over Cu/Ce1−xZrxO2
Catalysts. J. Mol. Catal. A: Chem. 2013, 378, 255−262.
(55) Goeppert, A.; Zhang, H.; Czaun, M.; May, R. B.; Prakash, G. K.
S.; Olah, G. A.; Narayanan, S. R. Easily Regenerable Solid Adsorbents
Based on Polyamines for Carbon Dioxide Capture from the Air.
ChemSusChem 2014, 7, 1386−1397.
(56) Goeppert, A.; Czaun, M.; May, R. B.; Prakash, G. K. S.; Olah,
G. A.; Narayanan, S. R. Carbon Dioxide Capture from the Air Using a
Polyamine Based. J. Am. Chem. Soc. 2011, 133, 20164−20167.
(57) Kothandaraman, J.; Goeppert, A.; Czaun, M.; Olah, G. A.;
Prakash, G. K. S. CO2 Capture by Amines in Aqueous Media and Its
Subsequent Conversion to Formate with Reusable Ruthenium and
Iron Catalysts. Green Chem. 2016, 18, 5831−5838.
(58) Goeppert, A.; Czaun, M.; Prakash, G. K. S.; Olah, G. A. Air as
the Renewable Carbon Source of the Future: An Overview of CO2
Capture from the Atmosphere. Energy Environ. Sci. 2012, 5, 7833−
7853.
(59) Shen, Y.; Linville, J. L.; Urgun-Demirtas, M.; Mintz, M. M.;
Snyder, S. W. An Overview of Biogas Production and Utilization at
Full-Scale Wastewater Treatment Plants (WWTPs) in the United
States: Challenges and Opportunities towards Energy-Neutral
WWTPs. Renewable Sustainable Energy Rev. 2015, 50, 346−362.
(60) Ashrafi, M.; Pfeifer, C.; Pröll, T.; Hofbauer, H. Experimental
Study of Model Biogas Catalytic Steam Reforming: 2. Impact of
Sulfur on the Deactivation and Regeneration of Ni-Based Catalysts.
Energy Fuels 2008, 22, 4190−4195.
(61) Appari, S.; Janardhanan, V. M.; Bauri, R.; Jayanti, S.;
Deutschmann, O. A Detailed Kinetic Model for Biogas Steam
Reforming on Ni and Catalyst Deactivation Due to Sulfur Poisoning.
Appl. Catal., A 2014, 471, 118−125.
(62) Mortensen, P. M.; Dybkjær, I. Industrial Scale Experience on
Steam Reforming of CO2-Rich Gas. Appl. Catal., A 2015, 495, 141−
151.
(63) Chein, R. Y.; Chen, Y. C.; Yu, C. T.; Chung, J. N.
Thermodynamic Analysis of Dry Reforming of CH4 with CO2 at
High Pressures. J. Nat. Gas Sci. Eng. 2015, 26, 617−629.
(64) Tassios, D. P. Applied Chemical Engineering Thermodynamics;
Springer, 1993.
(65) Vanden Bussche, K. M.; Froment, G. F. The STAR
Configuration for Methanol Synthesis in Reversed Flow Reactors.
Can. J. Chem. Eng. 1996, 74, 729−734.
(66) Bozzano, G.; Manenti, F. Efficient Methanol Synthesis:
Perspectives, Technologies and Optimization Strategies. Prog. Energy
Combust. Sci. 2016, 56, 71−105.
(67) Bussche, K. M. V.; Froment, G. F. A Steady-State Kinetic
Model for Methanol Synthesis and the Water Gas Shift Reaction on a
Commercial Cu/ZnO/Al2O3 Catalyst. J. Catal. 1996, 161, 1−10.
(68) Luyben, W. L. Design and Control of a Methanol Reactor/
Column Process. Ind. Eng. Chem. Res. 2010, 49, 6150−6163.
(69) Van-Dal, É . S.; Bouallou, C. Design and Simulation of a
Methanol Production Plant from CO2 Hydrogenation. J. Cleaner Prod.
2013, 57, 38−45.

J https://dx.doi.org/10.1021/acs.iecr.0c00755
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX

You might also like