You are on page 1of 18

ll

Perspective
Alcohol Production from Carbon Dioxide:
Methanol as a Fuel and Chemical Feedstock
Seda Sarp,1 Santiago Gonzalez Hernandez,1 Chi Chen,1 and Stafford W. Sheehan1,*

SUMMARY Context & Scale


Production of renewable alcohols from air, water, and sunlight pre- As scientists and engineers
sent an avenue to utilize captured carbon dioxide for the production studying CO2 conversion
of basic chemicals and store renewable energy in the chemical technologies, it is important to
bonds of liquid fuels. Of the technologies that utilize CO2 directly, understand the CO2-derived
CO2 electrolysis, as well as CO2 hydrogenation coupled with H2O methods of production for critical
electrolysis, have the benefit of requiring only CO2, H2O, and renew- basic chemicals, and renewable
able electricity as inputs with O2 as a sole byproduct. Among alco- alcohol production is one of the
hols, renewable methanol has seen the most development and anal- few CO2 utilization technologies
ysis in the chemical industry because it is currently a syngas-derived deployed at industrial scale.
product that could be adapted for direct CO2 utilization. In this Alcohols are building blocks for
perspective, we compare renewably powered CO2 electrolysis and materials that we encounter every
CO2 hydrogenation with the incumbent methanol production day; and among alcohols,
method from syngas from a cost and CO2 life cycle perspective by methanol is a critical chemical
analyzing recent literature to identify the research goals that enable intermediate for the production of
further scale-up. Survey of the industry shows that CO2 hydrogena- polymers, plastics, fibers, and
tion is among the closest CO2 utilization technologies to large-scale resins and holds promise as a
deployment. We further discuss these CO2 hydrogenation systems potential renewable liquid fuel. At
and the catalysts that drive them, with recommendations to drive present, most methanol is
further development and scale-up. produced using natural-gas-
derived syngas. Its alternative
INTRODUCTION production using CO2, water, and
renewable electricity presents an
The rate at which carbon dioxide (CO2) is being emitted into the atmosphere is out-
opportunity to advance entire
pacing its rate of absorption into terrestrial carbon sinks by a wider margin than ever
industries toward carbon
before in recorded history.1 By adopting sustainable manufacturing practices world-
neutrality. By analyzing its current
wide, including those that utilize CO2 to make liquid products that sequester more
stage of development, this
greenhouse gases (GHGs) than they emit, we can make a substantial contribution to
perspective presents research and
limiting the increase in global average temperature to 1.5 C above pre-industrial
development goals to further
levels. Limiting the increase in global average temperature was highlighted in the
utilize CO2 conversion to
Paris Agreement under the United Nations Framework Convention on Climate
methanol technologies. Working
Change2 and requires a drastic decrease in net anthropogenic CO2 emissions. Car-
together toward these goals, we
bon capture, utilization, and storage (CCUS) can play an essential role as society
can advance humanity toward the
moves toward achieving this goal, especially in the chemical industry.3,4
implementation of carbon-neutral
alcohols and low-carbon fuel.
Photosynthesis in microbes and plants is the original CCUS process that sequestered
CO2 from the atmosphere to form the fossil fuels we burn today.5 Thus, nature has
shown that combining CO2 and H2O using a renewable source of energy to produce
O2 and a carbon-containing product are compatible with global atmospheric chem-
istry. Processes that mimic photosynthesis in this manner are not yet widely de-
ployed, with a key missing link being the chemical or electrochemical reduction of
CO2.5,6 Thus far, laboratory research has shown several products that can be
made from CO2, H2O, and renewable electricity by either electrochemical reduction
of CO2, or chemical reduction of CO2 using H2 produced by H2O electrolysis.7

Joule 5, 59–76, January 20, 2021 ª 2020 Elsevier Inc. 59


ll
Perspective

However, none of these have yet been deployed on globally relevant scales because
they face challenging economics when compared with low-cost fossil fuels.8

Of the CO2 reduction technologies that have come the closest to reaching a global
scale, in that they have already been demonstrated on large (tons per day) scales,
three major C1 products stand out: carbon monoxide (CO), methane (CH4), and
methanol (CH3OH). CO can be produced from CO2 and H2O by direct CO2 electrol-
ysis9 or the reverse water-gas shift reaction (RWGS) coupled with commercial water
electrolysis. Of these approaches, most development has been in the nascent CO2
electrolysis industry with both membrane-based and solid oxide electrolyzers.10 CH4
is produced from CO2 in Sabatier reactors,11 a technology discovered in the late
19th century, with deployments over the last few decades driven using electrolytic
H2. Several examples of direct CO2 electrolysis for CH4 production have been re-
ported in laboratories, with limited progress to the ton per day scale. Finally, meth-
anol production using CO2 and electrochemically produced H2 has been demon-
strated on the 4,000 metric tons per annum (t/a) scale,12 which is one of the
largest and most mature CO2 utilization technology demonstrations. Production
of methanol from direct electrolysis is in earlier stages, with hurdles around catalyst
selectivity and current density.

Methanol has a unique advantage when compared with CO and CH4 in that it is a room-
temperature liquid with high energy density and CO2 sequestration potential (Table S1).
It also has a large demand as both a fuel and a basic feedstock chemical. This is exem-
plified by global methanol consumption, which was 64 million metric tons in 2015, and
of which 23 million metric tons were used for fuel-related applications.11 As methanol
and its derivatives are already well established in the global chemicals and transporta-
tion market, infrastructure and logistics do not present a significant challenge. More
than 90 methanol plants offer a combined maximum capacity of nearly 110 million
metric tons per year, generating $55 billion in economic activity.13 These plants use a
mixture of CO and H2 (syngas) as reagents. For a CO2-derived process to displace
this legacy one, it is imperative that researchers understand the current syngas process
and the ways that CO2 utilization technologies can improve upon it.

Syngas is obtained by reforming or partial oxidation of carbonaceous materials such as


coal, coke, natural gas, petroleum and heavy oils, or biomass. Economic considerations,
the long-term availability of raw materials, energy consumption, and environmental as-
pects determine the choice of raw material. Natural gas is currently the most widely
used feedstock, which is transformed into syngas by steam methane reforming (SMR,
Equation 1) and the water-gas shift (WGS) reaction to optimize CO:H2 ratios (Equation 2),
which is followed by methanol formation (Equation 3).14 Each of these reactions is revers-
ible and thus limited by thermodynamic equilibrium depending on the reaction condi-
tions, including temperature, pressure, and composition of the syngas.

kJ
CH4 + H2 O4CO + 3H2 DH298K = 205:43 : (Equation 1)
mol

kJ
CO + H2 O4CO2 + H2 DH298K =  41 : (Equation 2)
mol

kJ
CO + 2H2 4CH3 OH DH298K =  90:7 : (Equation 3)
mol 1Air
Company, 407 Johnson Avenue, Brooklyn,
Equations 2 and 3 are exothermic and cumulatively result in a decrease of volume as New York, NY 11206, USA

the reaction proceeds, thus the conversion into methanol is favored by increasing *Correspondence: staff@aircompany.com
https://doi.org/10.1016/j.joule.2020.11.005

60 Joule 5, 59–76, January 20, 2021


ll
Perspective

pressure and decreasing temperature.14 The WGS reaction results in a small amount
of CO2 production that is hydrogenated during methanol synthesis, making this a
process that already utilizes some CO2 though very minor compared with a direct
CO2-fed approach.15

To build viable systems that convert CO2 to methanol to effectively replace the pre-
sent natural-gas-fed systems, we must understand the differences between the two.
One major difference is the pathway for feedstock production. The existing method
uses SMR to produce syngas, which usually is carried out using nickel-based catalysts
at high temperatures of 800 C–1,000 C and pressures of 290–580 psi.16 Another is
reaction conditions; the production of methanol from syngas is much more
exothermic than from CO2 and H2, though the latter still occurs at moderate temper-
atures and pressures. Electrolysis offers a route that occurs at lower temperature and
pressure but as such faces challenges with production rate and cost of capital equip-
ment. These key differences, however, help to identify use cases where the CO2-
derived approaches bring additional value propositions.

Both electrolysis and direct hydrogenation of CO2 to methanol have been studied
widely in literature, toward improving catalysts as well as technoeconomic and life
cycle aspects. It has been shown that the CO2-based methanol production pro-
cesses can have lower global warming impacts compared with the fossil-based pro-
cess, as long as the electricity is provided renewably.17–25 In this Perspective article,
we discuss the ways that methanol produced from CO2, H2O, and renewable elec-
tricity can be more widely implemented and the research and development goals
that would help to achieve that. We outline the leading methods of production of
methanol from CO2 and assess literature technoeconomic analyses to identify
best practices. We then review examples of industrial commercial deployments to
disseminate the scale and status of the most successful deployments so that re-
searchers understand the current state of the industry. Finally, we discuss the appli-
cations that CO2-derived methanol can have to enable wider adoption.

HOW CAN WE PRODUCE METHANOL FROM CO2, H2O, AND


RENEWABLE ELECTRICITY?
Several pathways exist for methanol production in both the natural world and in the
chemical industry.26 Methanol is produced ubiquitously by fermentation with yeast,
necessitating distillation to remove it in the production of consumer alcohols.27
Fermentation with CO2 as a carbon source has also been described using specific
mutants of the Clostridium genus of Gram-positive bacteria.28 Some phytoplank-
tons produce methanol from algal biomass, making it a major source of methanol
in the marine euphotic zone.29 Direct photocatalytic pathways have also been stud-
ied.30 None of these, however, have attracted as much interest as direct CO2 elec-
trolysis or CO2 hydrogenation coupled with electrolysis. Table 1 gives a high-level
overview of the energy cost and life cycle GHG emissions of these processes
compared with the production of methanol from syngas, assuming all electricity
used comes from photovoltaics or a source with equivalent lifecycle GHG emissions
per kWh. It is worth noting that converting natural gas to methanol is an efficient (as
high as 75%) process, but the energy comes from fossil fuels and, therefore, has a
high carbon intensity.

Direct CO2 Electrolysis


Electrochemical reduction of CO2 has attracted substantial interest in the last
decade as a CO2 conversion pathway because of its potential for renewable

Joule 5, 59–76, January 20, 2021 61


ll
Perspective

Table 1. Comparison Showing Ranges of CO2 Emissions and Normalized Energy Use from CO2
Electrolysis, H2O Electrolysis Coupled with CO2 Electrolysis, and Conventional Methanol
Production in Idealized Cases
Process Subcomponent Energy Use Net GHG Emissions
(kWh/kg MeOH) (kg CO2e/kg MeOH)
CO2 electrolysis – 9.3–11.5 0.72 to 0.82
CO2 hydrogenation – 9.9–12.4 0.68 to 0.84
with H2O electrolysis
H2O electrolysis 9.6–11.1 +0.43 to +0.48
CO2 hydrogenation 0.3–1.3 1.11 to 1.32
Natural gas to methanol – 8.5–12.7 +0.77 to +1.6

electricity utilization and high overall energy efficiency.31 CO2 electrolysis proposes
efficient, on-site chemical production, provided reactor and catalyst combinations
with suitable selectivity, stability, overpotential, and capability to sustain commer-
cially relevant current densities are found.32 If CO2 electrolysis follows in the path
of water electrolysis for hydrogen production, the latter of which now achieves over-
all thermal efficiencies greater than 70% in modular systems,33 this potential may be
realized. Cathodic reduction of CO2 to saturated straight-chain alcohols, including
methanol, ethanol, and n-propanol requires six protons and six electrons provided
by water oxidation, as shown in Equation 4:

nCO2 + ð6nÞH + + ð6nÞe /Cn H2n + 1 OH + ð2n  1ÞH2 O: (Equation 4)


For the production of methanol (n = 1), there are still substantial hurdles in the two
major areas of research in CO2 electroreduction: (1) catalyst development, in that
there is no stable catalyst yet developed that can sustain both high current densities
and selectivity near 100% and (2) reactor design and engineering, which goes hand-
in-hand with catalyst development especially when dealing with multi-pass scenarios
where side products and unreacted CO2 are recycled through the electrolyzer.

The first challenge requires developing a catalyst that stably binds CO to its surface
yet retains lability to initiate four-electron, four-proton reduction to methanol. Since
this necessitates delicate balance in CO-binding energies, molecular catalysts have
shown promise toward driving this reaction.34 Oxide nanoparticles are also compe-
tent catalysts as a component in gas diffusion electrodes when comprisingCu2O/
ZnO.35 This is a notable development, as the same metal combination is used in
methanol synthesis by hydrogenation.

Although there has been strong progress developing flow reactors for CO2 electro-
reduction,36 the second challenge, reactor design, still requires significant develop-
ment to optimize system-level efficiency and mass transport.37 Among room-tem-
perature liquids that can be made by CO2 reduction, methanol has a unique
property that may facilitate improved efficiency at a systems level. The boiling point
of methanol at atmospheric pressure (65 C) is above room temperature but compa-
rable to the operational temperatures of a polymer electrolyte membrane (PEM)
electrolyzer. This presents a unique opportunity for efficient process integration,
as shown in Figure 1, by generating methanol in the gas phase while water (which
is typically found in the cathode outlet stream from the reduction reaction, humidi-
fied CO2 feed, electroosmosis, or a combination of these factors) is removed as a
liquid. This enables separation of methanol from gas and liquid recycle streams
with a simple glycol-cooled condenser and gas-liquid separator, reducing the
need for costlier distillation.

62 Joule 5, 59–76, January 20, 2021


ll
Perspective

O2
CO2
Feed Gas + Recycle
H2O + O2 + H2O
Reservoir
Electrolyte

H2O

CH3OH(l)

Cathode Flow Plate


Anode Flow Plate

Cathode GDE

ses
Anode GDE
Membrane

Recycle Ga

Distillation
H2O
CH3OH(l) + CO2 + H2O
H2O
o
Tcell < 65 C

B
O2
CO2
Feed Gas + Recycle
H2O + O2 + H2O
Reservoir
Electrolyte

H2O
Cathode Flow Plate
Anode Flow Plate

Condenser
Cathode GDE
Anode GDE
Membrane

CH3OH(l)
Gases

H2O
CH3OH(g) + CO2 + H2O
H2O
o
Tcell > 65 C

Figure 1. Two Zero-Gap System Configurations for CO2 Electrolysis to Methanol


(A) An electrolyzer that operates at a temperature below the boiling point of methanol with an
output mixture of water and methanol to be distilled to a pure product
(B) An electrolyzer that operates at a temperature above the boiling point of methanol (which is
more typical in PEM electrolysis), wherein methanol is collected as a vapor and condensed.

CO2 Hydrogenation Coupled with Water Electrolysis


Direct CO2 electroreduction is a potentially promising single-step route for meth-
anol production from CO2; however, CO2 hydrogenation combined with electrolysis
has proven to be the most scalable and implemented technology for renewable CO2
conversion to alcohols. Electrolysis of water occurs in standard commercial alkaline,
PEM, or solid oxide systems that are commercially available on the scale of several
megawatts,38 whereas exothermic hydrogenation of CO2 (Equation 5) occurs in
fixed-bed flow reactors that typically use a catalyst comprising copper oxide, zinc ox-
ide, and alumina (CZA) similar to that of the syngas process.

kJ
CO2 + 3H2 4CH3 OH + H2 O DH298K =  40:9 : (Equation 5)
mol
Experimental data have suggested that the legacy syngas process and CO2 hydro-
genation are both promoted by a formate intermediate, which is one explanation
why the syngas catalyst is also a competent catalyst for CO2 and H2.39

In model systems, as shown in Figure 2, captured CO2 and H2 from a water electro-
lyzer are compressed and preheated to up to 280 C to maintain an optimal thermal

Joule 5, 59–76, January 20, 2021 63


ll
Perspective

Figure 2. Diagram Showing the High-Level Components and Recycle Elements of an Integrated CO2 Hydrogenation Coupled with H2O Electrolysis
System

profile in the reactor. The reactor itself is loaded with the CZA catalyst that is pellet-
ized to ensure that it remains immobile and does not pulverize under the differential
pressure present during the reaction, while simultaneously optimizing the catalyst
mass to reactor volume ratio. In the reactor, Equation 5 and other side reactions
reach equilibrium producing methanol, water, and byproducts such as CO and
CH4. Unreacted CO2, H2, and other gaseous products are separated from liquid
products after exiting, and the liquids are distilled to separate methanol from water.
The unreacted CO2, H2, and product gases are, in many cases, recompressed and
recycled to the reactor. The water obtained from the distillation column can be
further recycled to the electrolyzer for hydrogen production to minimize the water
usage of the process. At a systems level, recycling of both water and gaseous prod-
ucts may increase system cost but are desirable to maximize CO2 utilization and
minimize life cycle GHG emissions.

Life Cycle and Technoeconomic Analysis


The primary advantage of CO2-based methanol is that it can offer lower overall GHG
emissions as compared with the legacy syngas process. In a cradle-to-gate analysis,
all steps of methanol production and their environmental impact are accounted for,
including parameters for the process steps shown in Figure 2 as well as several
external factors, including the method of CO2 capture, electricity generation, utiliza-
tion of side products, and more. To begin with the current syngas process as a com-
parison point, Aresta and coworkers14 conducted a life cycle analysis (LCA) on
various syngas to methanol routes and found that optimizing process engineering,
for example by efficiently capturing and recovering heat, can lower the overall en-
ergy required for methanol production by approximately 17%. However, even under

64 Joule 5, 59–76, January 20, 2021


ll
Perspective

A CO2 Electricity Cooling water Electricity


capture

CO2 Methanol Transporta on


CO2
Hydrogena on
Electrolysis H2O
H2O H2 (99.9 wt%)
( = 60-75%) 1-step in CO2 Electrolysis
End Use

O2

B
Natural Gas Electricity Cooling water Electricity

CH4 CO
Syngas Methanol synthesis
H2 Methanol Transporta on
produc on ( = 75%)
Steam CO2 (trace)
End use

Figure 3. Diagrams Showing Process Elements for Life Cycle Analysis of Methanol Production
(A) Direct CO 2 hydrogenation to methanol via water electrolysis.
(B) Methanol production via syngas from natural gas.

optimized conditions, their analysis showed that reacting recovered CO2 with H2
from water electrolysis powered by photovoltaics consumed approximately seven
times less energy than the conventional technology.

Several studies have been published investigating different process conditions for
the production of methanol from CO2 toward minimizing its CO2 equivalent GHG
emissions (CO2e); typical boundary conditions for these studies are shown in Fig-
ure 3, as compared with the legacy syngas method. As the majority of the global
warming impact (GWI) for the direct hydrogenation of CO2 to methanol is due to
hydrogen supply,23 the lowest global warming impacts are achieved if hydrogen is
supplied by water electrolysis using wind electricity.22,40 Solar-driven thermochem-
ical process combined with solar heat have also been studied and found to have
similar life cycle CO2 emissions.20

Different CO2 sources can have an effect on the GWI almost as substantial as the
electricity source for the system, which makes modular systems with flexible deploy-
ment parameters desirable. Although the conventional method maintains the high-
est GHG emissions, the GWI of CO2-based methanol can swing between carbon
positive and negative depending on carbon source and its heat and electricity re-
quirements.19,23,24 The energy cost of CO2 capture is such an important statistic
that one study showed even SMR-sourced H2 can provide a GWI that is nearly neutral
if an efficient CO2 source (such as biogas, industrial flue gas, or waste incineration) is
used together with wind power.23

Notably, the cost of CO2 is higher than that of natural gas, making technoeconomic
competition with the syngas process challenging independent of policy that favors

Joule 5, 59–76, January 20, 2021 65


ll
Perspective

A B
3.0 3.0

Normalized Cost ($/kg MeOH)


Normalized Cost ($/kg MeOH)

2.5 2.5

2.0 2.0

1.5 1.5

8 10 12 14 16 18 20 22 24 26 28 30 32 6 8 10 12 14 16 18 20 22 24 26 28 30 32

Total Conversion Energy (kWh/kg MeOH) Electrolysis Energy (kWh/kg MeOH)

Figure 4. Production Cost of Methanol Normalized for Cost of CO2, H2O, and Electricity between
Eleven Published Technoeconomic Analyses
(A) The total energy required for MeOH production.
(B) The energy required for electrolysis only.
Circles report references that use CO 2 hydrogenation coupled with various methods of water
electrolysis, and the red triangle reports a technoeconomic analysis for direct CO 2 electrolysis.
See Table S2.

CO2 removal. Electrochemical technoeconomic analyses assume a PEM or similar


electrolyzer,41 and hydrogenation-based approaches consider a commercial CZA
catalyst with temperatures varying between 200 C and 250 C with pressures around
750 psi. Multiple sources’ calculations showed that the cost of methanol produced
from CO2 is in the range of $1.25–$2.00 per kg MeOH, whereas for natural-gas-
derived methanol, the production cost can vary between values as low as $0.12–
$0.25 per kg MeOH (adjusted to 2019 cost using the Chemical Engineering Plant
Cost Index) as shown in Figure 4.42

To better understand and compare technoeconomic analyses performed for meth-


anol production from CO2, we performed an evaluation of 11 published studies
where we normalized feedstock costs to determine consensus target values for
the cost of MeOH production from CO2 (see Supplemental Information for
Methods). Figure 4 shows how the cost of CO2-based methanol production corre-
lated with both system-level and electrolysis energy requirements from these
studies.14,19–22,25,43–47 Figure 4A shows a linear relationship between the cost and
system-level energy required, with today’s cost of photovoltaic electricity that we
normalized to ($0.108 per kWh), as the slope. The relationship was expected
because many of the papers analyzed had approximately the same feed ratios,
with feed ratio deviation being the primary cause of outliers. For example, one
outlier in Figure 4A that used CO2 hydrogenation reported CO2/H2 feed ratios
that diverged significantly from the stoichiometric ratios.20 Figure 4B displays verti-
cal conformity between the electrolysis energy value from 9 to 12 kWh/kg MeOH,
due to the normalization of electricity cost and electrolysis energy requirements
for the CO2 hydrogenation processes. Interestingly, the minimum energy proposed
to be required for CO2 electrolysis also falls in this range, which emphasizes the po-
tential competitiveness of CO2 electrolysis once kinetic challenges (such as system
selectivity and overpotential at relevant current densities) are overcome.

The criteria for selecting the 11 CO2 to methanol papers analyzed was that the study
include or reasonably assess values for the CO2 feed, H2/H2O feed, and electricity
requirement for the overall system (Table S3). From there, each paper differed in in-
dividual parameters including reaction pathway, temperature, pressure, catalyst

66 Joule 5, 59–76, January 20, 2021


ll
Perspective

Table 2. Current State-of-the-Art Operational Conditions for CO2 Electrolysis, H2O Electrolysis, Combined with CO2 Hydrogenation, and Legacy
Methanol Production
Conversion Method Crude Methanol Reaction Rate Selectivity Per-Pass Demonstrated
Concentrationa Conversion Run Time
SOAb CO2 electrolysis 1%–5% 52 mA/cm2 at 2.7 Vcellc 44%d >10% >24 h
2
Electrolysis targets 25%–35% 500 mA/cm at 2.0 Vcell >90% 45%–55% >3,000 h
SOA CO2 hydrogenation >63% 0.7 g/mLcath at 10k GHSV >99.9% 35%–45% >3,000 h
Hydrogenation targets >63% 1.5 g/mLcath at 10k GHSV >99.9% 55%–65% >3,000 h
Syngas MeOH >84% 2.2 g/mLcath at 10k GHSV >99.4% >80% >3,000 h

Performance targets for CO2 electrolysis and CO2 hydrogenation research to become cost-competitive with the incumbent technology are outlined in cells in
second and fourth rows (Electrolysis targets and Hydrogenation targets, respectively).
a
Concentration of methanol in water or electrolyte, wt %.
b
SOA, state of the art
c
Assuming a 200 mV OER overpotential and 300 mV series resistance through the cell.
d
Best selectivity with high partial current density.

used, reactor type, etc. We found that the studies with the most realistic energy re-
quirements paid close attention to the energy cost of heating and cooling, which is
important in both electrolysis (where energy lost as overpotential is converted to
heat) and exothermic hydrogenation. Cooling for electrolysis (both H2 production
and CO2 electrolysis) results in low-grade waste heat, since the cell stacks to not op-
erate at higher than 80 C, while the heat emitted from hydrogenation is higher
grade and able to be reutilized.

The data extraction and method of cost calculation were outlined for each paper in
the Supplemental Information. The calculated cost range using today’s renewable
electricity prices was $1.32 per kg MeOH to $3.08 per kg MeOH, which is three to
ten times the cost of MeOH made from natural gas (Table S2). The reaction energy
requirement is an influential factor in the final cost, but not as much as divergence
from stoichiometric ratios of the CO2 and H2 feeds that leads to wasted H2. The
biggest hurdle for methanol made from CO2 and H2O that is preventing the process
from being economically competitive to traditional methanol production methods
today is the high energy requirement of electrolysis. Thus, any analysis is highly sen-
sitive to wasted H2 in the CO2 hydrogenation process.48 The first thing that must
happen to make CO2-derived methanol economically viable is either a decrease in
renewable electricity prices to one-third of their current level, a concomitant carbon
tax, or a combination of the two to drive this cost lower.

Once a low renewable electricity price is realized, the per-pass utilization of CO2 (current
density for electrolysis and space-time yield [STY] for hydrogenation) and durability of
integrated systems must be improved. Finally, the energy efficiency of the process, as
controlled by factors such as overpotential of electrocatalysts as well as reactor sizing
and heat loss, must be optimized. Based on our analysis of the 11 studies on technoe-
conomics for methanol production from CO2 and the current state-of-the-art sys-
tems,49–53 we determined performance targets for CO2 electrolysis and CO2 hydroge-
nation systems shown in Table 2. In determining these targets, we assumed that these
systems needed to reach a production cost below that of today’s commodity price,
$0.34/kg (see Supplemental Information). We also assumed that, in this future scenario,
renewable electricity prices lower to $0.03/kWh.

INDUSTRIAL DEMONSTRATIONS OF CO2 TO ALCOHOLS


Global efforts to scale and deploy CO2 utilization technologies have resulted in pi-
lot and demonstration-scale plants that use CO2 hydrogenation co-fed with H2

Joule 5, 59–76, January 20, 2021 67


ll
Perspective

Table 3. Operating or Recently Operated Abiotic CO2-to-Alcohol Pilot and Demonstration-Scale Production Plants Reported Using New or Newly
Scaled Technology
Organization Location Nameplate Capacity (tpa) Temperature ( C) Pressure (Psi)
a
Lurgi AG Frankfurt, Germany unpublished 260 870
NIRE and RITE Kyoto, Japan 18 250 725
CAMERE process Seoul, South Korea 73 250 400
Mitsui Chemicals Osaka, Japan 100 250 725
Carbon Recycling International Grindavik, Iceland 4,000 225 725
Air Company Brooklyn, NY, USA 32b 250 750
a
Lurgi AG CO2-derived methanol plant nameplate capacity has not been reported.
b
Air Company’s process is optimized for ethanol production and provides three separate product streams: MeOH (10 tpa), high-purity EtOH (13 tpa), and n-PrOH
(9 tpa).

provided by water electrolysis to produce alcohols. CO2 electrolysis is a potential


alternative in the long term, generally operating at lower temperatures and pres-
sures. However, CO2 electrolysis today is at a much lower technology-readiness
level (TRL), having not advanced past laboratory and prototype scales. Industrial
demonstrations of CO2 hydrogenation to methanol, on the other hand, have a his-
tory spanning the last 26 years, as summarized in Table 3. In 1994, Lurgi first
demonstrated the production of methanol from CO2 and H2 in a pilot plant using
a commercial Süd-Chemie CZA catalyst. The plant was able to reach 35%–45%
methanol yield at 260 C and 870 psi with a H2:CO2 ratio of 3.17,18 This was fol-
lowed by a 1996 demonstration from National Institute for Resources and Environ-
ment (NIRE) and Research Institute of Innovative Technology for the Earth (RITE) in
Japan in a similar process at the 18 tpa scale with as high as 99.7% methanol
selectivity.54

In 2009, Mitsui Chemicals in Korea opened a CO2 hydrogenation to methanol plant


that further scaled this technology to 100 tpa using CO2 captured post-ethylene pro-
duction and H2 generated by water electrolysis using with a CZA-SiO2-Pd/Ga/Zr
catalyst.55–57 Process conditions for this plant were followed by several in the future,
namely a H2:CO2 ratio of 3, operating temperature of 250 C at 725 psi pressure, and
a gas hourly space velocity (GHSV) of 1,000 h1. These operating parameters
achieved an output of 0.58 kg MeOH/LCat/h.57 By adding an additional off-the-shelf
RWGS reactor, Korean Institute of Science and Technologies developed the CA-
MERE (carbon dioxide hydrogenation to form methanol via a reverse-water-gas-shift
reaction) process, which demonstrated methanol production from CO2 and H2 by
first production CO over a zinc-alumina catalyst followed by methanol synthesis
with a Cu/ZnO/ZrO2/Ga2O3 catalyst.55,56,58,59

The MefCO2 plant in Niederaussem, Germany,43 and PetroChina pilot plant in coop-
eration with the Dalian Institute of Chemical Physics are more recent deployments
that further enable study of state-of-the-art catalyst and reactor technology on the
ton per day scale using captured CO2.56 Several upcoming deployments are
planned as well. Mitsubishi Gas Chemical Company is collaborating with Mitsubishi
Hitachi Power Systems (MHPS) and Mitsubishi Heavy Industries Engineering
(MHIENG) to utilize the CO2 emitted from the refinery in Tomakomai City, Hokkaido,
Japan, for a methanol synthesis plant that will have a 20 ton per day capacity, starting
operation in 2021.60 Sunfire in Germany partnered with Total Oil in 2019 to construct
a green methanol plant with the capacity of approximately 160 tpa that was planned
to start operation in 2021 that will utilize a wind-powered, 1-MW electrolyzer to pro-
vide the necessary H2 to react with captured CO2.56,61

68 Joule 5, 59–76, January 20, 2021


ll
Perspective

A 1.25 B 30

CO2 Converted to Methanol Per Pass (%)


25
1.00
280.0
Methanol STY (gMeOH/(gcath))

Temperature ( C)
20

o
0.75

15 235.0

0.50
10

0.25 190.0
5

0.00 0
0 5000 10000 15000 20000 25000 30000 35000 0 5000 10000 15000 20000 25000 30000 35000
-1 -1
GHSV (h ) GHSV (h )

Figure 5. Space-Time Yield and Per-Pass Conversion for CZA-Based Catalysts


(A) Space-time yield for CO2 to methanol at various GHSVs for CZA-based catalysts published in
literature between 1998 and 2020.
(B) Reported per-pass conversions for CO 2 to methanol from the same studies.
See Table S5.

Of all these organizations and deployments, Carbon Recycling International in Ice-


land has emerged as a global leader in CO2 to methanol with their George Olah
plant, with a reported capacity of 4,000 tpa, utilizing CO2 obtained from flue gas
at a geothermal plant and H2 obtained by water electrolysis using geothermal en-
ergy.12,62 Air Company also deployed a modular process that uses captured CO2
along with H2 produced by water electrolysis to produce high-purity renewable
ethanol for consumer goods, powered by solar photovoltaics. The process has sec-
ondary streams of methanol and n-propanol that can be used for fuel and fine chem-
ical applications. The process is being scaled to convert a flue gas slipstream from an
860-MW natural-gas-fired power plant in Alberta, Canada, for the final round of the
NRG COSIA Carbon XPrize.

In industry, conditions for CO2 hydrogenation to methanol are not dissimilar to


legacy methanol production, namely 725–1,450 psi of pressure and a temperature
range of 250 C–300 C.43,58,63–65 However, the conversion of CO2 to methanol is
thermodynamically limited and as such is more easily inhibited at lower tempera-
tures. On the other hand, an increase in temperature leads to the formation of un-
desirable CO and CH4,66 which promotes overconsumption of H2, leading to cata-
lyst sintering and deactivation.67 This puts more stringent temperature
requirements on the CO2-based process. Other liquid products, such as ethers, al-
cohols, and hydrocarbons can be formed in addition to methanol during CO2 hy-
drogenation, making a highly selective catalyst essential. Currently available com-
mercial CZA catalysts are frequently modified with promoters, such as Ga2O3,
ZrO2, and Cr2O3, as well as alkaline metals such as Na, K, Ca, and Mg,6,8,11,63–75
to affect Cu dispersion, particle size, and mobility that ultimately determine selec-
tivity and stability.

As in electrolysis, hydrogenation catalyst composition and preparation methods


have significant effects on the surface structure of the catalysts.68,73,76–80 When
studying these materials, the STY, or mass of methanol produced per mass of cata-
lyst per h, of catalysts in a CO2 hydrogenation reactor are linearly correlated with the
GHSV, which is the ratio of gas flow rate to the interior volume of the catalyst-packed
reactor. Figure 5 shows data from more than 40 published reports of CZA catalysts
for both variables. The STY linearly increases with increasing GHSV due to higher
supply of reactants to the reactor (Figure 5A). However, increasing GHSV might

Joule 5, 59–76, January 20, 2021 69


ll
Perspective

Table 4. High-Volume Chemicals Commonly Produced Using Methanol as a Synthetic Feedstock along with Market Sizes, Total Volume, and Product
Value
Compound Market Size (USD/Year) Global Production (tpa) Value Per
Metric Ton
(USD/t)
Dimethyl ether $9,700,000,000 (Bı̂ldea et al.101) 20,000,000 (Nakyai and Saebea102) $485
103 104
Methyl tert-butyl ether $14,900,000,000 (Markets and Markets ) 12,000,000 (Stefanakis ) $1,242
Acetic acid $8,700,000,000 (Naresh Kumar et al.105) 16,000,000 (Christodoulou and Velasquez- $544
Orta106)
Formaldehyde $18,000,000,000 (Hunt and Dale107) 52,000,000 (Martı́nez-Aquino et al.108) $346

Global production numbers include all production pathways (both from methanol and others).

lead to a decrease in CO2 conversion to methanol as the reactant gas retention time
is lowered (Figure 5B), leading to further loss of feedstock material and lower mass
efficiency unless gas recycle is employed at a system level.

Figure 5B also shows, generally, that lower GHSV increases CO2 converted per pass,
which makes sense as the gaseous reagents have a higher retention time at the cata-
lyst at low GHSV and illustrates how a CO2 hydrogenator can operate under varied
flow conditions. This may also be advantageous in scenarios where electricity power-
ing the H2 electrolyzer varies over time; when the feed H2 gas needs to be dialed
down, the overall GHSV can be decreased to reduce gas consumption and increase
plant carbon efficiency. Although higher temperature within the range of 190 C–
280 C is correlated with higher conversion, the per-pass conversion of CO2 to meth-
anol typically lower than 30% while on the system level (with feed gas recycle), the
overall conversion rates are much higher. In comparison, CO2 electrolysis has
reached comparable per-pass conversion rates for other products, suggesting po-
tential advancement of the earlier-stage technology to pilot- and demonstration-
scale systems. In these systems, the robustness of catalyst and support materials
to feed contaminants and reaction intermediates are of utmost importance, as
degradation pathways can be exacerbated by recycling feed gases for multi-pass
conversion.

Commercial CZA catalysts have been proven to have high activity and selectivity
under the conditions outlined above. However, Cu-based catalysts show poor ac-
tivity at low temperature81,82 and commonly face sintering due to thermal effects
at high temperatures, as well as undesirable water formation.83–86 In addition,
CZA catalysts are susceptible to poisoning by impurities commonly found in
flue gas.87 To address these drawbacks, the selectivity, stability, activity, and
durability of these catalysts can be further improved. Pd is a common addition
or substitution for CZA catalysts because Pd-based catalysts exhibit excellent ac-
tivity and selectivity for the hydrogenation of CO2 to methanol, which can be
further amplified when supported on cerium, lanthanum, zirconium, gallium, in-
dium, and other supports.63,82,84,88–99 However, as precious metals are costly
on the volumes required for thermal processes, these materials are not used
widely in industry.

WHAT CAN WE DO WITH METHANOL?


Methanol is a nearly ubiquitous feedstock material used for high-volume chemical
applications globally, with annual volumes and market sizes for major compounds
commonly derived from methanol, as shown in Table 4. Given the higher cost of pro-
ducing methanol from CO2, the most advantageous applications are those that

70 Joule 5, 59–76, January 20, 2021


ll
Perspective

derive highest value. Still, the ease of converting a gasoline combustion engine into
a methanol combustion engine leading to its flexibility as a drop-in fuel is under-
stated. In 2015, 36% of the global demand for methanol was for fuel applications.11
In addition to a robust infrastructure that already exists, using methanol as a fuel can
help reduce the emissions of undesirable toxic products such as unburned hydrocar-
bons, SOx, and NOx. Methanol also has a high octane number that enables efficient
engine performance, as well as a lower toxicity and faster biodegradability than gas-
oline, suggesting that leaks or spills would biodegrade more readily than traditional
petroleum fuels.100

Marine engines designed for methanol combustion can achieve engine efficiencies
similar to diesel combustion. Emissions from such engines meet or exceed current
regulations with no additional safety concerns relative to traditional marine
fuels.100 Dimethyl ether (DME) that is produced from methanol is used as pesti-
cide, polishing agent, and anti-rust agent.109–111 Methyl tert-butyl ether (MTBE)
is made from methanol and used in gasoline as an octane booster and
oxygenate,110,111 and larger oxymethylene ethers are potential drop-in replace-
ments for diesel fuel. MTBE can be further used to make high-purity isobutylene,
which is a reagent for production of butyl rubber, highly reactive polyisobutylene,
and methyl methacrylate, (MMA) which have applications across high-value con-
sumer goods.112 Tert-amyl methyl ether (TAME) has also been synthesized from
methanol, which helps with the complete combustion of fuel to minimize exhaust
gas emission such as NOx and SOx.113

As a feedstock chemical for a variety of goods, methanol is also a precursor to ace-


tic acid, which is used to produce vinyl acetate.114 Acetic acid is also a raw material
for cellulose acetate, a synthetic textile also used for photographic film, acetate es-
ters,115 acetic anhydride,116 and chloroacetic acid.117 Methanol is commonly con-
verted to olefins such as ethylene, propylene, and butylenes which can be reacted
to produce polyolefins, used to make many plastic materials.118,119 Modular pro-
duction of methanol integrated with downstream production of higher-value
chemicals such as these help to enable low-carbon, end-to-end manufacturing
on-site.

A less-frequently cited use of CO2 utilization technology is its application for human-
ity expanding to space and Mars. As a building block, methanol is a precursor to
formaldehyde, which can be used to create resins, sugars, and several other neces-
sities using fairly limited resources such as oxygen and water.120 These pathways
have been highlighted in the NASA CO2 Conversion Challenge, which has identified
phase 1 winners and phase 2 competitors for chemical feedstock production from
CO2 in space and on Mars.121 The currently known resources on Mars include large
quantities of water ice and CO2 in its atmosphere, providing carbon, hydrogen, and
oxygen as basic building blocks for life support, fuels and plastics and much else. In-
Situ Resource Utilization (ISRU) on Mars to produce building block alcohols, such as
methanol and ethanol, is enabled by the Martian atmosphere which is over 95% CO2
and can be extracted by cooling or compression.122

CONCLUSIONS
Several challenges and opportunities lie ahead for CO2 utilization technologies, and
ultimately for economic production and market adoption there are high-value
beachhead markets that CO2-to-alcohols technologies can target to enable further
scale-up and cost reduction. It is clear from the life cycle analyses outlined that

Joule 5, 59–76, January 20, 2021 71


ll
Perspective

production of methanol from CO2 is more advantageous for the environment than
the legacy method using natural gas and syngas. However, it is also evident from
the technoeconomic analyses surveyed that CO2-derived methanol is much more
expensive in large part due to the cost of electrochemical systems and renewable
electricity. To balance these two factors while implementing relevant technology,
the sustainable value propositions of CO2-derived methanol need to be effectively
captured to convince industry and government that a low-carbon alternative is worth
additional investment.

The most likely pathway for CO2-to-methanol implementation in the near term is
hydrogenation on CZA-based heterogeneous catalysts because of the technol-
ogy’s maturity and catalyst’s selectivity. However, there are still deployment sce-
narios where CO2 electrolysis to methanol may hold advantages in the long
term, namely modular deployment where economies of scale for heating and
cooling are challenging to achieve and where system integration with distillation
can provide unique energy-saving opportunities. There are no electrocatalysts yet
that can perform this reaction at commercially relevant rates; however, our
opinion is that this catalyst problem will be solved. In comparison, there has
been far more research progress with catalysts that perform CO2 electrolysis to
produce ethanol; however, the system-level barriers to ethanol production and
separation are higher. It is worth noting that several research groups study
CO2 electrocatalysis, but fewer study reactor design and systems integration.
Counterintuitively, CO2 electrolysis to ethanol may be further away because of
unaddressed reactor-level challenges despite more frequent reports of catalyst
improvement.

Both alcohols are important as chemical feedstocks and are among the highest
volume chemicals produced globally. We believe that ethanol is a better feedstock
for certain materials such as jet fuel and for consumer uses as an antiseptic, sol-
vent, and spirits, which we will discuss further in our future work. Methanol is
more useful to produce from CO2 than ethanol for heavy land and sea fuel and
as precursor to DME, acetic acid, oxymethylene ethers, and several other com-
pounds. Since methanol does not require formation of a C–C bond on a catalyst
surface, the kinetics are more easily accessible, as evidenced by routine selectivity
>99% among liquid-carbon-containing products with CO2 hydrogenation. Despite
the maturity of CO2 hydrogenation technology, there are still challenges in basic
research that require further study of catalysts to enable usage of low-cost and
contaminated CO2 streams, as well as reactor equilibria to better understand
gas recycle and optimal system integration. It is important to note that all alterna-
tive fuels face similar economic issues and scaling problems. Future deployments,
therefore, also rely on the cost of renewable energy decreasing. Ultimately, to be
able to compete with the production and cost of fossil-based processes, the CO2-
to-alcohols systems must effectively capture the value of their low life cycle carbon
emissions.

SUPPLEMENTAL INFORMATION
Supplemental Information can be found online at https://doi.org/10.1016/j.joule.
2020.11.005.

ACKNOWLEDGMENTS
The authors thank the New York State Energy Research and Development Authority
(NYSERDA) for their Clean Energy Internship Program (S.S. and S.G.H.).

72 Joule 5, 59–76, January 20, 2021


ll
Perspective

DECLARATION OF INTERESTS
The authors have submitted patents on carbon dioxide conversion technologies,
including patent application PCT/US2018/040442. The authors are employees of
Air Company, which has a financial interest in these patents.

REFERENCES
1. Solomon, S., Plattner, G.K., Knutti, R., and (LCA) applied to the synthesis of methanol. analysis of CO2-based methanol processes.
Friedlingstein, P. (2009). Irreversible climate Comparison of the use of syngas with the use Energy Environ. Sci. 12, 3425–3436.
change due to carbon dioxide emissions. of CO2 and dihydrogen produced from
Proc. Natl. Acad. Sci. USA 106, 1704–1709. renewables. In Environmental Challenges and 26. Tountas, A.A., Peng, X., Tavasoli, A.V.,
Greenhouse Gas Control for Fossil Fuel Duchesne, P.N., Dingle, T.L., Dong, Y.,
2. Rogelj, J., den Elzen, M., Höhne, N., Fransen, Utilization in the 21st Century, M.M. Maroto- Hurtado, L., Mohan, A., Sun, W., Ulmer, U.,
T., Fekete, H., Winkler, H., Schaeffer, R., Sha, Valer, C. Song, and Y. Soong, eds. (Springer), et al. (2019). Towards solar methanol: past,
F., Riahi, K., and Meinshausen, M. (2016). Paris pp. 331–347. present, and future. Adv Sci (Weinh) 6,
agreement climate proposals need a boost to 1801903.
keep warming well below 2  C. Nature 534, 15. Mikkelsen, M., Jørgensen, M., and Krebs, F.C.
631–639. (2010). The teraton challenge. A review of 27. Revilla, I., and González-SanJosé, M.L. (1998).
fixation and transformation of carbon dioxide. Methanol release during fermentation of red
3. Kätelhön, A., Meys, R., Deutz, S., Suh, S., and Energy Environ. Sci. 3, 43–81. grapes treated with pectolytic enzymes. Food
Bardow, A. (2019). Climate change mitigation Chem 63, 307–312.
potential of carbon capture and utilization in 16. Gangadharan, P., Kanchi, K.C., and Lou, H.H.
the chemical industry. Proc. Natl. Acad. Sci. (2012). Evaluation of the economic and 28. Tyurin, M., and Kiriukhin, M. (2013). Selective
USA 116, 11187–11194. environmental impact of combining dry methanol or formate production during
reforming with steam reforming of methane. continuous CO2 fermentation by the
4. Blanco, D.E., and Modestino, M.A. (2019). Chem. Eng. Res. Des. 90, 1956–1968. acetogen biocatalysts engineered via
Organic electrosynthesis for sustainable integration of synthetic pathways using Tn7-
chemical manufacturing. Trends in Chemistry 17. Artz, J., Müller, T.E., Thenert, K., Kleinekorte, tool. World J. Microbiol. Biotechnol. 29, 1611–
1, 8–10. J., Meys, R., Sternberg, A., Bardow, A., and 1623.
Leitner, W. (2018). Sustainable conversion of
5. Fukuzumi, S. (2015). Artificial photosynthetic carbon dioxide: an integrated review of 29. Mincer, T.J., and Aicher, A.C. (2016).
systems for production of hydrogen. Curr. Methanol production by a broad
catalysis and life cycle assessment. Chem.
Opin. Chem. Biol. 25, 18–26. Rev. 118, 434–504.
phylogenetic array of marine phytoplankton.
PLoS One 11, e0150820.
6. Sheehan, S.W., Cave, E.R., Kuhl, K.P., 18. Goeppert, A., Czaun, M., Jones, J.P., Surya
Flanders, N., Smeigh, A.L., and Co, D.T. 30. Wang, L., Ghoussoub, M., Wang, H., Shao, Y.,
Prakash, G.K., and Olah, G.A. (2014).
(2017). Commercializing solar fuels within Sun, W., Tountas, A.A., Wood, T.E., Li, H., Loh,
Recycling of carbon dioxide to methanol and
today’s markets. Chem 3, 3–7. derived products - closing the loop. Chem.
J.Y.Y., Dong, Y., et al. (2018). Photocatalytic
hydrogenation of carbon dioxide with high
7. Bushuyev, O.S., De Luna, P., Dinh, C.T., Tao, Soc. Rev. 43, 7995–8048.
selectivity to methanol at atmospheric
L., Saur, G., van de Lagemaat, J., Kelley, S.O.,
19. Hoppe, W., Thonemann, N., and Bringezu, S. pressure. Joule 2, 1369–1381.
and Sargent, E.H. (2018). What should we
(2018). Life cycle assessment of carbon 31. Chen, C., Khosrowabadi Kotyk, J.F.K., and
make with CO2 and how can we make it?
dioxide–based production of methane and
Joule 2, 825–832. Sheehan, S.W. (2018). Progress toward
methanol and derived polymers. J. Ind. Ecol. commercial application of electrochemical
8. Orella, M.J., Román-Leshkov, Y., and 22, 327–340. carbon dioxide reduction. Chem 4, 2571–
Brushett, F.R. (2018). Emerging opportunities 2586.
20. Kim, J., Henao, C.A., Johnson, T.A., Dedrick,
for electrochemical processing to enable
D.E., Miller, J.E., Stechel, E.B., and 32. Burdyny, T., and Smith, W.A. (2019). CO2
sustainable chemical manufacturing. Curr.
Maravelias, C.T. (2011). Methanol production
Opin. Chem. Eng. 20, 159–167. reduction on gas-diffusion electrodes and
from CO2 using solar-thermal energy: process why catalytic performance must be assessed
9. Smith, W.A., Burdyny, T., Vermaas, D.A., and development and techno-economic analysis. at commercially-relevant conditions. Energy
Geerlings, H. (2019). Pathways to industrial- Energy Environ. Sci. 4, 3122–3132. Environ. Sci. 12, 1442–1453.
scale fuel out of thin air from CO2 electrolysis.
Joule 3, 1822–1834. 21. Al-Kalbani, H., Xuan, J., Garcı́a, S., and Wang, 33. Ayers, K. (2019). The potential of proton
H. (2016). Comparative energetic assessment exchange membrane–based electrolysis
10. Küngas, R., Blennow, P., Heiredal-Clausen, T., of methanol production from CO2: chemical technology. Curr. Opin. Electrochem. 18,
Holt, T., Rass-Hansen, J., Primdahl, S., and versus electrochemical process. Appl. Energy 9–15.
Hansen, J.B. (2017). eCOs - a commercial CO2 165, 1–13.
electrolysis system developed by Haldor 34. Wu, Y., Jiang, Z., Lu, X., Liang, Y., and Wang,
Topsoe. ECS Trans 78, 2879–2884. 22. Matzen, M., and Demirel, Y. (2016). Methanol H. (2019). Domino electroreduction of CO2 to
and dimethyl ether from renewable hydrogen methanol on a molecular catalyst. Nature 575,
11. Younas, M., Loong Kong, L., Bashir, M.J.K., and carbon dioxide: alternative fuels 639–642.
Nadeem, H., Shehzad, A., and Sethupathi, S. production and life-cycle assessment.
(2016). Recent advancements, fundamental J. Clean. Prod. 139, 1068–1077. 35. Albo, J., and Irabien, A. (2016). Cu2O-loaded
challenges, and opportunities in catalytic gas diffusion electrodes for the continuous
methanation of CO2. Energy Fuels 30, 8815– 23. Sternberg, A., Jens, C.M., and Bardow, A. electrochemical reduction of CO2 to
8831. (2017). Life cycle assessment of CO2-based methanol. J. Catal. 343, 232–239.
C1-chemicals. Green Chem 19, 2244–2259.
12. Marlin, D.S., Sarron, E., and Sigurbjörnsson, 36. Jeng, E., and Jiao, F. (2020). Investigation of
Ó. (2018). Process advantages of direct CO2 24. von der Assen, N., Jung, J., and Bardow, A. CO2 single-pass conversion in a flow
to methanol synthesis. Front. Chem. 6, 446. (2013). Life-cycle assessment of carbon electrolyzer. React. Chem. Eng. 5, 1768–1775.
dioxide capture and utilization: avoiding the
13. Methanol Institute (2020). The methanol pitfalls. Energy Environ. Sci. 6, 2721–2734. 37. Angulo, A., van der Linde, P., Gardeniers, H.,
industry. https://www.methanol.org/. Modestino, M., and Fernández Rivas, D.
25. González-Garay, A., Frei, M.S., Al-Qahtani, A., (2020). Influence of bubbles on the energy
14. Aresta, M., Caroppo, A., Dibenedetto, A., and Mondelli, C., Guillén-Gosálbez, G., and conversion efficiency of electrochemical
Narracci, M. (2002). Life Cycle Assessment Pérez-Ramı́rez, J. (2019). Plant-to-planet reactors. Joule 4, 555–579.

Joule 5, 59–76, January 20, 2021 73


ll
Perspective

38. Danilovic, N., Ayers, K.E., Capuano, C., decorated carbon membranes for CO2 day class test plant for methanol synthesis
Renner, J.N., Wiles, L., and Pertoso, M. (2016). electroreduction to methanol. J. Am. Chem. from CO2 and H2. Stud. Surf. Sci. Catal. 114,
(Plenary) challenges in going from laboratory Soc. 141, 12717–12723. 357–362.
to megawatt scale PEM electrolysis. ECS
Trans 75, 395–402. 52. Liu, Y., Li, F., Zhang, X., and Ji, X. (2020). 65. Meunier, N., Chauvy, R., Mouhoubi, S.,
Recent progress on electrochemical Thomas, D., and De Weireld, G. (2020).
39. Kuld, S., Thorhauge, M., Falsig, H., Elkjær, reduction of CO2 to methanol. Curr. Opin. Alternative production of methanol from
C.F., Helveg, S., Chorkendorff, I., and Green Sustain. Chem. 23, 10–17. industrial CO2. Renew. Energy 146, 1192–
Sehested, J. (2016). Quantifying the 1203.
promotion of Cu catalysts by ZnO for 53. Payra, S., Shenoy, S., Chakraborty, C.,
methanol synthesis. Science 352, 969–974. Tarafder, K., and Roy, S. (2020). Structure- 66. Jadhav, S.G., Vaidya, P.D., Bhanage, B.M.,
sensitive electrocatalytic reduction of CO2 to and Joshi, J.B. (2014). Catalytic carbon
40. Pérez-Fortes, M., Schöneberger, J.C., methanol over carbon-supported dioxide hydrogenation to methanol: a review
Boulamanti, A., and Tzimas, E. (2016). intermetallic PtZn nano-alloys. ACS Appl. of recent studies. Chem. Eng. Res. Des. 92,
Methanol synthesis using captured CO2 as Mater. Interfaces 12, 19402–19414. 2557–2567.
raw material: techno-economic and
environmental assessment. Appl. Energy 161, 54. Toyir, J., Miloua, R., Elkadri, N.E., Nawdali, M., 67. Wu, J., Saito, M., Takeuchi, M., and
718–732. Toufik, H., Miloua, F., and Saito, M. (2009). Watanabe, T. (2001). The stability of Cu/ZnO-
Sustainable process for the production of based catalysts in methanol synthesis from a
41. Orella, M.J., Brown, S.M., Leonard, M.E., methanol from CO2 and H2 using Cu/ZnO- CO2-rich feed and from a CO-rich feed. Appl.
Román-Leshkov, Y., and Brushett, F.R. (2020). based multicomponent catalyst. Phys. Catal. A 218, 235–240.
A general technoeconomic model for Procedia 2, 1075–1079.
evaluating emerging electrolytic processes. 68. Bonura, G., Khassin, A.A., Yurieva, T.M.,
Energy Technol 8, 1900994. 55. Simon Araya, S.S., Liso, V., Cui, X., Li, N., Zhu, Cannilla, C., Frusteri, F., and Frusteri, L. (2020).
J., Sahlin, S.L., Jensen, S.H., Nielsen, M.P., Structure control on kinetics of copper
42. Galindo Cifre, P.G., and Badr, O. (2007). and Kær, S.K. (2020). A review of the methanol reduction in Zr–containing mixed oxides
Renewable hydrogen utilisation for the economy: the fuel cell route. Energies 13, 596. during catalytic hydrogenation of carbon
production of methanol. Energy Convers. oxides to methanol. Catal. Today 342, 39–45.
Manag. 48, 519–527. 56. Zhong, J., Yang, X., Wu, Z., Liang, B., Huang,
Y., and Zhang, T. (2020). State of the art and 69. Zhan, H., Li, F., Gao, P., Zhao, N., Xiao, F., Wei,
43. Bowker, M. (2019). Methanol synthesis from perspectives in heterogeneous catalysis of W., Zhong, L., and Sun, Y. (2014). Methanol
CO2 hydrogenation. ChemCatChem 11, CO2 hydrogenation to methanol. Chem. Soc. synthesis from CO2 hydrogenation over La-M-
4238–4246. Rev. 49, 1325–1616. Cu-Zn-O (M = Y, Ce, Mg, Zr) catalysts derived
from perovskite-type precursors. J. Power
44. Asif, M., Gao, X., Lv, H., Xi, X., and Dong, P. 57. Matsushita, T., Haganuma, T., and Fujita, D. Sources 251, 113–121.
(2018). Catalytic hydrogenation of CO2 from (2013). Process for producing methanol.
600 MW supercritical coal power plant to European patent 2,641,889B1, filed 70. Angelo, L., Kobl, K., Tejada, L.M.M.,
produce methanol: a techno-economic November 18, 2011, and granted December Zimmermann, Y., Parkhomenko, K., and
analysis. Int. J. Hydr. Energy 43, 2726–2741. 20, 2017. Roger, A.C. (2015). Study of CuZnMOx oxides
(M = Al, Zr, Ce, CeZr) for the catalytic
45. Tremel, A., Wasserscheid, P., Baldauf, M., and 58. Roy, S., Cherevotan, A., and Peter, S.C. (2018). hydrogenation of CO2 into methanol. C. R.
Hammer, T. (2015). Techno-economic analysis Thermochemical CO2 hydrogenation to Chim. 18, 250–260.
for the synthesis of liquid and gaseous fuels single carbon products: scientific and
based on hydrogen production via technological challenges. ACS Energy Lett 3, 71. Meliancabrera, I., Granados, M.L., and Fierro,
electrolysis. Int. J. Hydr. Energy 40, 11457– 1938–1966. J.L.G. (2002). Reverse topotactic
11464. transformation of a Cu–Zn–Al catalyst during
59. Joo, O.S., Jung, K.D., Moon, I., Rozovskii, wet Pd impregnation: relevance for the
46. Zhang, H., Wang, L., Van Herle, J., Maréchal, A.Y., Lin, G.I., Han, S.H., and Uhm, S.J. (1999). performance in methanol synthesis from CO2/
F., and Desideri, U. (2019). Techno-economic Carbon dioxide hydrogenation to form H2 mixtures. J. Catal. 210, 273–284.
optimization of CO2-to-methanol with solid- methanol via a reverse-water-gas-shift
oxide electrolyzer. Energies 12, 3742. reaction (the CAMERE process). Ind. Eng. 72. Saito, M., and Murata, K. (2004). Development
Chem. Res. 38, 1808–1812. of high performance Cu/ZnO-based catalysts
47. Jouny, M., Luc, W., and Jiao, F. (2018). for methanol synthesis and the water-gas shift
General techno-economic analysis of CO2 60. Oil and Gas 360 Feed Wire (2020). MHPS, reaction. Catal. Surv. Asia 8, 285–294.
electrolysis systems. Ind. Eng. Chem. Res. 57, MHIENG and MGC selected to conduct
2165–2177. research on effective recycling of CO2 to 73. Chen, D., Mao, D., Wang, G., Guo, X., and Yu,
produce methanol. https://www. J. (2019). CO2 hydrogenation to methanol
48. Ardo, S., Fernandez Rivas, D., Modestino, oilandgas360.com/mhps-mhieng-and-mgc- over CuO-ZnO-ZrO2 catalyst prepared by
M.A., Schulze Greiving, V., Abdi, F.F., Alarcon selected-to-conduct-research-on-effective- polymeric precursor method. J Sol-Gel Sci
Llado, E., Artero, V., Ayers, K., Battaglia, C., recycling-of-co2-to-produce-methanol/. Technol 89, 686–699.
Becker, J.-P., et al. (2018). Pathways to
electrochemical solar-hydrogen 61. Reuters. (2019). Germany’s Sunfire partners 74. Wang, W., Qu, Z., Song, L., and Fu, Q. (2020).
technologies. Energy Environ. Sci. 11, 2768– with Total to produce hydrogen fuels at An investigation of Zr/Ce ratio influencing the
2783. refineries. https://www.reuters.com/article/ catalytic performance of CuO/Ce1-xZrxO2
total-sunfire-hydrogen-idUSL5N26N4BY. catalyst for CO2 hydrogenation to CH3OH.
49. Dieterich, V., Buttler, A., Hanel, A., Spliethoff, J. Energy Chem. 47, 18–28.
H., and Fendt, S. (2020). Power-to-liquid via 62. Shulenberger, A.M., Jonsson, F.R.,
synthesis of methanol, DME or Fischer– Ingolfsson, O., and Tran, K.C. (2007). Process 75. Ren, S., Fan, X., Shang, Z., Shoemaker, W.R.,
Tropsch-fuels: a review. Energy Environ. Sci. for producing liquid fuel from carbon dioxide Ma, L., Wu, T., Li, S., Klinghoffer, N.B., Yu, M.,
13, 3207–3252. and water. US patent 8,198,338B2, filed March and Liang, X. (2020). Enhanced catalytic
20, 2007, and granted June 12, 2012. performance of Zr modified CuO/ZnO/Al2O3
50. Lunkenbein, T., Girgsdies, F., Kandemir, T., catalyst for methanol and DME synthesis via
Thomas, N., Behrens, M., Schlögl, R., and Frei, 63. Chen, C., Kotyk, J.F.K., and Sheehan, S.W. CO2 hydrogenation. J. CO2 Util. 36, 82–95.
E. (2016). Bridging the time gap: a copper/ (2018). Methods and catalysts for the selective
zinc oxide/aluminum oxide catalyst for production of methanol from carbon dioxide 76. Liu, J., Shi, J., He, D., Zhang, Q., Wu, X., Liang,
methanol synthesis studied under industrially and hydrogen gas for chemical synthesis and Y., and Zhu, Q. (2001). Surface active structure
relevant conditions and time scales. Angew. gas purification. US patent 10,858,302, filed of ultra-fine Cu/ZrO2 catalysts used for the
Chem. Int. Ed. Engl. 55, 12708–12712. June 19, 2018, and granted December 8, CO2+H2 to methanol reaction. Appl. Catal. A
2020. 218, 113–119.
51. Yang, H., Wu, Y., Li, G., Lin, Q., Hu, Q., Zhang,
Q., Liu, J., and He, C. (2019). Scalable 64. Ushikoshi, K., Moria, K., Watanabe, T., 77. Yang, C., Ma, Z., Zhao, N., Wei, W., Hu, T., and
production of efficient single-atom copper Takeuchi, M., and Saito, M. (1998). A 50 kg/ Sun, Y. (2006). Methanol synthesis from CO2-

74 Joule 5, 59–76, January 20, 2021


ll
Perspective

rich syngas over a ZrO2 doped CuZnO W. (2018). Solvent free synthesis of PdZn/TiO2 America, Europe, APAC, Latin America,
catalyst. Catal. Today 115, 222–227. catalysts for the hydrogenation of CO2 to Middle East & Africa) - Global Forecast to
methanol. Top. Catal. 61, 144–153. 2022. https://www.marketsandmarkets.com/
78. Guo, X., Mao, D., Lu, G., Wang, S., and Wu, G. Market-Reports/methyl-tertiary-butyl-ether-
(2011). CO2 hydrogenation to methanol over 91. Malik, A.S., Zaman, S.F., Al-Zahrani, A.A., market-12328028.html.
Cu/ZnO/ZrO2 catalysts prepared via a route Daous, M.A., Driss, H., and Petrov, L.A. (2018).
of solid-state reaction. Catal. Commun. 12, Development of highly selective PdZn/CeO2 104. Stefanakis, A.I. (2020). The fate of MTBE and
1095–1098. and Ca-doped PdZn/CeO2 catalysts for BTEX in constructed wetlands. Appl. Sci. 10,
methanol synthesis from CO2 hydrogenation. 127.
79. Guo, X., Mao, D., Wang, S., Wu, G., and Lu, G. Appl. Catal. A 560, 42–53.
(2009). Combustion synthesis of CuO–ZnO– 105. Naresh Kumar, A., Bandarapu, A.K., and
ZrO2 catalysts for the hydrogenation of 92. Ojelade, O.A., Zaman, S.F., Daous, M.A., Al- Venkata Mohan, S. (2019). Microbial electro-
carbon dioxide to methanol. Catal. Commun. Zahrani, A.A., Malik, A.S., Driss, H., Shterk, G., hydrolysis of sewage sludge for acidogenic
10, 1661–1664. and Gascon, J. (2019). Optimizing Pd:Zn molar production of biohydrogen and volatile fatty
ratio in PdZn/CeO2 for CO2 hydrogenation to acids along with struvite. Chem. Eng. J. 374,
80. Arena, F., Barbera, K., Italiano, G., Bonura, G., methanol. Appl. Catal. A 584, 117185. 1264–1274.
Spadaro, L., and Frusteri, F. (2007). Synthesis,
characterization and activity pattern of Cu– 93. Dı́ez-Ramı́rez, J., Sánchez, P., Rodrı́guez- 106. Christodoulou, X., and Velasquez-Orta, S.B.
ZnO/ZrO2 catalysts in the hydrogenation of Gómez, A., Valverde, J.L., and Dorado, F. (2016). Microbial electrosynthesis and anaerobic
carbon dioxide to methanol. J. Catal. 249, (2016). Carbon nanofiber-based palladium/ fermentation: an economic evaluation for acetic
185–194. zinc catalysts for the hydrogenation of carbon acid production from CO2 and CO. Environ. Sci.
dioxide to methanol at atmospheric pressure. Technol. 50, 11234–11242.
81. Fujitani, T., Saito, M., Kanai, Y., Watanabe, T., Ind. Eng. Chem. Res. 55, 3556–3567.
Nakamura, J., and Uchijima, T. (1995). 107. Hunt, A., and Dale, N. (2018). Economic
Development of an active Ga2O3 supported 94. Dı́ez-Ramı́rez, J., Valverde, J.L., Sánchez, P., Valuation in Formaldehyde Regulation,
palladium catalyst for the synthesis of and Dorado, F. (2016). CO2 hydrogenation to OECD environment working papers, no. 134.
methanol from carbon dioxide and hydrogen. methanol at atmospheric pressure: influence https://www.oecd-ilibrary.org/docserver/
Appl. Catal. A 125, L199–L202. of the preparation method of Pd/ZnO c0f8b8e5-en.pdf?expires=1605469661&
catalysts. Catal. Lett. 146, 373–382. id=id&accname=guest&checksum=
82. Garcı́a-Trenco, A., Regoutz, A., White, E.R., 2917C34AFE7BAB4BD4B4D3F1BD6E0846.
Payne, D.J., Shaffer, M.S.P., and Williams, C.K. 95. Baiker, A., and Gasser, D. (1989). Supported
(2018). PdIn intermetallic nanoparticles for the palladium catalyst prepared from amorphous 108. Martı́nez-Aquino, C., Costero, A.M., Gil, S.,
hydrogenation of CO2 to methanol. Appl. palladium–zirconium. Structural properties and Gaviña, P. (2018). A new environmentally-
Catal. 220, 9–18. and catalytic behaviour in the hydrogenation friendly colorimetric probe for formaldehyde
of carbon dioxide. J. Chem. Soc. Faraday gas detection under real conditions.
83. Iwasa, N., Suzuki, H., Terashita, M., Arai, M., Trans. 1, 999–1007. Molecules 23, 2646.
and Takezawa, N. (2004). Methanol synthesis
from CO2 under atmospheric pressure over 96. Jiang, X., Nie, X., Wang, X., Wang, H., 109. Azizi, Z., Rezaeimanesh, M., Tohidian, T., and
supported Pd catalysts. Catal. Lett. 96, 75–78. Koizumi, N., Chen, Y., Guo, X., and Song, C. Rahimpour, M.R. (2014). Dimethyl ether: a
(2019). Origin of Pd-Cu bimetallic effect for review of technologies and production
84. Liao, F., Wu, X.P., Zheng, J., Li, M.M.-J., synergetic promotion of methanol formation challenges. Chem. Eng. Process. 82,
Kroner, A., Zeng, Z., Hong, X., Yuan, Y., Gong, from CO2 hydrogenation. J. Catal. 369, 21–32. 150–172.
X.-Q., and Tsang, S.C.E. (2017). A promising
low pressure methanol synthesis route from 97. Song, J., Liu, S., Yang, C., Wang, G., Tian, H., 110. Awad, O.I., Mamat, R., Ali, O.M., Sidik, N.A.C.,
CO2 hydrogenation over Pd@Zn core–shell Zhao, Z.J., Mu, R., and Gong, J. (2020). The Yusaf, T., Kadirgama, K., and Kettner, M.
catalysts. Green Chem 19, 270–280. role of Al doping in Pd/ZnO catalyst for CO2 (2018). Alcohol and ether as alternative fuels in
hydrogenation to methanol. Appl. Catal. 263, spark ignition engine: a review. Renew.
85. Behrens, M., Zander, S., Kurr, P., Jacobsen, N., 118367. Sustain. Energ. Rev. 82, 2586–2605.
Senker, J., Koch, G., Ressler, T., Fischer, R.W.,
and Schlögl, R. (2013). Performance 98. Sharma, S., Hilaire, S., Vohs, J.M., Gorte, R.J., 111. Sezer, I., and Bilgin, A. (2008). Effects of
improvement of nanocatalysts by promoter- and Jen, H.-W. (2000). Evidence for oxidation methyl tert-butyl ether addition to base
induced defects in the support material: of ceria by CO2. J. Catal. 190, 199–204. gasoline on the performance and CO
methanol synthesis over Cu/ZnO:Al. J. Am. emissions of a spark ignition engine. Energy
Chem. Soc. 135, 6061–6068. 99. Matsumura, Y., Shen, W.J., Ichihashi, Y., and Fuels 22, 1341–1348.
Okumura, M. (2001). Low-temperature
86. Bukhtiyarova, M., Lunkenbein, T., Kähler, K., methanol synthesis catalyzed over ultrafine 112. Ryu, J.Y. (2008). Process and catalyst for
and Schlögl, R. (2017). Methanol synthesis palladium particles supported on cerium cracking of ethers and alcohols, US patent
from industrial CO2 sources: a contribution to oxide. J. Catal. 197, 267–272. 8,829,260B2, filed April 19, 2012, and granted
chemical energy conversion. Catal. Lett. 147, September 9, 2014,
416–427. 100. Corbett, J.J., and Winebrake, J.J. (2018). Life
cycle analysis of the usage of methanol for 113. Rechnia, P., Malaika, A., and Kozłowski, M.
87. Fujitani, T., and Nakamura, I. (2002). Methanol marine transportation (U.S. Department of (2015). Synthesis of tert-amyl methyl ether
synthesis from CO and CO2 hydrogenations Transportation Maritime Administration). (TAME) over modified activated carbon
over supported palladium catalysts. Bull. https://www.maritime.dot.gov/innovation/ catalysts. Fuel 154, 338–345.
Chem. Soc. 75, 1393–1398. meta/lifecycle-analysis-use-methanol-marine-
transportation. 114. Johnston, V.J., Zink, J.H., Repman, D.R.,
88. Xu, J., Su, X., Liu, X., Pan, X., Pei, G., Huang, Y., Kimmich, B.F., Chapman, J.T., and Chen, L.
Wang, X., Zhang, T., and Geng, H. (2016). 101. Bı̂ldea, C.S., Gy}
orgy, R., Brunchi, C.C., and (2010). Integrated process for the production of
Methanol synthesis from CO2 and H2 over Pd/ Kiss, A.A. (2017). Optimal design of intensified vinylacetate from acetic acid via ethylacetate.
ZnO/Al2O3: catalyst structure dependence of processes for DME synthesis. Comput. Chem. US patent 8,178,715B2, filed December 31,
methanol selectivity. Appl. Catal. A 514, Eng. 105, 142–151. 2008, and granted May 15, 2012.
51–59.
102. Nakyai, T., and Saebea, D. (2019). 115. Mooberry, J.B., Gompf, T.E., and Johnston,
89. Bahruji, H., Bowker, M., Hutchings, G., Exergoeconomic comparison of syngas B.H. (1999). Acetate ester compounds. US
Dimitratos, N., Wells, P., Gibson, E., Jones, production from biomass, coal, and natural gas patent 5,959,120A, filed June 12, 1997, and
W., Brookes, C., Morgan, D., and Lalev, G. for dimethyl ether synthesis in single-step and granted September 28, 1999.
(2016). Pd/ZnO catalysts for direct CO2 two-step processes. J. Clean. Prod. 241, 118334.
hydrogenation to methanol. J. Catal. 343, 116. Scates, M.O., and Webb, S.C. (2011).
133–146. 103. Markets and Markets (2020). Methyl Tertiary Integrated process for producing
Butyl Ether Market by Manufacturing Process, carbonylation acetic acid, acetic anhydride, or
90. Bahruji, H., Esquius, J.R., Bowker, M., (Steam Cracker, Fluid Liquid Cracker), coproduction of each from a methyl acetate
Hutchings, G., Armstrong, R.D., and Jones, Application (Gasoline), and Region (North by-product stream. US patent

Joule 5, 59–76, January 20, 2021 75


ll
Perspective

US20080221352A1, filed August 23, 2006, and 119. Lesthaeghe, D., Van Speybroeck, V., Marin, 121. Rowe, J. (2019). Winning teams design
granted September 11, 2008. G.B., and Waroquier, M. (2007). The rise and systems to convert carbon dioxide
fall of direct mechanisms in methanol-to- into something sweet (National Astronautics
117. Bankowski, B. (2019). Method of industrially olefin catalysis: an overview of theoretical and Space Administration). https://www.
producing monochloroacetic acid. US patent contributions. Ind. Eng. Chem. Res. 46, 8832– nasa.gov/spacetech/centennial_challenges/
10 494,325B2, filed February 4, 2017, and 8838. co2challenge/winning-teams-design-
granted December 3, 2019. systems-to-convert-carbon-dioxide-into-
120. Heim, L.E., Konnerth, H., and Prechtl, something-sweet.html.
118. Tian, P., Wei, Y., Ye, M., and Liu, Z. (2015). M.H.G. (2017). Future perspectives
Methanol to olefins (MTO): from for formaldehyde: pathways for 122. Starr, S.O., and Muscatello, A.C. (2020).
fundamentals to commercialization. ACS reductive synthesis and energy storage. Mars in situ resource utilization: a review.
Catal 5, 1922–1938. Green Chem 19, 2347–2355. Planet. Space Sci. 182, 104824.

76 Joule 5, 59–76, January 20, 2021

You might also like