You are on page 1of 66

pubs.acs.

org/CR Review

Understanding the Effect of Water on CO2 Adsorption


Joel M. Kolle, Mohammadreza Fayaz, and Abdelhamid Sayari*

Cite This: https://doi.org/10.1021/acs.chemrev.0c00762 Read Online

ACCESS Metrics & More Article Recommendations

ABSTRACT: Carbon capture from large sources and ambient air is one of the most
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

promising strategies to curb the deleterious effect of greenhouse gases. Among different
technologies, CO2 adsorption has drawn widespread attention mostly because of its low
energy requirements. Considering that water vapor is a ubiquitous component in air and
almost all CO2-rich industrial gas streams, understanding its impact on CO2 adsorption is of
Downloaded via UNIV OF ULSTER on May 17, 2021 at 16:39:11 (UTC).

critical importance. Owing to the large diversity of adsorbents, water plays many different
roles from a severe inhibitor of CO2 adsorption to an excellent promoter. Water may also
increase the rate of CO2 capture or have the opposite effect. In the presence of amine-
containing adsorbents, water is even necessary for their long-term stability. The current
contribution is a comprehensive review of the effects of water whether in the gas feed or as
adsorbent moisture on CO2 adsorption. For convenience, we discuss the effect of water
vapor on CO2 adsorption over four broadly defined groups of materials separately, namely (i) physical adsorbents, including carbons,
zeolites and MOFs, (ii) amine-functionalized adsorbents, and (iii) reactive adsorbents, including metal carbonates and oxides. For
each category, the effects of humidity level on CO2 uptake, selectivity, and adsorption kinetics under different operational conditions
are discussed. Whenever possible, findings from different sources are compared, paying particular attention to both similarities and
inconsistencies. For completeness, the effect of water on membrane CO2 separation is also discussed, albeit briefly.

CONTENTS 3.4.5. Effect of Framework Flexibility S


3.4.6. Comparative Studies S
1. Introduction B 4. Amine-Functionalized Adsorbents V
2. Measuring CO2 Uptake Under Humid Conditions C 4.1. Amine-Impregnated Adsorbents Y
2.1. Thermogravimetric Analysis D 4.2. Amine-Grafted Adsorbents AD
2.2. Column Breakthrough Technique (Open 4.2.1. Amine-Grafted Mesoporous Silicas AD
System) D 4.2.2. Amine-Grafted Nanofibrillated Cellu-
2.3. Multicomponent Equilibrium Adsorption lose AF
(Closed System) D 4.2.3. Amine and Ammonium-Containing Pol-
3. Physical Adsorbents E ymers AG
3.1. Carbons E 4.2.4. Amine-Functionalized Zeolites AI
3.1.1. Effect of Water on CO2 Adsorption at 4.3. Hydrophobic Amine-Containing Adsorbents AI
Subatmospheric Pressure E 4.4. Effect of Water on the Regeneration Heat AJ
3.1.2. Effect of Water on CO2 Adsorption at 5. Membranes AL
High Pressure G 6. Reactive Adsorbents AQ
3.1.3. Theoretical Studies G 6.1. Alkali Metal-Based Adsorbents AQ
3.2. Zeolites and Zeolite-like Materials H 6.1.1. Sodium Carbonate AQ
3.3. Metal−Organic Frameworks I 6.1.2. Potassium Carbonate AQ
3.3.1. Degradation Mechanisms of MOFs in 6.2. Alkaline Earth Metal-Based Adsorbents AR
Contact with Water J 6.2.1. Calcium Oxide and Hydroxide AR
3.3.2. Thermodynamic Stability of MOFs J 6.2.2. Magnesium Oxide AS
3.3.3. Kinetic Stability of MOFs K
3.4. Effect of Water on CO2 Adsorption over
MOFs L Received: July 21, 2020
3.4.1. Effect of Coordinatively Unsaturated
Sites L
3.4.2. Effect of Other SBU Features N
3.4.3. Effect of Hydrophobicity N
3.4.4. Effect of MOF Functionalization P

© XXXX American Chemical Society https://doi.org/10.1021/acs.chemrev.0c00762


A Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Table 1. Typical Compositions of Flue Gases from Different Sources69


components pulverized coal waste incineration coal-fired IGCC natural gas-fired CC cement rotary kiln
O2 (vol %) 6 7−14 12 14 7
N2 (vol %) 76 balance 66 76 59
CO2 (vol %) 11 6−12 7 3 19
H2O (vol %) 6 10−18 14 6 13
Ar (vol %) 1 1 1 1 1
SO2 (ppm) 300−5000 200−1500 10−200 5−1200
NOx (ppm) 500−800 200−500 10−100 10−300 100−1500

6.3. Other Metal Oxides AU implemented in several fossil fuel power plants.29 Compared to
7. Conclusions and Outlook AU conventional mono- and diethanolamine, KS-1 exhibits higher
Author Information AW CO2 loading, lower corrosion, limited degradation, and amine
Corresponding Author AW loss and less energy consumption for solvent stripping. As an
Authors AW alternative CO2 capture technology, adsorption using a variety
Notes AW of materials capable of reversibly or permanently removing
Biographies AW CO2 has been proposed as a more viable approach, with lower
Acknowledgments AW energy requirements.30 A wide variety of CO2 adsorbents were
References AW developed for CO2 capture,15,31−34 including metal−organic
frameworks (MOFs), 35−43 covalent organic polymers
(COPs),44−46 zeolites,47,48 clays,49−51 carbons,52−56 silica,57,58
1. INTRODUCTION basic oxides,59,60 and amine-containing porous materials.61−68
The threat of climate change and global warming to the planet Water vapor is a ubiquitous component in flue gas and in
Earth has become one of our most intractable challenges. many other CO2-rich industrial gases. Depending on their
Because of the sheer volume of anthropogenic carbon dioxide origin, flue gases may contain 6−18% water vapor (Table 1),69
(CO2) emissions, i.e., more than 36 gigatons per year, CO2 was whereas the water content of biogas is typically 6−7%.70
recognized as the largest greenhouse gas (GHG) contributor Because of its chemical properties and large dipole moment,
to global warming.1 The atmospheric CO2 concentration water has a wide range of effects on CO2 adsorption depending
increased by more than a third since the Industrial Revolution on the nature of the adsorbent and the separation conditions.
era, and continues to rise at an alarming pace, despite the Therefore, it is critically important to understand the different
temporary relief due to the exceptional circumstances of the roles that water may play with respect to CO2 adsorption, and
Covid-19 coronavirus pandemic.2,3 In particular, the rapid their implications on the selection of materials and processes.
increase of atmospheric CO2 concentration from 378 to 411 In some instances, water may strongly compete with CO2,
ppm in the short period between 2005 and 2019 is a clear sign particularly in the presence of physical adsorbents,71−73 while
of more severe weather events to come.4 Therefore, urgent in others it may have a promoting effect on CO2 uptake
actions are needed to curb GHG emissions, culminating in through more favorable surface chemistry, formation of CO2
global accords such as the 2015 Paris Agreement, adopted at hydrate, or through decreasing diffusion resistance and
the 21st session of the United Nations Framework Convention enhanced CO2 adsorption kinetics.74−76 In other circum-
on Climate Change (UNFCCC) Conference of the Parties stances, water may have a unique stabilization effect on the
(COP21), requiring limitation of the global average temper- adsorbent during repeated adsorption−desorption cycling.77,78
ature rise over the 21st century to below 2 °C.5 In 2018, an As for the effects of other flue gas components such as O2, SOx,
updated assessment by the Intergovernmental Panel on and NOx on the performance of amine-containing adsorbents,
Climate Change (IPCC) urged rapid, far-reaching and drastic as well as their CO2-induced deactivation under dry
changes in energy, industry, buildings, transport, and cities to conditiona, they were addressed in a recent review.79
limit global warming to 1.5 °C, requiring a 45% reduction in In the presence of physical adsorbents, interactions between
anthropogenic CO2 emissions by 2030, relative to 2010 levels.6 CO2 and the material surface are governed by van der Waals
Therefore, unprecedented efforts are being deployed for the forces.80 Some physical adsorbents, such as zeolites, exhibit
development of materials and processes to achieve more very high affinity toward water vapor,73,81,82 therefore, water
efficient CO 2 capture capabilities 7−16 as well as CO 2 can significantly reduce their CO2 uptake through competitive
storage17,18 and utilization.19−21 adsorption.83,84 As for hydrophobic adsorbents such as
Absorption of CO2 by alkanolamine solutions is a mature activated carbons (ACs), the negative effect of water on CO2
technology that has been practiced on large scale for decades, uptake is not as drastic as for zeolites, particularly at low
but mostly for gas purification, e.g., natural gas and hydrogen.22 relative humidity (RH). Nevertheless, at high RHs, CO2
Despite its widespread implementation, amine scrubbing has adsorption over ACs often decreases in favor of water uptake,
several drawbacks. In particular, the regeneration step is highly but at high pressure, CO2 uptake may increase significantly due
energy demanding.23 Moreover, amine solutions eventually to the formation of CO2 hydrate.85 For physical adsorbents
lead to the release of toxic pollutants such as nitrosamine, with more complex behavior such as MOFs, the effect of water
nitramines, and ammonia,24−26 and extensive corrosion of the on CO2 capture is strongly dependent on the interactions
equipment.27,28 Therefore, amine absorption for only the sake between the different components of the adsorbent and water
of CO2 capture is cost-prohibitive. Nonetheless, CO2 capture versus CO2. With MOFs emerging as a highly promising class
from flue gas using hindered amine solvents, in particular, the of adsorbents, numerous investigations dealt with the effect of
proprietary KS-1 by Mitsubishi Heavy Industries, has been water not only on CO2 adsorption but also on their structural
B https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Figure 1. (a) Water uptake followed by humid CO2 uptake at the same temperature (30 °C) and relative humidity (20% RH) over NOTT-401.
Reproduced with permission from ref 198. Copyright 2016 American Chemical Society. (b) CO2 and water uptake on InOF-1 at 30 °C in the
presence of 20% RH. Reproduced with permission from ref 196. Copyright 2015 Royal Society of Chemistry.

integrity, depending on the nature of their ligands and metal formation of stable urea linkages by an overall dehydration
nodes. In early studies, several MOFs were found to collapse reaction.77,165 In addition to the above-mentioned materials,
upon exposure to water vapor, even at low RH and room the gas phase humidity and/or prehydration of materials affects
temperature.86−89 Nevertheless, many MOFs with high other adsorbents such as metal oxides, hydroxides, and
hydrolytic stability are now readily available,90−97 particularly carbonates in a variety of ways.166−170
once the underlying criteria for MOFs hydrolytic stability were There are review articles on the effect of water vapor on CO2
fully understood.88,96−98 In addition to water stability, there are adsorption performance of MOFs.171−173 However, there are
diverse effects of moisture on CO2 adsorption over MOFs. For no comprehensive reports critically reviewing the effect of
some types of MOFs, limited water preadsorption enhances water vapor on CO2 capture over the whole range of adsorbent
CO2 uptake and its selectivity versus N2 and CH4.99 However, materials. Such a report would be highly informative because
further increase in water content often has the opposite different types of adsorbents exhibit distinct behaviors when
effect.99−101 In some other instances, interactions between CO2 adsorption is conducted under humid conditions, and/or
water molecules and ligands can trigger the reversible when the material is prehumidified. The current review aims at
deformation of the material framework, a phenomenon filling this gap, starting with discussion of water effect on CO2
known as the breathing effect, with direct implications on uptake over physisorbents, such as zeolites and MOFs, then
CO2 adsorption.102,103 All of these findings, some of which are over amine-containing adsorbents and reactive materials. To
counterintuitive, indicate that the effect of humidity on CO2 facilitate comparison of literature data, whenever possible, the
capture performance of MOFs should be investigated almost water content of the feed gas was also expressed in RH at the
case by case. Note that there are other unrelated effects of adsorption temperature. Nonetheless, it would be informative
water on MOFs, such as single-crystal to single-crystal to first discuss the different methodologies that are commonly
transformation upon hydration−dehydration.104,105 used to investigate the effect of moisture on CO2 adsorption,
In addition to CO2 physisorbents, chemical adsorbents with emphasis on their advantages and shortcomings.
based on amine-functionalized porous materials were devel- Although membrane gas separation is distinctly different
oped to combine the high affinity of amine groups toward CO2 from adsorption, both technologies use solid separating agents.
and the low energy requirement for regeneration of solid Moreover, because CO2 separation membranes often contain
adsorbents. Amine-containing adsorbents are prepared using actual CO 2 adsorbents, such as amines and poly-
different methodologies, including impregnation of amines or amines,171,174,175 MOFs,176−178 zeolites,177,179,180 carbon,181
polyamines on porous supports, particularly silicas,106−119 and ionic liquids,182,183 enzymes,184 and enzyme mimics,185,186
less frequently, activated carbons,120−122 zeolites,123−125 and they are bound to have similarities with adsorbents. Although
MOFs126−128 or covalent bonding of amines through surface progress in this area has been reported in several authoritative
silylation, 76,129−143 tethering, 144−153 ligand modifica- reviews,171,174,187−190 the effect of water on CO2 transport and
tion,154−156 in situ polymerization,157,158 and other strat- selectivity across membranes will be discussed succinctly.
egies.159−161 CO2 adsorption over amine-functionalized
materials involves chemical interactions mostly with primary 2. MEASURING CO2 UPTAKE UNDER HUMID
and secondary amines, typically to produce ammonium CONDITIONS
carbamate.61,62,66 Remarkably, not only water is tolerated by Experimental studies investigating the effect of humidity on
such adsorbents,162 but the presence of water vapor in the feed CO2 adsorption involved different methods to measure
gas often promotes CO2 adsorption, typically through the multicomponent CO2 uptake. This includes (i) thermogravi-
formation of carbonate and bicarbonate with more favorable metric analysis (TGA), alone76,162,191−202 or combined with
CO2/N stoichiometric ratio than ammonium carbamate.61,62,66 other analytical techniques such as mass spectrometry (MS),
Water vapor can also trigger CO2 adsorption by tertiary amines infrared spectroscopy (IR), gas chromatography (GC), or
via bicarbonate formation.163,164 More importantly, water other specific detectors for CO2 and/or RH measurements
vapor was found to be necessary to ensure long-term stability with additional calibrations and validation proce-
of amine-containing adsorbents against CO2-induced deacti- dures,139,155,203 (ii) dynamic column-breakthrough technique
vation, which takes place under dry condition through the with accurate flow measurements in combination with different
C https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

analytical tools,204 and (iii) static multicomponent equilibrium CO2 adsorption−desorption cycle was conducted on the
adsorption.100,205 adsorbent with lower moisture content. For each cycle, the
2.1. Thermogravimetric Analysis CO2 uptake was determined based on the difference between
the mass after and before CO2 adsorption. Here also, the
When using TGA alone, the sample is often prehumidified
displacement of water by CO2, particularly for hydrophobic
at a desired temperature in the presence of an inert gas
adsorbents, with less strongly held water, cannot be ignored.
with a constant relative humidity, then exposed
to the CO2-containing feed with the same RH, at the same 2.2. Column Breakthrough Technique (Open System)
temperature, while monitoring the weight gain As schematically shown in Figure 3a, in breakthrough
(Figure 1a).76,139,162,191−194,196−198,201,202,206 Alternatively, experiments, a gas mixture is flowed through a packed bed
both measurements can been undertaken separately (Figure of adsorbent until the outlet gas stream reaches the same
1b).196 Assuming that the adsorption of water at a given RH is composition as the inlet feed.205,207 In the course of
not affected by CO2, and the adsorption of the inert gas, often adsorption, the outlet gas composition is continuously
N2, is negligible, the sample weights at both equilibria would
measured using a variety of techniques such as gas
yield the corresponding water and CO2 uptakes. However, on
the basis of adsorption measurements using, in addition to the chromatography−thermal conductivity detection (GC-
method described above, independent measurements of CO2 TCD)75,208−215 MS207,216−219 or CO2 detectors.111,163,220−227
uptake by MS during temperature programmed desorption The CO2 uptake is calculated based on the gas composition
(TPD), these assumptions were found to be questionable.131 and flow rate and the breakthrough curve versus time.
Using a triamine-grafted silica in the presence of 5% CO2/N2 Alternatively, the composition of the adsorbed phase may be
with 27−74% RH, CO2 uptakes measured by TGA only, were determined by TPD in an inert gas, coupled with MS
lower by as much as 50% compared to the combined TGA-MS measurements.131 Although accurate determination of the
procedure, thus severely underestimating the actual uptake. It equilibrium amounts of adsorbed species based on break-
was surmised that preadsorbed water is in part, displaced by through experiments is not trivial,205 this technique is
CO2 adsorption, which cannot be detected by TGA. In extremely useful for comparative evaluation of adsorbents.
addition, the CO2 uptake after prehumidification was found to Pertinent to this review is the use of CO2−water−inert gas
be as much as 10% higher compared to humid CO2 uptake mixtures, with or without prehydration of the materi-
over fresh material, particularly at high RH, indicating that it is al.75,111,157,163,208−229
difficult to reproduce the same equibrium conditions.131 In many instances, the outlet concentration of CO2 reaches
Furthermore, to measure CO2 uptake by TPD, the gas phase the same level as the inlet gas, yet the water front is still very
has to be flushed, altering the actual equilibrium state. close to the column entrance, while most of the adsorbent is in
In another variation of this method applied to a large the presence of a desiccated gas mixture. This may lead to the
number of MOFs, Chanut et al.195 first saturated their hasty conclusion that water has a limited or no effect on CO2
materials with water vapor (100% RH) at room temperature adsorption. If the experiment proceeds for longer time, the
for 2 days, using an airtight vessel containing liquid water. The advancing water front may displace some of the already
sample was then loaded in the TGA instrument and submitted
adsorbed CO2, leading to lower CO2 capacity under near-
to a five-step procedure as illustrated in Figure 2. It was first
equilibrium conditions. Moreover, if water is not completely
treated in nitrogen at 30 °C for 8 h to remove weakly held
water, followed by CO2 adsorption in the presence of dry pure removed during regeneration, consecutive adsorption−desorp-
CO2. The desorption step was conducted for 90 min, while the tion cycles, lead to decreasing CO2 working capacity as the
sample was heated to a desired temperature to remove all CO2 amount of adsorbed water builds up over time.
and a portion of preadsorbed water. Finally, the sample was 2.3. Multicomponent Equilibrium Adsorption (Closed
cooled down to 30 °C while purging with nitrogen, and a new System)
In a typical closed system multicomponent equilibrium
adsorption measurement (Figure 3b), a gas mixture with
known volume, pressure, temperature, and composition is
dosed to a weighed sample, placed in an evacuated space.100,205
The gas is often recirculated continuously to improve mixing.
Once the equilibrium is established, the gas phase is analyzed,
typically by GC or MS, and the composition of the adsorbed
phase is deduced by material balance knowing the dead volume
or the total amount adsorbed. Contrary to the open system
where the gas composition at equilibrium is the same as the
inlet gas, in the closed system, the equilibrium state is difficult
to predict.
Among the three methods, the closed-system multi-
component adsorption with independent measurements of
Figure 2. Consecutive CO2 adsorption−desorption cycles conducted gas phase composition before and after adsorption provides the
by TGA. (a) Temperature and (b) mass profiles. Steps (1), (2)/(6),
(3), (4), and (5) represent sample pretreatment, CO2 adsorption,
most accurate measurements of actual thermodynamic
CO2 desorption, partial water desorption, and sample conditioning, equilibria. This is particularly important for multicomponent
respectively. Reproduced with permission from ref 195. Copyright adsorption isotherms to be used for modeling, process
2013 John Wiley and Sons. optimization, and scale-up purposes.
D https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Figure 3. Schematic representation of systems for multicomponent equilibrium adsorption measurements (a) column-breakthrough (open system),
and (b) closed recirculation system. Reproduced with permission from ref 205. Copyright 2015 American Chemical Society.

Figure 4. Breakthrough data for a mixture of CO2, N2, O2, and H2O on a biomass-derived AC, at 25 °C and 130 kPa; (a) outlet concentration on
dry basis; (b) partial pressure of H2O in the outlet gas. Reproduced with permission from ref 209. Copyright 2014 American Chemical Society.

3. PHYSICAL ADSORBENTS carbon was proposed earlier based on experimental248,249 and


theoretical250 data. In some instances, the structural properties
3.1. Carbons of ACs can be adjusted to increase CO2 selectivity over water.
Because of their often large surface area and pore volume and Koresh et al.251,252 found that mild thermal activation of a
their low cost, ACs have been traditionally used for separation fibrous carbon molecular sieve (TCM-128) resulted in a
purposes,230,231 including CO2 capture.54,56,232−236 They are controlled enlargement of the pores, with tiny hydrophobic
often prepared through carbonization of different low-cost entrances (constrictions), possibly due to nonoxidized graphite
precursors, including coal and a wide range of biomass interlayers. Despite the smaller size of water molecules relative
materials, e.g., wood, fruit stones, nut, and coconut shells, to nitrogen and CO2, water was not able to penetrate such
followed by chemical or physical activation.232,237,238 Depend- constrictions due to the formation of water clusters; however,
ing on their preparation conditions, ACs may be microporous the other two gases did penetrate. Therefore, the presence of
and/or mesoporous in nature. It should be noted that in such hydrophobic adsorption sites significantly increased the
addition to ACs, other carbon-based materials such as carbon selectivity of CO2 over water.
nanotubes (CNTs),239 carbon molecular sieves,240 and 3.1.1. Effect of Water on CO2 Adsorption at Subatmo-
graphene241 were also investigated for CO2 capture. Two spheric Pressure. Most studies on carbon-based materials
properties were recognized to be of critical importance for CO2 conducted at subatmospheric pressures reported a negative
adsorption over carbon materials, namely the occurrence of impact of humidity on CO2 uptake as a result of competitive
nitrogen species and favorable pore structure.55,242 For adsorption of water. Duran et al.208 used a commercial-grade
example, Sethia and Sayari,55 and others243,244 reported CO2 (Norit R) and two biomass-derived (IH2 and IH3) ACs for
uptakes exceeding 5 mmol/g at 25 °C and 1 atm over nitrogen- CO2 capture from a flue gas stream generated in a solid waste
doped carbons with high ultramicropore (0.7−0.9 nm) incinerator. Single-component adsorption isotherms of N2,
volume. CO2, and H2O were used to estimate the selectivity of CO2
ACs often exhibit type I CO2 adsorption isotherms, versus water and nitrogen for all adsorbents at 30−70 °C. In
indicating their high affinity toward CO2 at low partial the presence of all carbons, water was found to preferentially
pressures.240 In addition, they often have low affinity toward adsorb compared to CO2, particularly at the lowest temper-
water vapor at low RH; however, at high RH levels, water ature. Using 2% H2O/N2 and 2% H2O/8% CO2/N2 gas
uptake increases significantly. This is reflected in the “S” mixtures, corresponding to 48 and 16% RH at 30 and 50 °C,
shaped adsorption isotherms of water on ACs, corresponding respectively, water uptake was found to be hardly affected by
to type V.245,246 According to the model proposed by Do and CO2. However, the presence of humidity reduced CO2 uptake
Do,247 because of strong affinity of water to surface functional significantly, particularly at the highest RH. Moreover,
groups, water adsorbs first over such sites, then on top of the prehydration of the adsorbent at 30 °C reduced the CO2
chemisorbed water molecules through hydrogen bonding, to uptake by as much as 46%. Younas et al.223 studied the effect of
eventually form water clusters that move into the micropores. activation method on the competitive adsorption of CO2 and
The formation of such clusters on the surface of activated water vapor on a palm shell activated carbon (PSAC). Both
E https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Figure 5. (a) Gas phase concentration and (b) amount of adsorbed CO2 and water versus axial position. Reproduced with permission from ref 256.
Copyright 2015 John Wiley and Sons.

Figure 6. (a) CO2 adsorption isotherms at 2 °C on a microporous carbon with different moisture contents: (1) Rw = 0, (2) Rw = 1.74, (3) Rw =
1.32, and (4) Rw = 1.65. (b) CO2 adsorption isotherms using Rw = 1.31 at (1) 8 °C, (2) 6 °C, (3) 4 °C, and (4) 2 °C, while (5) is the CO2
desorption isotherm at 2 °C. Reproduced with permission from ref 85. Copyright 2007 Elsevier.

chemical activation using 20−50% NaOH solutions and adsorption on fresh or fully regenerated carbon material was
thermal activation were performed on the material. The not affected by the presence of water in the feed. Dynamic
presence of water vapor (20% RH) was found to reduce CO2 adsorption at 25 °C under dry 1% CO2/N2 or humid (40−
uptake; however, the negative effect of humidity was more 50% RH) conditions over a fresh adsorbent showed that the
severe for the thermally than the chemically activated PSAC, CO2 breakthrough time and the saturation uptake were the
with 45% and 13% decline, respectively. Similarly, water vapor same.254 Moreover, using 15% CO2/N2 at 25 °C, the CO2
(94% RH) in 15% CO2/N2 at 30 °C led to ca. 14% decrease in breakthrough and saturation occurred at ca. 17 and 48 min,
CO2 uptake over KOH-treated AC.253 respectively (Figure 4a), regardless of the presence of moisture
Some studies reported that humidity has no effect on CO2 (ca. 50% RH).209 However, water took about 470 min to break
uptake of carbonaceous materials. Using vacuum swing through, while adsorption continued for well over 24 h (Figure
adsorption (VSA) to investigate the effect of humidity on 4b). This indicates that at the time of CO2 apparent saturation,
CO2 adsorption on an activated carbon bed, Xu et al.222 because of the much-delayed water adsorption front, most of
observed that increasing the water vapor content in a 12% the adsorbent was in the presence of a dried CO2/N2 gas
CO2/air inlet gas mixture up to 7.2% did not change the CO2 mixture. Nonetheless, as the experiment proceeded further,
recovery and purity, presumably because this was a kinetic there was evidence of a roll-up phenomenon, indicating that as
separation and CO2 adsorbed much faster than water vapor. At the water front advanced, it displaced part of the CO2 initially
the time the outlet and inlet CO2 concentrations became adsorbed. The ideal adsorbed solution theory (IAST)
equal, all water was essentially adsorbed at the entrance of the predicted that over this time, 60% of adsorbed CO2 would
reactor, forming a cone with rapidly decreasing concentration be displaced. 254 This finding has important practical
along the bed length. In a series of reports, Plaza et al.209,254,255 implications because it indicates that the effect of water on
carried out a comprehensive investigation on the effect of CO2 adsorption over carbon depends on the degree of the
water on CO2 adsorption in a fixed-bed unit under different adsorbent hydration. Moreover, because the isosteric heat of
conditions using a biochar obtained by carbonization of olive water adsorption is significantly higher than CO2, it is likely
stones. In agreement with others,210,222 they found that CO2 that in a cyclic operation, water will not be completely
F https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

removed because of its slower desorption and high heat


capacity, leading to increasing moisture content and eventually
decreasing CO2 uptake.254
Likewise, using a microporous coffee-based AC,256 a much
shorter breakthrough time was observed for CO2 relative to
water at 52% RH (3 versus 200 min). Although CO2 uptake
was not reduced in the presence of humidity over short
adsorption durations, experimental and modeling data showed
that extended adsorption time led to displacement of CO2 by
water, eventually resulting in decreasing CO2 uptake from 1.25
to 0.6 mmol/g at t = 523 min, the equilibrium time for water
adsorption. CO2 and water concentration profiles in the gas
phase (Figure 5a) and adsorbed phase (Figure 5b) were
determined throughout the bed via simulation using Aspen
adsorption software at three different times, i.e., right before
and after CO2 breakthrough at t = 2.2 and 4 min and at
equilibrium (t = 523 min). After CO2 uptake reached the Figure 7. CO2 isotherms at 2 °C on a microporous/mesoporous
maximum, as time elapsed, the water concentration front carbon with different moisture contents: (1) Rw = 0, (2) Rw = 1.35,
moved through the adsorbent bed, leading to partial (3) Rw = 1.80, (4) Rw = 2.15, and (5) Rw = 2.36. Reproduced with
desorption of CO2, consistent with experimental data. Notice permission from ref 257. Copyright 2008 Elsevier.
that for two commercial ACs (Takeda and Norit), even when
homogeneously distributed throughout the material, the
limited amount of preadsorbed water of 3.5−4 wt % did not
affect CO2 uptake.195
3.1.2. Effect of Water on CO2 Adsorption at High
Pressure. There is a broad consensus regarding the beneficial
effect of water on CO2 adsorption over carbon at high
pressure. Using a microporous AC derived from coconut shells,
Sun et al.85 measured CO2 adsorption isotherms at 2 °C up to
40 bar over prehydrated adsorbents. The moisture content Rw,
was defined as the ratio of adsorbed water to carbon.
Compared to dry carbon, the CO2 uptake over moisturized
materials was lower at low pressure; however, at 15−20 bar,
the uptake increased dramatically (Figure 6a), which was
attributed to enhanced solubility of CO2 in water, correspond-
ing to the formation of CO2 hydrate. Figure 6b shows that at Figure 8. CO2 adsorption isotherms for uGil carbon prehydrated with
increasing temperatures, the inflection point occurs at higher different amounts of water. CO2 uptake is per gram of dry carbon
pressures, with values comparable to the pressure required for material. Reproduced with permission from ref 258. Copyright 2017
the formation of CO2 hydrate in pure water. However, unlike Springer Nature.
exclusively microporous carbons, CO2 adsorption isotherms
over a bamboo-derived carbon with microporous/mesoporous reported earlier.260 At 25 °C, the presence of 1% H2O (32%
bimodal pore structure in the presence of water, exhibited RH) resulted in the least CO2 uptake reduction of 13% for NQ
similar transitions, but did not reach a plateau associated with sites, followed by 25%, 28%, and 58% loss for pyrrolic,
the formation of CO2 hydrate (Figure 7).257 pyridonic, and oxidized pyridinic nitrogen species, respectively.
Jalilov et al.258 demonstrated that compared to dry Although water vapor content at 25 °C cannot exceed 3.1%,
condition, prehydration of asphalt-derived AC (uGil) led to calculations were made at 10%. This resulted in dramatic
remarkable increase in CO2 uptake at −15 °C and 20 bar, from reduction in CO2 adsorption capacity for all materials, with the
15 to 48 mmol/g, and enhanced the CO2/CH4 selectivity from highest loss being associated with the quaternary (98%) and
ca. 2 to 200, with equilibrium and kinetic data suggesting that pyridinic (97%) nitrogen. Because of the presence of high-
CO2 uptake proceeded via both physisorption and chem- energy surface sites, however, CO2 adsorption was competitive
isorption. Moreover, as the water content of the adsorbent with water at low pressure over oxidized pyridine, pyridonic,
increased, the shape of CO2 isotherm changed from type I to and quaternary nitrogen-containing carbons. Pyrrolic nitrogen
type V (Figure 8), consistent with the formation of CO2 exhibited completely different behavior as it adsorbed more
hydrate. water than CO2 at low CO2 partial pressure (<0.1 mbar), with
3.1.3. Theoretical Studies. Using density functional CO2 uptake going through a maximum at 0.25 mbar. This was
theory (DFT) and grand canonical Monte Carlo (GCMC) attributed to the steric and electronic properties of pyrrolic
simulations, Psarras et al.259 studied the adsorption of CO2 nitrogen, which enabled the material to adsorb CO2 despite
with different RHs over carbon containing pyrrolic (N5), prior adsorption of water. Similar simulations conducted at 60
oxidized pyridinic nitrogen (NO), pyridonic (NP), and °C showed that in the presence of 1% H2O (5% RH), the
quaternary nitrogen (NQ) species. The simulated CO2 relative CO2 capacity decreased in the order of NO > NP >
adsorption isotherms in the presence of humidity (1 and NQ > N5, which is different from the trend observed at 25 °C,
10%) at two temperatures (25 and 60 °C) were compared to i.e., NQ > NP > NO > N5, indicating that CO2 adsorption
the calculated isotherms under dry and humid conditions, over different nitrogen species exhibits different temperature
G https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

dependence behaviors. Moreover, the negative effect of 1% Table 2. CO2 Uptake for Activated Carbon Adsorbents at
H2O (5% RH) was particularly severe for NQ, N5, and NO, Atmospheric Conditions
with 38%, 19%, and 19% CO2 adsorption decline, respectively.
CO2 uptake
Surprisingly, at 10% H2O (51% RH), an increase in CO2 (mmol/g)
capacity was observed over NQ, N5, and NO, with minor
CO2 adsorption
negative effect of ca. 10% for NP relative to adsorption at 25 content temperature RH
°C. material (%) (°C) (%) dry humid ref
You and Liu261 used a Langmuir-type empirical model to sawdust-based 10 50 22 0.35 0.32 208
simulate fixed-bed adsorption and desorption of CO2 from wet carbon (IH2)
flue gas over AC. Using 12% CO2/N2 at 60 °C with 0−100% sawdust-based 10 50 22 0.42 0.38 208
carbon (IH3)
RH at atmospheric pressure, the recovery of CO2 was found to commercial 10 50 22 0.31 0.29 208
decrease from 84% under dry condition to 61% at 100% RH. carbon
This was attributed to the fact that water uptake increases with (Norit R)
humidity, whereas CO2 shows the opposite behavior. More- olive stones-based 13 25 48 1.1 1.1 209
carbon
over, the increase in water uptake with humidity led to a rise in
the average temperature of the bed, owing to the heat of coffee-based 16 25 48 1.25 1.25a 256
adsorption, with adverse effect on CO2 uptake, notably beyond activated carbon
50% RH. 16 25 48 1.3 0.6b 256
In summary, water in the form of vapor or preadsorbed
moisture can adversely affect the equilibrium CO2 capacity of palm shell 15 30 20 0.80 0.70 223
carbonaceous materials at low pressures due to competitive activated carbon
(PSAC)
adsorption. The negative effect of water is particularly striking Ni-impregnated 15 30 20 0.97 0.66 223
for feed streams with high RHs. However, in fixed bed (24 wt %)
operations, because the CO2 adsorption front moves through PSAC
the bed much faster than water, the negative effect of water KOH-activated 15 30 20 1.09 0.83 223
PSAC
vapor may be circumvented by kinetic separation. In contrast,
thermally 15 30 20 0.98 0.84 223
for CO2 adsorption at high pressures, the presence of activated PSAC
preadsorbed water promotes CO2 uptake through mechanisms KOH-treated AC 15 30 94 0.84 0.72 253
other than adsorption, including dissolution and hydrate ACFCd 15 25 75 0.68 0.68 210
formation. Typically, in the presence of wet carbon, CO2 ACFC/KOH 15 25 75 0.84 0.84 210
adsorption isotherms change from type I to type V as the commercial AC 100 30 100 1.93 2.01c 195
moisture content of the adsorbent increases. Literature data on (Takeda)
CO2 uptake by carbon-based materials at atmospheric pressure commercial AC 100 30 100 1.65 1.70c 195
(Norit)
are summarized in Table 2. a
CO2 uptake after 15 from the beginning of breakthrough test. bCO2
3.2. Zeolites and Zeolite-like Materials uptake after 523 min from the beginning of breakthrough test.
Zeolites are microporous crystalline materials with a wide
c
Prehumified at 100% RH. dPolyacrylonitrile activated carbon fiber
range of applications, in particular gas separation and
purification.262−266 Their narrow micropore size distributions decreasing Si/Al ratios because of higher content of extra-
and extra-framework cations make them an appropriate choice framework cations.275 However, in the presence of water (5%
for separating small molecules via electrostatic interactions or H2O/5% CO2/N2), CO2 uptake showed the opposite trend,
size exclusion (molecular sieving).267 Several alkali and alkaline whereas water uptake increased at decreasing Si/Al ratios,
earth-exchanged zeolites were reported to be efficient CO2 presumably because of stronger interactions between water and
adsorbents because of the strong interactions between the extra the increasingly hydrophilic framework. Encapsulation of
framework cations and CO2 molecules.47,48,267−271 Never- zeolites in a hydrophobic support helps mitigate the adverse
theless, ion-exchanged zeolites with low Si/Al ratio tend to effect of moisture. A hierarchically structured composite
adsorb not only molecules with large quadrupole moments monolith consisting of 43 wt % zeolite in carbon, prepared
such as CO2, but also species with high dipole moments such by carbonization of zeolite-β/polymer foam, exhibited in the
as H2O, through electrostatic forces.271 In particular, when presence of 15% CO2/N2 with 50% RH at 25 °C, a stable
dealing with CO2 adsorption at low concentration, as in direct working capacity of ca. 0.7 mmol/g over 15 adsorption−
air capture (DAC), the presence of humidity reduces CO2 desorption cycles.276 In contrast, pure zeolite β showed higher
uptake significantly.272,273 Investigating CO2 adsorption over working capacity for the first cycle (1.37 mmol/g) but
13X zeolite in the presence of 5% CO2/N2, Sayari group218 negligible CO2 uptake thereafter, which was attributed to
found that at increasing RH, more water was adsorbed at the strongly bound water molecules.
expense of CO2, whose uptake at 25 °C decreased by 80% and Contrary to hydrophilic aluminosilicate zeolites, such as
93% at 27% and 74% RH, respectively. Likewise, Li et al.274 NaX and 5A, CO2 uptake over coppersilicate (SGU-29), a
found that at 30 °C, 13X zeolite adsorbed 99% less CO2 in the crystalline microporous material with a structure akin to AM-6
presence of 8.8% CO2 and 3.4% H2O in air (80% RH) vanadosilicate and Engelhardt ETS-10 titanosilicate,277 was
compared to dry conditions, indicating that the competitive found to be unaffected by moisture.278 Exposure of SGU-29 at
adsorption is strongly in favor of water. 25 °C to dry (9.9% CO2/18.8% O2/N2) and humid (9.6%
Consistent with the above-discussed findings, Monte Carlo CO2/18.3% O2/2.9% H2O/N2, 90% RH) gas mixtures led to
simulations for CO2 and water adsorption on faujasite zeolite CO2 uptake of 2.47 and 2.43 mmol/g, respectively. Fourier-
showed that, under dry conditions, CO2 uptake increases with transform infrared spectroscopy (FTIR) measurements
H https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Table 3. CO2 Uptake for Zeolites in the Presence of Water at 1 atm


CO2 uptake (mmol/g)
material CO2 content (%)/balance adsorption temperature (°C) RH (%) dry humid ref
zeolite 13X 5/N2 25 27 1.95 0.48 218
5/N2 25 74 1.95 0.12 218
0.04/air 30 49 NPa 0.03 272
15/N2 30 100 3.18 0.60 272
100 30 100 3.86 1.22 272
0.0395/air 25 80 0.32 0 273
100 30 2.4a NPa 0.51 195
100 30 25.7b NPa 0.05 195
20/N2 40 24 4.1 0 205

Na−Y 15/N2 60 61 1.23 0.03 281


Li-LSX 0.04/air 25 80 0.82 0.01 273
K-LSX 0.04/air 25 80 0.25 0.03 273
SGU-29 9.8% CO2/18% O2/N2 25 90 2.47 2.43 278
ETS-4 9.8% CO2/18% O2/N2 25 90 1.72 1.65 278
ETS-10 9.8% CO2/18% O2/N2 25 90 1.88 1.81 278
AM-6 9.8% CO2/18% O2/N2 25 90 2.09 2.05 278

zeolite NaX 9.8% CO2/18% O2/N2 25 90 1.85 1.50 278


15% CO2/3% H2O/2% Ar/N2 40 41 2.1 0.23 279

SGU-29 15% CO2/3% H2O/2% Ar/N2 40 41 1.3 0.18 279


ETS-10 15% CO2/3% H2O/2% Ar/N2 40 41 1.4 0.17 279
zeolite 5A 20/N2 40 23 4.1 0.3 205
a
Not provided. bWater content in prehydrated material in wt %.

provided evidence that CO2 and H2O adsorbed non- encapsulate the zeolite in a hydrophobic matrix. Moreover,
competitively on distinct sites, whereas NaX has, in addition because of their lower affinity for water, zeolite-like materials
to specific sites for CO2 and H2O, shared adsorption sites. The with neutral frameworks such as AlPOs may exhibit high CO2
isosteric heat of adsorption of CO2 over SGU-29 was higher to water selectivity. Functionalization of zeolites with amine-
compared to that of zeolite NaX, while the isosteric heat of containing compounds (vide infra) is another strategy for
water adsorption was lower, indicating that SGU-29 binds CO2 increasing the selectivity of CO2 over water, even though
more strongly than does NaX but has a lower affinity to water. amine modification of zeolites instead of cheaper silicas may
Surprisingly, soon after, Kim et al.279 reported contradictory not be justified. CO2 adsorption data for selected zeolites in
findings. They claimed that all three zeolites, NaX, ETS-10, dry and humid condition are listed in Table 3.
and SGU-29, lost 85−90% CO2 uptake upon addition of 3% 3.3. Metal−Organic Frameworks
water (41% RH) to the 25% CO2/N2 feed gas at 40 °C.
Computational screening of hundreds of pure silica zeolites Metal−organic frameworks are crystalline porous materials
using GCMC simulations showed that, depending on the comprised of a three-dimensional network of metallic ions or
CO2/H2O binding site separations and CO2−H2O interaction clusters, referred to as secondary building units (SBU),
energies, humidity may promote or impede CO2 adsorption.280 connected by organic linkers or ligands.282−287 Because of
Enhancement in CO2 uptake occurs when the material exhibits their porous structure, MOFs may develop extraordinarily large
clearly distinct CO2 and water binding sites, with long-range internal surface area and pore volume.288,289 Owing to the
favorable CO2−H2O interactions. Conversely, for zeolites with diversity of their structure and composition, they are ideal
overlapping CO2 and H2O binding sites, the CO2 uptake is candidates for a wide range of applications, including gas
significantly reduced in the presence of water because of storage,290−292 postcombustion CO2 capture,35,293−298 and
adsorption competition. If the binding sites for CO2 and H2O others.299−309 Although amine functionalization was used to
are nearby, but not completely overlapped, water may affect impart CO2 chemisorption, most MOFs behave as physical
CO2 uptake either way, depending on whether the reduced adsorbents.
pore volume for CO2 adsorption due to the presence of water Unlike conventional physical adsorbents such as carbons and
is overcompensated by the enhanced CO2−H2O interactions zeolites, interactions between water and MOFs cannot be
or not. described by a single mechanism because MOFs exhibit
In summary, the strong electrostatic interactions between different behaviors associated with the wide variety of their
extra-framework cations and water molecules result in structures and properties. Therefore, depending on the
significantly higher affinity of zeolites toward water compared composition and structure of MOFs as well as the nature
to CO2, leading to less favorable CO2 adsorption. Short of and strength of their interactions with water, the ultimate
dehydrating the feed gas, one way to minimize the negative outcome may vary considerably. This diversity also applies to
impact of water on CO2 adsorption is to use high-silica the effect of water on CO2 adsorption over MOFs. Therefore,
hydrophobic zeolites with less extra framework cations or to although the main focus of this review is on the effect of
I https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Figure 9. HSAB theory for MOFs stability (BTP, benzenetripyrazolate; BDP, 1,4-benzenedipyrazolate. Reproduced with permission from ref 95.
Copyright 2018 John Wiley and Sons.

moisture on CO2 adsorption, for MOFs it is important to first frameworks in the presence of water occurs through two
discuss, albeit briefly, the adsorption of water and its potential mechanisms, namely ligand displacement and hydrolysis.316 In
effect on their structural integrity. More details on the behavior ligand displacement reactions under neutral condition, water
of MOFs in the presence of water may be found in several molecules are inserted into the metal−ligand bond of the
authoritative reviews.86,88,96−98,308,310 framework, leading to the formation of hydrated cations and
MOFs exhibit a wide range of water adsorption isotherms. the release of deprotonated ligands (eq 1). The mechanism is
Hydrophilic MOFs such as Ni-DOBDC and Mg-DOBDC slightly different under basic or acidic condition.317 For
(DOBDC4−: 2,5-dioxido-1,4-benzenedicarboxylate), also instance, in the presence of water vapor, MOF-5 undergoes
termed M-MOF-74, contain coordinatively unsaturated metal complete structural collapse via ligand displacement (eq 2).318
sites (CUS) that adsorb water strongly, very much like extra-
framework cations in zeolites. Therefore, such MOFs exhibit Mm − Ln − + H 2O → Mm + − (OH 2) ···Ln − (1)
water adsorption isotherms of type I. In contrast, hydrophobic
MOFs such as MIL-125-NH2 and Cr-MIL-101 do not adsorb (Zn4O)(BDC)3 + 4H 2O
much water until the pressure for capillary condensation is → [(Zn4O)(H 2O)4 (BDC)2 ]2 + . . . BDC2 − (2)
reached, leading to type V isotherms.88 In general, water
adsorption on MOFs occurs via three mechanisms,88,98,308 Hydrolysis of MOFs is akin to an acid−base reaction, where
namely (i) chemisorption on high affinity metallic clusters with the metal−ligand species react with water molecules to form a
CUS as in M-DOBDC (M = Zn, Ni, Co), leading to type I hydroxylated SBU and free protonated ligand (eq 3). For
isotherms. In some instances, adsorption and desorption of instance, the disintegration of Cu(BDC)(TED)0.5 (BDC, 1,4-
water on CUS leads to local or global deformation of the benzene dicarboxylate; TED, triethylenediamine) takes place
framework,311 and (ii) reversible cluster adsorption, similar to through hydrolysis of Cu−O−C, as demonstrated by IR
the water adsorption mechanism in the presence of carbon, measurements (eq 4).319
discussed earlier.247−250 This mechanism takes place in MOFs,
with pores below the critical pore size for water capillary Mm + − Ln − + H 2O → Mm + − (OH)− + HL(n − 1) − (3)
condensation of ca. 2.1 nm. For example, in the presence of
Cu−O−C + H 2O → Cu−O−H + C − OH (4)
MOF-801, the primary adsorption sites for water are the
bridging OH groups on the Zr6O4(OH)4 metallic clusters, Understandably, the hydrolytic stability of MOFs depends on
leading first to tetrahedral water clusters, confined within the thermodynamic factors such as the strength of the metal−
MOFs tetrahedral pores. Further adsorption gives rise to larger linker bond and on kinetic factors that impede the degradation
cubic clusters, which facilitate water clustering in the larger reaction with water such as increased steric hindrance or
octahedral pores. Eventually, all water clusters coalesce, leading hydrophobicity.
to continuous pore filling,98 and (iii) capillary condensation 3.3.2. Thermodynamic Stability of MOFs. The
through pore-filling in MOFs with pore diameter ≥2 nm. This thermodynamic stability of MOFs is primarily due to the
is the most common physical adsorption mechanism in strength of the metal−linker bond. This bond can be regarded
mesoporous materials. At rising moisture content, reversible as the result of a reaction between the metallic cation acting as
multilayer adsorption takes place, followed by capillary a Lewis acid and the ligand anion acting as a Lewis base,
condensation at specific pressures, depending on the pore indicating the strong relationship between the strength of the
structure, leading to type IV or type V isotherms. acid and base and the stability of the corresponding MOF.
In many instances, water vapor affects the structural integrity Hence, it is expected that as pKa of the protonated ligand
of MOFs. The hydrolytic stability of a MOF is defined as the increases, the thermodynamic stability increases. As for the
preservation of its pore structure and crystallinity after metal, the higher the charge and the smaller the size, i.e., the
exposure to a certain level of humidity for a period of time higher the charge density, the stronger the metal−linker bond.
at a given temperature.312 Techniques such as powder X-ray As shown in Figure 9, this relationship can also be explained
diffraction (PXRD) and nitrogen adsorption at 77 K are the using the hard−soft acid−base (HSAB) concept, which
most common tools used to evaluate the stability of MOFs in indicates that strong bonds occur between hard acid and
the presence of water vapor or, in some instances, liquid hard base or soft acid and soft base.95
water.313−315 Most MOFs use carboxylate or N-heterocycle-based linkers.
3.3.1. Degradation Mechanisms of MOFs in Contact Carboxylate ligands, such as BDC and BTC (1,3,5-benzene
with Water. The degradation of MOFs with vulnerable tricarboxylate), are strong Lewis bases as indicated by the low
J https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Figure 10. Stability of different MOFs as a function of temperature and percent steam. The red and magenta numbers indicate experimental and
simulated activation energy (in kcal/mol) for ligand displacement by water, respectively. Reproduced with permission from ref 316. Copyright
2009 American Chemical Society.

pKa of their conjugated acids of 3.9 and 3.6, respectively. They area versus 8% for Mg-DOBDC, although no further
can also be regarded as hard bases, which form stable MOFs improvement took place at higher Ni content. As for Mg-
with high-valent metal ions (hard acids), such as Fe3+, Cr3+, Co-DOBDC, the retained surface area after water exposure
Al3+, Zr4+, Hf4+, and Ti4+. The MIL series is a typical example increased linearly with Co content. Further computational and
of hydrolytically stable carboxylate-based MOFs, containing experimental studies331 indicated that in mixed metal materials,
Fe3+, Cr3+, or Al3+.102,320,321 Likewise, Zr-carboxylate MOFs distortion of the apical M−O bond length exceeded what
such as UiO-66 and derivatives are also known to be would be expected from simple metal size effects, suggesting
extraordinarily stable.90,322−325 Nitrogen containing hetero- that the M−O bond length plays a key role in M-DOBDC
cycles (high pKa), such as azolates,326 adeninates,315 and hydrolytic stability. Consistent with the prevailing under-
pyrazolates,327,328 are soft bases. Their combinations with low- standing of H2O adsorption on MOFs,86,332 it was inferred that
valent metals (soft acids), such as Zn2+, Ni2+, Cu2+, and Mn2+ the increased water stability of Mg-Ni-DOBDC arises from the
also give rise to water-stable MOFs. Guided by the HSAB contraction of the apical M−O bond.
concept, many hydrolytically stable MOFs have been 3.3.3. Kinetic Stability of MOFs. In a seminal
assembled and investigated.95,96 contribution, Low et al.316 investigated the hydrolytic stability
Although the metal−ligand bond strength based on pKa of a series of MOFs as a function of temperature and steam
values and the HSAB principle are useful to predict the content and calculated the free energy for their hydrolysis
stability of MOFs, another important thermodynamic factor is (ΔGhyd) as well as the free (ΔGdisp) and activation (ΔE⧧disp)
the metal lability with water, i.e., its likelihood to undergo energy for ligand displacement. As shown in Figure 10,
hydrolysis, which depends on the energy difference between experimental data were in good agreement with calculated
the frontier orbitals of the metal and water. The larger the activation energies for ligand displacement, indicating that the
energy gap between these frontier orbitals, the higher the water observed hydrolytic stabilities are under kinetic control.
stability. For example, Kang et al.329 investigated the effect of Kinetic factors include any feature that affects the activation
metal ions on the stability of three isostructural MOFs with the energy of hydrolysis. For example, increased hydrophobicity
same ligand, namely, Cr-BDC, Al-BDC, and V-BDC, in pure has a positive effect on MOF stability by preventing water from
water and in basic and acidic aqueous solutions. In all cases, adsorbing in the pores or from clustering around the metal
the stability decreased in the order of Cr-BDC > Al-BDC > V- center.96,98 Preventing water clustering in the SBU neighbor-
BDC, which is different from the metal−ligand bond strength hood was achieved by alkyl functionalization.333 Actually, even
order, i.e., Al-BDC > Cr-BDC > V-BDC. Moreover, the metal polar groups such as amines and alkoxy groups draw water
ions in all three MOFs had the same oxidation state and toward the new sites, thus precluding water clustering near the
coordination geometry, ruling out the effect of such parameters SBU.334 Using water-repellent fluorinated ligands blocked
on the observed trend. The authors concluded that the stability water adsorption altogether.335,336 Incorporation of polymers
of these MOFs depends on the inertness or lability order of the can also increase the MOF hydrophobicity and thus their
metallic ions, i.e., CrIII > AlIII > VIII > VIV. In another example, hydrolytic stability.337,338 In addition to hydrophobicity, steric
the Walton group330 carried out partial substitution of Mg in factors may also reduce the rate of hydrolysis. This includes the
Mg-DOBDC with other metals such as Ni or Co, which are occurrence of well-protected metals with high coordination
less vulnerable to hydrolysis than Mg, and evaluated the numbers, such as Zr-based UiO-66339 and MIL-140,340
stability of the resultant materials under humid condition. After numerous porous coordination networks (PCNs),341−343 Hf-
exposure to water vapor at 95% RH, Mg-Ni-DOBDC, based DUTs,344 phosphonate-based MOFs,345−347 and
containing as little as 16% Ni, retained 80% of its surface others.348 Catenation, i.e., interpenetration of two or more
K https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

identical and independent frameworks, is another steric factor


that improves water stability by locking labile ligands in
place.349−352 The higher water-stability of Zn(BDC)(BPY)0.5
or MOF-508 (BDY = 4,4′-bipyridine) with catenation,
compared to Zn(BDC)(DABCO)0.5 or DMOF (DABCO:
1,4-diazabicyclo[2.2.2]octane) without catenation, offers an
illustrative example.349 Nonetheless, although the zinc paddle
wheel in the parent DMOF hydrolyzes readily, 319,353
functionalization of H2BDC linker by methyl groups increased
the kinetic water stability, particularly for the tetramethyl
BDC.354,355 This finding could not be rationalized solely by
increased hydrophobicity; the additional methyl groups had a
steric shielding effect on the SBUs.98,355 In another example,
MOFs, prepared from alkaline-earth Mg and Sr, and tetrakis[4-
(carboxyphenyl)oxamethyl]methane acid as ligand showed
significant structural collapse after 24 h exposure to 98% RH,
whereas the Ca-based counterpart was stable even after two
weeks of exposure.356 The superior stability of Ca-derived
MOF was attributed to a more subtle steric effect as the ligand
distortion within the framework gave rise to a steric
Figure 11. Structural changes upon hydration-dehydration of (left)
conformation of benzene rings around Ca2+ ions, shielding MIL-53(Cr), (right) MIL-53(Fe). Reproduced with permission from
them from water. ref 359. Copyright 2008 Royal Society of Chemistry.
Some MOFs such as Zr-based UiO-67 and NU-1000 are
stable toward linker hydrolysis, but collapse during dehy-
dration, because of the occurrence of capillary forces, strong experimental evidence, Yazaydin et al.99 found that the CO2
enough to bring about solid-phase compression, and ultimately uptake over HKUST-1 at 25 °C increased by 75% (from 2.3
structural collapse.357,358 Because these forces increase with mmol/g) at 0.2 bar and by 37% (from 11.3 mmol/g) at ca. 5
increasing liquid−vapor interfacial energy, one way to bar, when the material was prehydrated to 4 wt % (2.2 mmol/
circumvent the material collapse would be to exchange water g), corresponding to an average of 0.5 H2O per Cu. Further
with a more volatile solvent before drying.358 Freeze-drying or increase of the material moisture content to 8 wt % reduced
supercritical CO2 drying are other options. the CO2 uptake, presumably because of the presence of
In addition to the effect of water on the structural integrity uncoordinated water that blocked the pore space. Simulation
of MOFs, adsorption−desorption of water vapor may bring data revealed that stronger interaction between the quadrupole
about reversible changes in their pore structure, known as the moment of CO2 and the electric field created by adsorbed
breathing effect, with the most typical example being MIL-53 water is responsible for the enhanced CO2 uptake. Using
MOFs.102 Millange et al.359 reported that the breathing molecular modeling, Yu et al.362 also found that CO2
properties of MIL-53 is metal-dependent. Variable-temperature adsorption on HKUST-1 was promoted by small amounts of
PXRD data indicated that unlike MIL-53(Cr), whose pores both coordinated (preadsorbed) and uncoordinated (in gas
expand upon dehydration, the loss of water initially caused phase) water. Consistent with the results reported by Yazaydin
pore narrowing in MIL-53(Fe) structure, leading to the et al.,99 DFT calculations indicated that the presence of small
formation of a metastable intermediate anhydrous phase. amount of preadsorbed water increased CO2 uptake and CO2/
Further heating resulted in smaller opening of the structure N2 selectivity through higher binding energy between CO2 and
until a final anhydrous phase was obtained. Figure 11 illustrates water-coordinated CUS compared to direct CO2−CUS
the structural changes during dehydration of MIL-53(Fe) and interaction. Enhanced CO2/N2 selectivity by uncoordinated
MIL-53(Cr). water was attributed to competitive adsorption sites between
3.4. Effect of Water on CO2 Adsorption over MOFs
water and N2 in combination with higher CO2 uptake. On the
contrary, closed-system multicomponent equilibrium adsorp-
Numerous studies were devoted to the effect of water vapor or tion measurements over HKUST-1 at 40 °C in the presence of
preadsorbed water on CO2 capture by different MOFs.171−173 16% CO2/N2 with different humidity levels, indicated no
As discussed below, such effects depend on a multitude of major changes in the amount of adsorbed CO2 at a water
parameters, including the nature of metal centers, functional pressure of 1 and 4 mbar, corresponding to H2O uptake of 1.9
groups, framework structure, porosity, postsynthesis modifica- and 7.7 mmol/g.205 Using slightly different static H2O/CO2
tions, and the adsorption procedure. equilibrium, Liu et al.100 did find that up to 3.4 mmol/g, water
3.4.1. Effect of Coordinatively Unsaturated Sites. had a minor promoting effect on CO2 uptake but not as high as
Upon proper activation, many MOFs, including M-DOBDC, reported by Yazaydin et al.99 Nonetheless, at higher moisture
UiO-66, MIL-100, MIL-101, and HKUST-1 (Cu-BTC) exhibit content, no CO2 adsorption took place.100,204,205
metal sites with unsaturated coordination, referred to as CUS, In contrast to HKUST-1, interaction of water with CUS in
which represent preferential adsorption sites with applications M-DOBDC (M = Mg, Ni, Zn, and Co) gave rise to metal-
in adsorption, catalysis, and sensing.360 Despite the limited dependent adverse effects on CO2 uptake.100,205,363−366 Static
water-stability of HKUST-1,339,361 numerous studies were three-component equilibrium measurements indicated that the
devoted to the effect of water on CO2 adsorption over this CO2 uptake at about 0.2 bar, over Mg- and Ni-DOBDC at 40
MOF, starting with the intriguing promoting effect of moisture °C, decreased by more that 75% in the presence of H2O partial
at low content.99 On the basis of molecular simulations and pressure of 2 mbar (0.2%), well below practical applications.205
L https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Figure 12. CO2, N2, and H2O multicomponent equilibrium measurements for MIL-100(Fe) at (a) 25 °C and (b) 40 °C. Green circles represent
pure CO2 adsorption isotherms, and the black line is the corresponding dual-site Langmuir fit. Square data points represent the CO2 uptake from a
mixture of CO2, N2, and H2O (color of the square indicates the equilibrium partial pressure of H2O). Reproduced with permission from ref 205.
Copyright 2015 American Chemical Society.

Similar findings for Ni-DOBDC were reported based on binary capacity up to 1% water vapor (3.5% RH). The difference in
equilibrium data.100 Breakthrough measurements indicated binding energies between CO2 and CUS with and without
that the CO2 uptake of Ni-DOBDC in the presence of 15% coordinated water is a good indicator of the effect of water on
CO2/N2 at 25 °C of 3.74 mmol/g dropped to 2.2 mmol/g at CO2 adsorption on the corresponding MOFs.99,364,366 Hence,
RH as low as 3%.365 Even this residual uptake is most likely the higher binding energy between CO2 and water-coordinated
overestimated because the adsorption state did not reach Cu in HKUST-1 resulted in beneficial effect of water on CO2
equilibrium.205 PXRD data showed no loss of crystallinity uptake,99,362 whereas for M-DOBDC (M = Mg, Ni, Zn and
upon exposure to humidity. Moreover, no relationship was Co), lower binding energies led to the opposite effect, with
found between the loss of CO2 uptake and physical properties, varying intensity, depending on the nature of the metal.366
including metal atom electronegativity and metal−benzoate Soubeyrand-Lenoir et al.216 investigated the effect of
stability in aqueous medium, suggesting that kinetic and humidity on CO2 adsorption over MIL-100(Fe), another
thermodynamic parameters may be associated with the loss in MOF with uncoordinated metal sites, using breakthrough
CO2 capacity with moisture.365 In other studies on M- measurements after hydration of the adsorbent at the same
DOBDC, the focus was on steam conditioning and long-term relative humidity. At 20% and 40% RH, the CO2 uptake at 0.2
stability.363,364 Using breakthrough experiments in the bar was found to increase by an impressive 3 and 5 times,
presence of 20% CO2/N2 with 9−70% RH, Kizzie et al.363 respectively, compared to dry conditions. This was attributed
found that the CO2 capacity of Mg-DOBDC was similar to to a combination of water displacement and CO2 adsorption in
that under dry conditions. However, upon regeneration at 150 microporous pockets created by the partial filling of mesopores
°C, the material lost up to 81% of its initial CO2 capacity under by water. Similarly, using equimolar CO2/CH4, Xian et al.368
dry conditions. Similarly, after exposure to 20% CO2/N2 with found that at 50% RH, the CO2 uptake over MIL-100(Fe)
70% RH, and regeneration in N2 at 150 °C, M-DOBDC (M = increased from 1.49 to 3.75 mmol/g and CO2 versus N2
Ni, Zn and Co) adsorbed under dry conditions, 78%, 39%, and selectivity from 8.6 to 16.3. In contrast, the CO2 capacity of
15% less CO2, respectively, compared to pristine materials.363 MIL-100(Cr) decreased from 2.44 to 1.37 mmol/g. The
Similar findings were reported by Liu et al.364 Interestingly, beneficial effect of humidity was tentatively associated with the
PXRD showed that after steaming, the crystal structure of Mg- dissociation of water on MIL-100(Fe), forming new hydroxyl
DOBDC remained intact, whereas Ni-DOBDC broke down, adsorption sites for CO2. This contention was supported by
but both materials had drastically reduced surface areas. It was the occurrence of two CO2 desorption peaks in the TPD
surmised that the CO2 uptake of these MOFs was not profile in humid condition versus a single peak in dry
governed by crystallinity but by the density of their CUS. On condition. For MIL-101(Cr), however, the reduction of CO2
the basis of the standard reduction potential of Mg2+ and Ni2+ uptake was attributed to competitive adsorption of water.
of −2.37 versus −0.26 eV, respectively, it was inferred that Mg- Contrary to the aforementioned promoting effect of humidity
DOBDC would react more readily with oxygen or water than in the presence of MIL-100(Fe), static three-component
its Ni counterpart, leading to decreased density of CUS. This equilibrium measurements at 25 and 40 °C showed a rather
interpretation seems also to apply to Zn- and Co-DOBDC.364 slight decrease in CO2 uptake at 0.2 bar in the presence of
Smit and co-workers367 developed a methodology to derive humidity (Figure 12).205
improved force fields from periodic DFT calculations, enabling Al-containing MOFs were also considered for wet CO2
accurate prediction of CO2 and H2O adsorption isotherms adsorption.369,370 Prehydration of water-stable aluminum
over M-DOBDC. They found that the CO2 uptake at 0.15 bar fumarate in humid N2 (14% RH) had no effect on its CO2
over Mg-DOBDC at 40 °C decreased by 50% and 100% in the uptake of 2.1 mmol/g at 30 °C and 1.0 bar.369 As for CO2
presence of 0.2 and 0.5% water vapor (2.7 and 6.8% RH), uptake in the presence of 75% CO2/He with 14% RH, it
respectively. This finding was attributed to the strong binding decreased by only 17%, indicating a modest negative effect of
energy between CUS and water (70 kJ/mol) and the water. Further GCMC simulations showed that in the presence
substantial difference (>20 kJ/mol) in the isosteric heat of of Al-fumarate, gas species such as CO2, N2, and CH4 adsorb in
adsorption of water compared to CO2. However, Zn-DOBDC, the center of the pores, while water adsorbs in proximity of the
with lower binding strength for water, showed a constant CO2 Al-O-Al-O units in the form of clusters.370 Similar to HKUST-
M https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

1,99,363 at low concentrations, water was found to increase the of adsorption.384,387 Although information is scarce, there are
isosteric heat of CO2 adsorption, leading to higher CO2 indications that the effect of humidity on CO2 uptake by such
uptake; however, high moisture content led to the opposite materials is metal-dependent. In the presence of 0.25% CO2/
effect due to the accumulation of water clusters in the one- N2 with ca. 10% RH at 25 °C, the CO2 uptake over Zn-OH
dimensional Al(OH) fumarate channels. In the presence of rht- containing CFA-1 did not change (2.18 mmol/g) but
Type-MOF (Cu-TDPAT, TDPAT: 2,4,6-tris(3,5-dicarboxyl- decreased from 2.31 to 1.93 mmol/g for its Ni-OH
phenylamino)-1,3,5-triazine), although water adversely affected counterpart.387 Nonetheless, in the presence of 100% RH,
CO 2 adsorption, it increased CO 2 /CH 4 and CO 2 /N 2 the Zn-OH containing MOF did not exhibit any CO2 uptake.
selectivity.371 Masala et al.372 reported that another CUS- Interestingly, the breakthrough curves of MAF-X27ox in the
containing MOF, UTSA-16, retained 70% of its CO2 capacity presence of dry or humid (82% RH) 10:90 CO2/N2 were
at 40 °C, i.e., 0.94 versus 1.3 mmol/g, in the presence of humid almost the same, whereas the OH-free MAF-X27 completely
CO2 feed gas (9.8% CO2/1.7% H2O/N2; 23% RH), which was lost its ability to adsorb CO2 from the humid gas mixture.
attributed to its high density of CUS and low affinity for water. 3.4.3. Effect of Hydrophobicity. Adjusting the hydro-
3.4.2. Effect of Other SBU Features. Ibarra and co- phobicity is an excellent strategy to control the effect of water
workers made significant contributions highlighting the on CO2 uptake. A comprehensive summary on hydrophobic
beneficial effect of bridging hydroxo groups (μ2-OH) on MOFs can be found elsewhere.388,389 Some MOFs such as
humid CO2 uptake of hydrolytically stable MOFs such as zeolitic imidazolate frameworks (ZIFs) are inherently hydro-
NOTT-400,373 NOTT-401,198,374 InOF-1,196 CAU-100,375 phobic, and as a general rule, their CO2 uptake is insensitive to
and others.376 CO2 uptake was found to increase significantly moisture. The Yaghi group390 observed negligible difference
in the presence of humidity up to some level, depending on the between dry and humid (80% RH) CO2 uptakes over ZIF-300,
actual MOF, before it eventually decreased to zero. As an ZIF-301, and ZIF-302. Molecular simulations by Liu and
example, the uptake in the presence of pure CO2 at 30 °C over Smit391 also indicated that water did not affect the CO2 uptake
NOTT-400, increased by 100% and 150% in the presence of and CO2/N2 selectivity of ZIF-68 and ZIF-69 at low pressure;
10% and 20% RH compared to dry conditions.373 This was however, higher water vapor pressures resulted in competitive
attributed to the strong hydrogen bonding between water adsorption of water with decreasing CO2 and N2 uptake and an
molecules and μ2-OH, leading to the so-called confinement overall increase in CO2 versus N2 selectivity. GCMC and DFT
effect.377,378 It was suggested that such water molecules calculations also confirmed the behavior of ZIF-68.392 As for
homogeneously distributed inside the pores, providing addi- other MOFs, a straightforward method to achieve hydro-
tional adsorption sites for CO2. At higher RH of 35% and 60%, phobicity is to use fluorinated ligands or SBUs. Bhatt et al.393
no CO2 adsorption took place, presumably because of pore investigated the use of fluorinated NbOFFIVE-1-Ni for low-
blocking by water. Likewise, in the presence of water-stable concentration CO2 capture. NbOFFIVE-1-Ni showed excellent
NOTT-401,198 CO2 uptake at 30 °C increased by 3.2- and 1.6- CO2 uptake of 1.3 mmol/g at 400 ppm of CO2, with
fold in the presence of 5% and 10% RH, respectively, but was unprecedented selectivity attributable to the adsorption sites
completely inhibited beyond 30% RH. Notice that these being at the vicinity of the fluorine center. In the presence of
findings seem at odds with an earlier report on NOTT-401, dry 1% CO2/N2, the CO2 uptake was 1.86 mmol/g; however,
indicating a 7-fold increase in CO2 capacity at 40% RH.374 at 75% RH, although water capacity was 10 mmol/g, the CO2
With regard to the indium-containing MOF, InOF-1, also uptake was reduced only to 1.27 mmol/g. Nonetheless, the
possessing a binuclear [In2(μ-OH)] building block, similar amounts of adsorbed water and CO2 over multiple cycles were
promoting effect of humidity on CO2 adsorption was observed, constant, suggesting that no more competition took place.
with 2-fold increase at 20% RH.196 However, beyond 40% RH, Likewise, although AlFFIVE-1-Ni adsorbed a significant
there was significant decrease in CO2 uptake, also attributed to amount of water (11 mmol/g at 95% RH and 25 °C), single
pore blocking by water. The humidity-induced CO2 adsorption breakthrough experiments using humid 5/5/90:H2S/CO2/
enhancement over CAU-10, another μ2-OH containing MOF, CH4 with 65% RH, showed little effect of water on CO2 uptake
was 30% in the presence of 3−20% RH.375 Similar behavior because water adsorption was negligible by the time CO2
was also reported for MOFs containing [Al2(μ2-OH)] and adsorption was completed.394 Moghadam et al.395 investigated
[Mg3(μ3-OH)]5+ units such as MIL-53(Al)-YDC379 and Mg- CO2 and water coadsorption on a flexible hydrophobic
CUK-1,380,381 respectively. The effect of spatial confinement fluorinated MOF (FMOF-1), consisting of a network of
and surface interaction of water in the pore volume was also channels and small pockets decorated with −CF3 groups.
reported for MIL-53(Al), containing AlO4(OH)2 octahedra, as Interestingly, GCMC simulations showed that the presence of
the presence of 20% RH resulted in 50% increase in CO2 humidity, even as high as 80% RH, did not affect the CO2
uptake at 30 °C, i.e., from 0.8 to 1.2 mmol/g.382,383 uptake, consistent with the hydrophobic nature of the material
MOFs containing monodentate hydroxide ligands (M-OH) (Figure 13).
with the ability to generate bicarbonate upon exposure to CO2, Yu and Balbuena396 found that because of the hydrophobic
much like carbonic anhydrase, represent a recent development nature of Br-functionalized UiO-66 and its low binding energy
in the field.384−387 Liao et al.384 synthesized MAF-X25ox and with water, its CO2 uptake was hardly affected by moisture,
MAF-X27ox by partial substitution of the corresponding whereas NH2- and OH-modified UiO-66 materials showed
metals Co2+ and Mn2+ by [CoIII(OH)]2+ and [MnIII(OH)]2+, strong interactions with water, with adverse effect on CO2
respectively. Likewise, Bien et al.386 used postsynthesis ligand uptake and CO2/N2 selectivity. Other studies also demon-
exchange to generate Zn-OH using Zn-OAc containing CFA-1. strated that functionalization of MOFs with groups that
They also synthesized bimetallic analogues of CFA-1 decrease their hydrophilic character, decreased water uptake,
containing Ni-OH and Co-OH by metal and ligand thereby increasing CO2 selectivity over water. Bae et al.397
exchanges.387 All such materials showed high affinity toward found that pyridine incorporation in Ni-DOBDC (Py-Ni-
CO2 at low pressure because of particularly high isosteric heat DOBDC) lowered water uptake significantly more than CO2,
N https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

een@PDMS20 was prehydrated for 6 h, with marginal


decrease in crystallinity. Using 1-een under dry optimum
TSA cycling, with adsorption at 70 °C and desorption at 100
°C, a steady CO2 working capacity of ca. 3.18 mmol/g, was
achieved over 600 cycles. However, under moist conditions
(15% CO2/3.75% H2O/He, 30% RH), the CO2 uptake over 5
cycles for 1-een was only 0.66 mmol/g, compared to 2.27
mmol/g for 1-een@PDMS20 (Figure 15), further supporting
the transfer of the hydrophobic character of PDMS to 1-een@
PDMS20.

Figure 13. Calculated CO2 and water uptake on FMOF-1.


Reproduced with permission from ref 395. Copyright 2017 Creative
Commons Attribution License.

decreasing the selectivity of water versus CO2 from 1844 to


308, most likely because pyridine coordinates to CUS at the
expense of water.
Coating MOFs with appropriate polymers is also a
promising approach to not only improve their hydrolytic
stability because of higher hydrophobicity, but to main-
tain398,399 or enhance400−402 their CO2 capacity under wet
conditions, Zhang et al.398 coated MOF-5, HKUST-1, and
ZnBT [Zn(BDC)(TED) 0.5 ] with polydimethysiloxane Figure 15. Cyclic CO2 adsorption capacity for (a) 1-een and (b) 1-
(PDMS) by vapor deposition at 235 °C and exposed the een@PDMS20 under wet condition (15% CO2/He, 30% RH).
pristine and coated materials to humid conditions; MOF-5 and Reproduced with permission from ref 400. Copyright 2019 Royal
ZnBT were exposed to air at 55% RH for 1 day, while Society of Chemistry.
HKUST-1 was treated in liquid water for 3 days. The CO2
uptake under dry condition at 25 °C and 1 atm was the same
for pristine and coated MOFs, suggesting that the coating did Coating NH2-MIL-125(Ti), ZIF-67, and HKUST-1 with a
not affect the CO2 binding sites nor their accessibility. silicone resin by solution immersion gave rise to stable
Following exposure to water, the CO2 capacity of all coated materials after exposure to liquid water for 5 days, while the
MOFs did not change significantly, while that of pristine pristine MOFs showed varying degree of structure collapse as
materials decreased by 80%, 75% and 98% for MOF-5, early as after 1 day.403 CO2 uptake of pristine MOFs at 25 °C
HKUST-1, and ZnBT, respectively. Hong group400 reported and 1 bar, decreased by at least 80% after water treatment,
that coating N-ethylethylenediamine-functionalized while their coated counterparts retained 76−86% of their
Mn2(dobpdc) (dobpdc4−: 4,4′-dioxidobiphenyl-3,3′-dicarbox- original capacity. Fernandez et al.399 coated MIL-101(Cr) and
ylate), denoted 1-een, with 20% hydrophobic PDMS (1-een@ Ni-DOBDC by immersion in a chloroform solution of Pluronic
PDMS20) significantly improved its water resistance. In the P123 and found that upon exposure to air at 95% RH, the
presence of 15% CO2/He, 1-een had the same CO2 capacity modified materials adsorbed ca. 50% and 60% less water than
under both dry and humid conditions (Figure 14a), whereas 1- the pristine MOFs. Moreover, the CO2 capacity of the water-
een@PDMS20 exhibited higher CO2 capacity under humid exposed materials at room temperature was similar to that of
condition (Figure 14b). Prehydration of hydrophilic 1-een over the unmodified MOFs, suggesting that coating enhanced the
an extended period adversely affected its CO2 uptake as hydrolytic stability of the materials without compromising its
diffusing water competed with CO2, resulting in over 40% CO2 capacity. Ding et al.402 performed in situ polymerization
decrease in CO2 capacity and almost complete loss in of 1,2-diethynylbenzene inside the pores of MOF-5 to afford a
crystallinity after 6 h of exposure to water vapor. On the hydrophobic polynaphthylene-containing material, PN@MOF-
contrary, only 10% loss in CO2 capacity was observed when 1- 5. Despite its significantly lower surface area, breakthrough

Figure 14. CO2 adsorption versus time in dry (15% CO2/He) and wet (15% CO2/He, 30% RH) conditions at 50 °C over (a) 1-een and (b) 1-
een@PDMS20. Reproduced with permission from ref 400. Copyright 2019 Royal Society of Chemistry.

O https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Table 4. CO2 Adsorption of 0.15 atm CO2 over Diamine-Appended Mg2(dobpdc) under Dry and Humid Conditions at 40 °C
CO2 uptake (mmol/g)
diamine RH (%) dry humid ref
a
ethylenediamine (en) 100 4.30 2.88 149
N-methylethylenediamine (men) 100 3.95 3.63a 149

N-ethylethylenediamine (een) 30 4.52 4.25a 150


50 4.52 4.59 150
412
70 4.36 4.70
100 3.18b 2.45b,c 400
100 4.11d 4.25d 400

N-(propyl)ethylenediamine (pen) 70 4.43 5.16 412

N-isopropylethylenediamine (ipen) 50 3.09 2.90 150


70 3.09 3.23 150
100 4.11 4.86 412

N-(butyl)ethylenediamine (nBu-2) 70 4.14 4.82 412


N-(pentyl)ethylenediamine (nPen-2) 70 3.86 4.48 412
N-(3-pentyl)ethylenediamine (3-Pent-2) 70 3.39 4.50 412
N-(hexyl)ethylenediamine (nHex-2) 70 3.50 4.14 412
N-(cyclopentyl)ethylenediamine (cPent-2) 70 3.86 5.45 412

N,N′-dimethylethylenediamine (mmen) 27 2.73 2.54a 149


50 3.36 4.80 149
50 3.75 4.22 205
100 3.02 3.32 408

2,2-dimethyl-1,3-diaminopropane (dmpn) 27 3.00 2.70 153


2-(aminomethyl)piperidine (2-ampd)e 35 2.40 2.40 411
a
Sample was exposed to 100% RH for 48 h, then activated at 130 °C under Ar and exposed to 15% CO2/N2. bAdsorption at 70 °C. cMaterial was
exposed to 100% RH at room temperature for 1 h, then activated at 120 °C for 2 h under N2 and exposed to 15% CO2/N2. dAdsorption at 50 °C.
e
CO2 pressure of 0.04 atm

experiments using 16% CO2/N2 at 25 °C indicated that the hydrolytic stability, breakthrough data at 25 °C using 15%
CO2 capacity for PN@MOF-5 was 50% higher compared to CO2/N2 with 33−95% RH revealed up to 20% enhancement
MOF-5 because of stronger CO2 binding, presumably due to a in CO2 uptake compared to dry conditions, which was
more favorable pore size distribution with subnanometer pores attributed based on DFT calculations, to the significantly
and the occurrence of exposed aromatic surfaces. Under humid higher heat of CO2 adsorption in the presence of water.
condition (65% RH), only marginal decrease in CO2 capacity GCMC simulations indicated the occurrence of distinct sites at
was observed for PN@MOF-5, while MOF-5 displayed 40% the channel center for CO2 adsorption and at the corner for
and 73% loss in CO2 uptake for cycle one and two, H2O, which seems to be a prerequisite for the promoting effect
respectively. of water on CO2 uptake in zeolites280 and MOFs.406 Using
3.4.4. Effect of MOF Functionalization. In addition to functionalized IRMOF74-III with two primary amine groups,
earlier examples of ligand modification aiming at increasing Yaghi and co-workers155 found that unlike other amine-
hydrophobicity,393−396 other functionalization in view of containing adsorbents, typically adsorbing CO2 as ammonium
facilitating CO2 capture from humid streams were reported. carbamate, IRMOF74-III-(CH2NH2)2 adsorbed dry CO2
Nandi et al.404 designed IISERP MOFs using bifunctional primarily as carbamic acid, whereas the formation of
ligands comprised of basic azolyl and chelating groups in the ammonium carbamate increased with the humidity level,
presence of metal acetates, the latter providing a hydrophobic suggesting a change in the conformation of the MOF linkers
lining to the pore walls. This enabled selective adsorption of with humidity. Dynamic CO2 uptake of 1.2 mmol/g in the
CO2 even in the presence of 75% RH. Simulation data revealed presence of 16% CO2/N2 was obtained at 25 °C under dry and
the occurrence of polarized CO2 molecules adjacent to the humid (65% RH) conditions. It was suggested that under both
azolyl moieties, which in addition to strong van der Waal conditions, CO2 forms first carbamic acid, which undergoes
interactions between O(δ−) and C(δ+) led to the formation of proton transfer in the presence of water to generate
T-shaped CO2···CO2 configurations, thought to be at the ammonium carbamate.
origin of the CO2 affinity of the material under humid In addition to amine-functionalized ligands,155,396,407 the
conditions. Shi et al.405 synthesized derivatives of ZnF(TZ), a most common amine modification of MOFs is through
1,2,4-triazolate (TZ)-based MOF, using amino, diamino, and tethering to CUS. Although the behavior of amine-grafted
dimethyl-azolate linkers, i.e., ZnF(aTZ), ZnF(daTZ), and materials will be addressed in section 4.2, it is more
ZnF(dmTZ), respectively. In addition to their outstanding appropriate to discuss amine-tethered MOFs, within this
P https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Figure 16. Proposed mechanism for CO2 adsorption over mmen-M2(dobpdc) with four neighboring M−mmen sites: (a) nucleophilic addition of
CO2 and amine deprotonation, (b) carbamate rearrangement and destabilization of M−N bonds, and (c) formation of carbamate chains along the
pore channels. Reproduced with permission from ref 151. Copyright 2015 Springer Nature.

Figure 17. CO2 adsorption isotherms at 25 °C (blue), 40 °C (blue-violet), 50 °C (red-violet), and 75 °C (red) for (a) mmen-Co2(dobpdc) and
(b) mmen-Zn2(dobpdc). Reproduced with permission from ref 151. Copyright 2015 Springer Nature

section. The Hong148−150,408,409 and Long151−153,205,410−413 (mmen)-M2(dobpdc) with M = Co and Zn, are shown in
groups and others144−147,414−416 devoted much effort to Figure 17.151
understanding the effect of amine-appended MOFs for CO2 At 25 °C and 0.15 bar CO2, the CO2 capacity of
capture under dry and humid conditions. Amine tethering Mg2(dobpdc) was 4.85 mmol/g,410 and generally decreased
provides protection for electrophilic CUS, leading to less following functionalization with different dia-
competitive adsorption of water.144 With few exceptions, most mines148,150,410,417,418 due to a decrease in CUS density and
investigations dealt with grafting Mg2(dobpdc), with ethyl- surface area (Table 4). Nonetheless, the main advantage of
enediamines having a wide variety of alkyl groups, which are amine-tethered MOFs is their tolerance to moisture as
often highly efficient adsorbents for humid CO2 (Table 4). As opposed to pristine MOFs. After exposure to humid N2
shown in Figure 16, using typical N,N′-dimethyldiethylenedi- (100% RH) at 21 °C for up to 50 h, mmen-Mg2(dobpdc),149
amine (mmen)-appended M2(dobpdc), CO2 adsorption versus men-Mg 2 (dobpdc), 149 ipen-Mg 2 (dobpdc), 150 and een-
Mg2(dobpdc)150 showed marginal loss in crystallinity and
pressure takes place in three steps. At low pressure,
CO2 uptake, compared to en-Mg2(dobpdc)149 and the pristine
nucleophilic addition of CO2 occurs concurrently with the
MOF,148 which lost up to 65% in CO2 capacity, suggesting that
deprotonation of a metal-bound amine by an uncoordinated
MOFs-containing diamines with side alkyl groups have
amine of a neighboring mmen, acting as a strong base. This superior hydrolytic stability. As for CO2 adsorption in humid
leads to the formation of ammonium carbamate as in amine- gas mixture, the Hong group reported that for mmen-
containing silicas (Figure 16a).151,205 At an appropriate Mg2(dobpdc) exposed to 15% CO2/He with 51% RH at 40
pressure, carbamate rearrangement occurs (Figure 16b), and °C, the CO2 uptake was up to 43% higher compared to dry
the ion pair destabilizes the M−N bonds on adjacent metal conditions, with the material being stable over multiple
sites, leading to the formation of carbamate chains along the adsorption cycles.149,408 Under similar conditions, only
pore channels through a cooperative adsorption mechanism in marginal increase in CO2 uptake was obtained for een- and
a cascade manner (Figure 16c). This gives rise to CO2 ipen-Mg2(dobpdc) compared to dry conditions, i.e., 4.59
isotherms with abrupt steps,151−153,411 corresponding to a versus 4.52 mmol/g and 3.23 versus 3.09 mmol/g,
sudden rise in CO2 uptake as both ends of the diamine become respectively.150 Moreover, the rate of CO2 uptake was almost
involved in binding CO2. Such steps are both metal- and two times higher under humid compared to dry conditions, i.e.,
temperature-dependent. Typical CO2 adsorption isotherms for 1.27 versus 0.66 mmol/g·min for ipen-Mg2(dobpdc). In
Q https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Figure 18. CO2, N2, and H2O multicomponent equilibrium data at 40 °C for (a) mmen-Mg2(dobpdc) and (b) mmen-Ni2(dobpdc). Green circles
represent pure CO2 adsorption data. Squares represent CO2 uptake from a mixture of CO2 (166 mbar), H2O (up to 30 mbar), and N2 (679 mbar),
and the color of the square indicates the equilibrium partial pressure of H2O. Reproduced with permission from ref 205. Copyright 2015 American
Chemical Society.

Figure 19. Isobars under (a) dry and (b) humid (2.6% H2O, 0.7−36% RH) conditions at atmospheric pressure for pure CO2 (green), 4% CO2 in
N2 (purple), 0.4% CO2 in N2 (orange), and pure N2 (blue) in 2-ampd-Mg2(dobpdc). Solid lines depict adsorption (cooling), and dotted green
lines depict desorption (heating) for the pure CO2 isobars. The dashed horizontal lines indicate adsorption of 1 CO2 per diamine. (c)
Breakthrough data at 40 °C with 2-ampd-Mg2(dobpdc) using 4% CO2/N2 and ∼1.1 bar: overlay of dry and humid (third cycle) CO2 breakthrough
profiles. Reproduced with permission from ref 411. Copyright 2019 American Chemical Society.

another study, the amount of CO2 adsorbed at 14% CO2/N2 °C in 100% CO2 with 2% RH for 1 min indicated that dmpn-
with 32% RH and 17% CO2/N2 with ca. 32% RH was found to Mg2(dobpdc) exhibits a stable working capacity of ca. 2.8
be 4.2 and 1.1 mmol/g for mmen-Mg2(dobpdc) and mmen- mmol/g over 1000 cycles, with no amine leaching or material
Ni2(dobpdc), respectively.205 These capacities were 14% and degradation.
83% higher, respectively, compared to dry conditions (Figure Although diamine-appended Mg2(dobpdc) was the most
18), suggesting that H2O promotes CO2 adsorption over these investigated, other MOFs were also tethered with diamines,
hybrid materials as opposed to the pristine MOFs, which lost including M2(dobpdc), with M = Mn, Fe, Co, Ni, and Zn.151
up to 95% of their CO2 capacity in the presence of 20% CO2/ Impressively, mmen-CuBTTri (H3BTTri: 1,3,5-tri(1H-1,2,3-
N2 with 26−32% RH. Water capacity up to 9.5 mmol/g was triazol-4-yl)benzene) produced 250% increase in CO2 uptake
obtained for mmen-Mg 2(dobpdc), suggesting that the at 0.15 bar and 25 °C compared to the pristine MOF413 and
adsorption sites for water and CO2 are different.205 only minimal decrease in CO2 uptake when exposed to 19%
Comparison of CO2 isobars in the presence of 2-ampd- CO2/N2 at 40 °C with 31% and 43% RH compared to dry
Mg2(dobpdc), showed significant enhancement in CO2 uptake conditions.205 Other examples include een-Mn2(dobpdc)400
in humid (Figure 19b) versus dry (Figure 19a) conditions, and piperazine (ppz)-Mg2(dondc) (dondc4−: 1,5-dioxido-2,6-
notably at 40 °C and low CO2 concentration, e.g., orange naphthalenedicarboxylate).409
curves.411 In addition, column breakthrough data at 40 °C In addition to ethylenediamines, MOFs were also tethered
showed no CO2 slip, with a significant sharp breakthrough with other amines. Liao et al.147 appended hydrazine into
profile even after three cycles (Figure 19c). It was suggested Mg2(dobdc) to achieve high amine loading of ca. 6 mmol/g.
that under humid conditions, hydrogen-bonding interactions Surprisingly, water vapor (82% RH) had no effect on CO2
between carbamate and water led to enhanced CO 2 uptake at 40 °C regardless of the CO2 concentration in the
adsorption, notably at low CO2 concentration. For other feed (10% or 400 ppm). In contrast, TEPA-appended
diamines, such as dmpn-Mg2(dobpdc), only marginally higher Mg2(dobdc) exhibited higher CO2 uptake under humid
CO2 uptake was obtained at 40 °C under humid (15% CO2/ condition.146 In the presence of 15% CO2/N2 at 60 °C, the
N2, 27% RH) compared to dry condition, i.e., 3.0 versus 2.9 adsorption capacity increased from 6.06 to 8. 31 mmol/g at ca.
mmol/g.153 Cycling with adsorption at 40 °C in the presence 16% RH. In a more recent contribution, Kim et al.419
of 15% CO2/N2 with 27% RH for 5 min and desorption at 100 investigated a series of tetraamine-appended Mg2(dobpdc)
R https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

adsorbents for the removal of CO 2 , mostly at 4% discussed separately for easier understanding of the key
concentration, relevant to the treatment of natural gas conclusions.
combined-cycle emissions. In particular, the material appended Zaworotko and co-workers investigated the effect of water
with 3-3 [bis(3-aminopropyl)amine] referred to as on CO2 adsorption over zeolite 13X, tetraethylenepentamine
Mg2(dobpdc)(3-4-3) was tested under humid condition in (TEPA)-impregnated SBA-15, and a total of 13 microporous
the presence of 4% CO2/2.6% H2O/N2 at 100 °C (2.6% RH) and ultramicroporous MOF physisorbents, as shown in Figure
for 1000 cycles with desorption at 180 °C under humid CO2, 22.272,422 Gas mixtures with CO2 to H2O ratios from 1% to
or for 15 cycles with regeneration at 120 °C under simulated 100% and CO2 concentrations from 400 ppm to 100% were
steam, ca. 60% H2O/N2. In both cases, the material was found used. With the exception of TEPA/SBA-15, water had an
to be stable in terms of both amine content and adsorption adverse effect on CO2 uptake over all other materials,
performance. presumably because of competitive adsorption. Mason et
3.4.5. Effect of Framework Flexibility. MIL-53(Al and al.205 also concluded that among the 15 different adsorbents
Cr) are known to undergo reversible structural changes from they investigated, including 13X and 5A zeolites, AX-21 AC,
closed to open pore structure upon hydration−dehydra- and numerous MOFs and amine-appended MOFs, only amine-
tion.359,420 Adsorbed water within the pores leads to functionalized materials showed any significant CO2 uptake in
hydrogen-bonding with carboxylic acid ligands, as well as μ2- the presence of water partial pressures that are relevant to CO2
OH groups,103,420 bringing about the deformation of the pore capture processes. Among physical adsorbents, everything
system (Figure 20). As illustrated in Figure 21a, upon exposure being equal, the severity of water competition was dependent
on the nature of the material and H2O to CO2 ratio.272,422 For
example, compared to dry conditions, the CO2 capacity of
SIFSIX-3-Ni, HKUST-1, Mg-DOBDC, and Zeolite 13X
decreased by 30%, 82%, 71%, and 81%, respectively, upon
exposure to 15% CO2/N2 with ca. 55% RH, whereas, in the
presence of pure CO2 with the same RH, the capacity of
HKUST-1 and Mg-DOBDC was only 53% and 60% lower,
while SIFSIX-3-Ni was hardly affected.272 The deleterious
effect of water on CO2 adsorption over HKUST-1 and Mg-
Figure 20. Breathing effect of MIL-53(Cr, Al) upon adsorption and DOBDC was attributed to the strong coordination of water by
desorption of water. Reproduced with permission from ref 103. CUS, as reported by others.363,364
Copyright 2006 John Wiley and Sons. As for the higher tolerance of SIFSIX-3-Ni to water, it was
ascribed to the strong electrostatic field generated by the
to CO2, dehydrated MIL-53(Cr) shrinks into its closed form, charged pillars in the material framework and the tight binding
with limited adsorption. At higher pressure of about 5−6 bar, sites within the micropores, leading to particularly strong
the CO2 uptake increases abruptly as a result of the pore interaction with polarizable species such as CO2. SIFSIX-3-Cu
structure reopening.421 Similarly, hydrated MIL-53(Cr) exhibited even better performance in DAC conditions (400
exhibits its closed structure in the presence of CO2 up to ca. ppm of CO2/N2, 49% RH) due to the higher isosteric heat of
11 bar, when gradual opening of the pores takes place, leading CO2 adsorption of 56 kJ/mol versus 51 kJ/mol for SIFSIX-3-
to a dramatic rise in CO2 uptake (Figure 21b). Interestingly, Ni.422 However, it was found to be inherently unstable at
hydration of MIL-53(Cr) reduced CH4 uptake even at high higher temperature and humidity (40 °C, 75% RH). Further
pressures, with ca. 0.2 mmol/g at 20 bar (Figure 21b), because optimization of SIFSIX series led to SIFSIX-18-Ni-β, with
of the repulsive interactions between preadsorbed water and exceptional trace-CO2 removal capacity at ambient conditions,
methane, a nonpolar molecule with significantly lower excellent hydrolytic stability, and lower water capacity than
quadrupole moment than CO2.103 hydrophobic ZIF-8.423 In the presence of dry 1000 ppm of
3.4.6. Comparative Studies. Several comparative studies CO2, the capacity of SIFSIX-18-Ni-β was found to be 0.7
dealt with the effect of water on CO2 adsorption over large versus 1.6 mmol/g for the best performing material, TIFSIX-3-
numbers of different materials, mostly MOFs. These are Ni.424 However, under humid condition (74% RH), the CO2

Figure 21. Adsorption isotherms of CO2 and CH4 at 31 °C on MIL-53 (Cr): (a) before and (b) after hydration. Reproduced with permission from
ref 103. Copyright 2006 Wiley and Sons.

S https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Figure 22. Representative structures of adsorbents investigated by Zaworotko and co-workers. (a) Reproduced with permission from ref 272.
Copyright 2015 Wiley and Sons. (b) Reproduced with permission from ref 422. Copyright 2017 Royal Society.

uptake of SIFSIX-18-Ni-β was 0.3 mmol/g, higher than the top


performing MOFs, NbOFFIVE-1-Ni, SIFSIX-3-Ni, and
TIFSIX-3-Ni. Molecular simulations revealed that unlike
CO2−CUS interactions, typical of MOFs, SIFSIX-18-Ni-β
interacts with CO2 like an enzyme, through C···F interactions
between CO2 and SIFSIX units and O···H−C interactions
between CO2 and methyl groups, thereby simultaneously
enhancing its CO2 affinity and hydrophobicity.423
Eddaoudi group investigated other SIFSIX MOFs as
potential materials for CO2 capture under humid condi-
tions.425,426 SIFSIX-Cu-i, an interpenetrated polymorph of
SIFSIX-2-Cu, exhibited only a marginal decrease in the CO2
uptake and selectivity under humid relative to dry conditions
(CO2/H2:30/70) (i.e., 1.61 mmol/g and 191 at 74% RH vs Figure 23. Schematic representation of the effect of preadsorbed
1.99 mmol/g and 237 at 0% RH), while its PXRD spectra water on CO2 uptake for different groups of MOFs. Adapted from ref
showed a reversible phase change with hydration/dehydration 195.
cycles.425 Moreover, at low CO2 concentrations (1000 ppm),
switching from dry to humid condition (74% RH) over undertaken. The moisture content for each cycle was
SIFSIX-3-Cu and SIFSIX-3-Zn, did not significantly change calculated by comparing the weight at the end of the previous
CO2 selectivity over water.426 cycle to the weight of dry material. As shown in Figure 23, all
Llewellyn and co-workers195 conducted a comprehensive materials fell into five groups based on trends of CO2 uptake
screening study on the effect of preadsorbed water on CO2 versus preadsorbed water. The first group of materials (group
uptake over 45 MOFs, in the presence of 100% CO2 at 30 °C I) included hydrophobic MOFs, whose CO2 uptake was not
using TGA measurements (Figure 23). The data were affected by the level of preadsorbed water, which did not
compared with conventional adsorbents such as zeolites and exceed 4 wt %. This group was composed of ZIFs, ScBDCs,
ACs. In a typical experiment, the material was prehydrated in and MIL-121s families as well as MIL-69(Al) and MIL-
air at 100% RH at room temperature for 2 days, then treated in 103(La). The lack of water effect on CO2 adsorption over ZIFs
N2 at 30 °C for 8 h to establish the initial moisture content. was also reported by others.390−392 Because most of these
Adsorption of pure CO2 under dry conditions was then carried MOFs contain rather small pores, e.g., MIL-69(Al) with one-
out, followed by regeneration for 90 min at a desired dimensional tunnel of 2.7 × 19.4 Å2,427 the authors postulated
temperature to remove CO2 and decrease the moisture that CO2 and H2O may occupy distinct adsorption sites or
content, before a new adsorption−regeneration cycle was CO2 displaces adsorbed H2O.
T https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Group II included materials whose CO2 uptake decreased implications on CO2 capture from natural gas and biogas.428
linearly with increasing preadsorbed water, indicating that CO2 Molecular simulations dealing with a collection of 25 MOFs
and water compete for the same adsorption sites, akin to the indicated that for materials with weak interactions with water,
behavior of the reference zeolites (13X and Chabazite-Cs). the latter moves freely throughout the materials with little
Some MOFs in this group, including MIL-102(Cr) and Ni- effect on CO2 adsorption. However, when the interaction is
DOBDC, have CUS, where water adsorbs more favorably than sufficiently strong, water may either compete with CO2
CO2. The third group (group III) comprised of MIL-100(Cr), adsorption or generate additional adsorption sites for CO2.
MIL-101(Cr), MIL-130(Al), and MIL-68(Ga), exhibited Therefore, strong water−framework interactions can lead to
slightly decreasing CO2 adsorption capacity as the amount of increased or decreased CO2/CH4 selectivity.428 Similar
preadsorbed moisture increased, albeit with different slopes. arguments can be made regarding CO2/N2 selectivity, relevant
Group IV, containing MOFs with flexible structures, showed a to CO2 capture from flue gas.
plateau region for low moisture content combined with The Henry constant for water adsorption was also used by
decreasing trend in CO2 uptake with increasing preadsorbed Boyd et al.406 as an important criteria for the selection and
water. In particular, for MIL-53(Cr) and MIL-53(Al), the design of appropriate MOFs for wet flue gas CO2 capture.
plateau was followed by a strongly decreasing trend, suggesting MOFs were selected from a library of 325 000 hypothetical
dynamic changes in the MOFs’ pore size, which shrinks with candidates by (i) screening each material for its CO2 working
increasing amount of preadsorbed water. The strong capacity and CO2/N2 selectivity, (ii) selecting more than 8000
adsorption of water in the narrow pores, due to the materials with CO2 working capacity greater than 2 mmol/g
confinement effect,377,378 was at the origin of the significant and CO2/N2 selectivity greater than 50, i.e., better than zeolite
decrease in CO2 uptake. The last group (group V), comprised 13X under dry conditions, (iii) characterizing the structure of
of MIL-110(Al), MIL-163(Zr), HKUST-1, and UiO-66(Zr), the three most common CO2 adsorption sites, referred to as
UiO-66(Zr)-nCF3 (n = 1 or 2), UiO-66(Zr)-nMe (n = 1 or 2) adsorbaphore, and (iv) calculating the Henry constant for
and UiO-66(Zr)-2Et, showed a weak relationship between water adsorption over these high-performing materials.
preadsorbed water and CO2 uptake, with a maximum at Materials selected based on criteria (ii) and (iii) with low
particular water loadings. The presence of optimum amount of water Henry constant would be suitable for wet flue gas CO2
preadsorbed water for HKUST-1 is in agreement with other capture. Among the three identified CO2 adsorbaphores, the
reports.99 one consisting of parallel aromatic rings as depicted in Figure
To rationalize the often adverse effect of water on CO2 25 provides a near optimum interaction with all three atoms of
uptake over MOFs, the Llewellyn group195 used water
adsorption data for 33 materials to calculate the corresponding
Henry law constants (KH), which reflect the strength of water
interaction with the materials. As shown in Figure 24, almost

Figure 25. CO2 adsorbaphore containing parallel aromatic rings as


identified by computational screening. Reproduced with permission
from ref 406. Copyright 2019 Springer Nature.

CO2, with lower affinity toward N2 and H2O. Moreover,


because adsorption of water at high partial pressures is
Figure 24. CO2 capacity loss versus KH for water adsorption on dominated by the energetics of hydrogen bond formation,
different MOFs. Reproduced with permission from ref 195. Copyright materials with high CO2 uptake at high humidity should be
2017 John Wiley and Sons. endowed with pore shapes that frustrate the formation of
hydrogen bond network and limit the water binding energy.
On the basis of these findings, the authors designed,
all the data fitted a single semilog correlation between CO2 synthesized, and successfully tested two new MOFs, namely
uptake loss and KH, indicating that the stronger the water Al-PMOF and Al-PyrMOF, containing one-dimensional rods
adsorption, the larger the CO2 uptake decline. Despite the of AlIII linked by TCPP (tetrakis(4-carboxyphenyl)porphyrin)
presence of few outliers, the correlation indicates that CO2 and TBAPy (1,3,6,8-tetrakis(p-benzoic acid)pyrene) ligands,
capture performance of many MOFs in the presence of water respectively (Figure 26). Not only both materials were found
can be predicted based on water adsorption data. The binding to be water-stable, but their high CO2 uptake was hardly
strength of water in MOFs plays a determining role in affected by moisture. As anticipated, Rietveld refinement and
controlling the CO2 selectivity versus, e.g., CH4, with direct Fourier analysis of PXRD data for CO2-loaded materials
U https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

DOBDC is severely hampered with increasing amounts of


adsorbed water. For others, such as HKUST-1, MIL-53(Cr or
Al), NOTT400 and 401, InOF-1, and CAU-100, the CO2
uptake initially improves then decreases with increasing water
loading. For hydrophobic MOFs, including ZIFs and MIL-
121s families, adsorbed water does not have any significant
effect on CO2 uptake. Generally, the stronger the water
adsorption as reflected by higher KH, the more severe the CO2
adsorption loss. Understandably, for any given MOF, the
adverse effect of water is likely to be stronger at lower CO2 to
water ratio as typically observed for direct air capture.
Several investigations focused on the CO2 uptake versus
preadsorbed water. Although such studies are very helpful in
determining the general behavior of different MOFs, the
observed trends may not be representative for the behavior of
the material in actual CO2 capture from humid streams. During
Figure 26. Structural representation of (a) Al-PMOF and (b) CO2 adsorption on prehydrated materials, the moisture
AlPyrMOF showing the adsorbaphore contained in 3D non-
content may not remain constant in the presence of dry
interpenetrated structures (red box), originating from the orientation
of the tetracarboxylate ligands around the Al3+ rods. Reproduced with CO2-containing streams. This may lead to underestimated
permission from ref 406. Copyright 2019 Springer Nature. CO2 uptake, particularly when only gravimetric methods are
used to monitor adsorption. Moreover, during water and CO2
indicated that CO2 adsorption occurs preferentially in the coadsorption, the presence of other species in the gas feed may
adsorbaphore described above, whereas 13C nuclear magnetic also affect the working capacity of CO2. Therefore, the effect of
resonance (NMR) data showed that (i) the adsorbaphore was water coadsorption on CO2 uptake should ideally be assessed
not a preferential site for water adsorption, and (ii) even at by exposing MOFs to gas mixtures with different moisture
high humidity level, water had limited effect on CO 2 contents and monitor both H2O and CO2 independently.
adsorption. Table 5 summarizes CO2 uptake data for MOFs under dry and
Furthermore, Llewellyn and co-workers195 investigated the humid conditions, reported in the literature.
relationship between the CO2 uptake versus preadsorbed water
and the dipole−quadrupole interactions between water and 4. AMINE-FUNCTIONALIZED ADSORBENTS
CO2 molecules and generated energy maps for different CO2 scrubbing in aqueous amine solutions has long been
orientations between the two molecules. It was found that practiced for commercial-scale CO2 separation from gas
depending on the alignment of the water dipole moment and mixtures, particularly for purification purposes such as natural
the CO2 quadrupole moment, the interactions between two gas sweetening and hydrogen purification.431 Although several
adsorbates can be attractive, neutral, or repulsive. Therefore, alkanolamine solutions have been used, monoethanolamine
adsorbed water molecules can block the CO2 adsorption sites (MEA), diethanolamine (DEA), and methyldiethanolamine
or be displaced by CO2, depending on the orientation of the (MDEA) are the most popular for large-scale opera-
CO2 molecular axis with respect to the dipole moment of water tions.432−434 Nonetheless, despite its widespread application,
molecules. Consequently, the presence of adsorbed water may this technology suffers a number of drawbacks, including high
increase, decrease, or not affect CO2 uptake, depending on the energy consumption, excess foaming, corrosion, solvent loss,
actual MOF. amine degradation, and toxic emissions.22−28 The high energy
In summary, water stability of MOFs is an important factor demand of amine scrubbing processes is associated with the
when it comes to using them for carbon capture in a humid regeneration stage, which involves heating a relatively dilute
environment. From a thermodynamic point of view, MOFs aqueous solution, with high heat capacity. These shortcomings
with a combination of high charge density metals (e.g., Fe3+, often make large-scale amine scrubbing operations for the only
Al3+, Zr4+, and Ti4+) and carboxylate ligands (e.g., BDC and purpose of CO2 capture cost-prohibitive. Nonetheless,
BTC) or a combination of low-valent metals such as Zn2+, extensive efforts are being deployed to address some of the
Ni2+, Cu2+, and Mn2+ with weakly basic ligands, exhibit strong above-mentioned drawbacks. One breakthrough achievement
metal−ligand bonding and thus possess high hydrolytic is the development of KS-1, a proprietary sterically hindered
stability. Kinetics-wise, any factor that increases the activation amine with lower energy consumption and solvent circulation
energy of hydrolysis enhances the hydrolytic stability of MOFs, flow rate than typical 30 wt% MEA solutions.29,435,436 In
e.g., increased hydrophobicity through coating or fluorinated addition, the KS-1 solvent has better oxygen tolerance than
ligands and SBUs, catenation, and metal shielding from water conventional MEA, with practically no corrosion issues.
through high coordination number or steric effects. Although Nonetheless, shortcomings associated with amine solutions
water stability is a requisite for CO2 capture from humid can be circumvented when amine is immobilized on a solid
streams over MOFs, other considerations such as the strength support and utilized for CO2 capture via adsorption.61 In
of the water−MOF interactions, and the relative orientation of particular, because of the lower heat capacity of adsorbents
the dipolar and quadrupole moments of H2O and CO2, are of compared to water (ca. 1.5 versus 4.1 kJ/kg·K), the required
critical importance in promoting or blocking CO2 adsorption. energy for adsorbent regeneration is significantly lower than for
Depending on the nature of their metals, functional groups, stripping amine solutions.30,437,438
and structures, different MOF families exhibit different In the absence of moisture, CO2 adsorption by primary or
relationships between CO2 uptake and preadsorbed water. secondary amines occurs primarily via carbamate formation
Carbon capture of some MOFs such as MIL-102(Cr) and M- (eq 5), whereas, in the presence of humidity, interactions
V https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Table 5. CO2 Uptake Data for MOFs under Dry and Humid Conditions
CO2 uptake (mmol/g)
material CO2 pressure (atm) adsorption temperature (°C) RH (%) dry humid ref
Mg2(dobdc) 0.10 38 5 4.93 2.41a 364
0.10 25 90 3.94 1.39 278
0.14 45 1 2.13b 1.05b 429
0.14 45 10 2.13b 0.47b 429
0.15 30 100 5.34 1.54 272
0.17 25 9 5.36 1.79c 363
0.17 25 36 5.36 1.34c 363
0.17 25 70 5.36 0.85c 363
0.17 40 36 4.95 0.50 205
1.00 30 100 5.68 2.28 272

Co2(dobdc) 0.17 25 70 3.03 2.59c 363


Zn2(dobdc) 0.17 25 70 1.65 0.36c 363

Ni2(dobdc) 0.10 38 5 2.98 2.73c 364


0.10 25 90 2.38 1.86 278
0.15 25 3 3.74 2.69 365
0.17 25 70 2.86 1.74a 363
0.18 40 16 3.20 0.10 205
0.18 40 5 3.20 0.30 205

HKUST-1 0.15 30 100 1.59 0.29 272


0.18 40 37 0.67 0 205
0.20 25 3 0.91 0.80 216
0.20 25 10 0.91 1.18 216
0.20 25 20 0.91 0.93 216
0.20 25 40 0.91 0.45 216
0.20 25 3 2.3 4.03d 99
1.00 30 100 2.5 1.17 272
0.14 45 1 0.33b 0.25b 429
0.14 45 10 0.33b 0.18b 429

MIL-53(Al) 0.14 45 1 0.19b 0.14b 429


0.14 45 10 0.19b 0.02b 429
1.00 30 20 0.80 1.19 382

MIL-100(Fe) 0.20 25 3 0.5 0.59 216


0.20 25 10 0.5 0.50 216
0.20 25 20 0.5 1.50 216
0.20 25 40 0.5 2.39 216
0.20 40 31 0.67 0.30 205
0.50 30 50 1.49 3.75 368

MIL-100(Cr) 0.50 30 50 2.44 1.37 368


SIFSIX-3-Zn 0.10 25 90 1.49 1.43 278

SIFSIX-3-Ni 0.15 30 100 2.48 1.76 272


1.00 30 100 2.5 2.49 272

SIFSIX-3-Cu 0.10 25 90 2.03 1.90 278


0.15 30 75 51.4e 60.6e 422

SIFSIX-2-Cu-i 0.15 30 75 9.2e 15.6e 422


NOTT-400 1.00 30 20 0.95 2.32 373

NOTT-401 1.00 30 5 0.27 0.89 198


1.00 30 10 0.27 0.43 198
1.00 30 30 0.27 0.09 198
1.00 30 40 0.33 2.3 374

W https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Table 5. continued
CO2 uptake (mmol/g)
material CO2 pressure (atm) adsorption temperature (°C) RH (%) dry humid ref
198
1.00 30 5 0.27 0.89

ZnF(TZ) 0.15 25 53 0.29 0.30 405


0.15 25 99 0.29 0.33 405

ZnF(aTZ) 0.15 25 53 0.49 0.50 405


0.15 25 99 0.49 0.47 405

ZnF(daTZ) 0.15 25 53 0.90 0.99 405


0.15 25 99 0.90 0.88 405

MOF-177 0.14 45 1 0.02b 0.01b 429


0.14 45 10 0.02b 0.01b 429

IRMOF-1 0.14 45 1 0.05b 0.03b 429


IRMOF-1 0.14 45 10 0.05b 0.02b 429

ZIF-8 0.14 45 1 0.03b 0.03b 429


0.14 45 10 0.03b 0.02b 429
0.15 30 75 0.04 0.06 430

PCN-14 0.14 45 1 0.18b 0.13b 429


0.14 45 10 0.18b 0.01b 429

UiO-66(Zr) 0.14 45 1 0.08b 0.06b 429


0.14 45 10 0.08b 0.03b 429

UTSA-16 0.10 25 90 1.05 0.95 278


0.10 40 23 1.3 0.94 372

Zn-OH/CFA-1 0.25 25 10 2.18 2.18 387


0.25 25 100 2.18 0.00 387

Ni-OH/CFA-1 0.25 25 10 2.31 1.93 387

InOF-1 1.00 30 20 1.23 2.5 196


1.00 30 20 0.12 0.25 196

aluminum fumarate 0.75 30 14 1.71 1.45 369


1.00 30 14 2.1 2.1f 369

NbOFFIVE-1-Ni 0.01 25 75 1.86 1.27 393


CAU-10 1.00 30 5 1.23 1.66 375
MOOFOUR-1-Ni 0.15 30 75 1.32 0.89 430
DICRO-3-Ni-i 0.15 30 75 0.54 0.44 430
Ni-4-PyC 0.15 30 75 0.59 0.35 430
DMOF-1 0.15 30 75 0.39 0.22 430
UiO-66 0.15 30 75 0.72 0.19 430
Zn(pyrz)2(SiF6) 0.18 40 20 2.4 0.20 205

PCN-250(Fe3) 0.15 25 50 1.18 1.82 376


0.15 25 90 1.18 1.47 376

PCN-250(Fe2Co) 0.15 25 50 1.32 2.23 376


0.15 25 90 1.32 2.27 376
a
Adsorption of 10% CO2/N2 after conditioning the sample in 5% H2O/N2 at 100 °C for 2 h. bUptake in kmol/m3. cAdsorption in dry condition,
over samples that were exposed to simulated flue gas with the indicated RH, then regenerated in Ar at 150 °C. dThe material was prehydrated to 4
wt %. eUptake in L/kg. fAdsorption in dry condition, after prehydration in He with 14% RH at 30 °C.

X https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

between amine groups and CO2 lead, often partially, to the


formation of bicarbonate and carbonate (eq 6), alongside
carbamate.61,439,440 It is widely believed that the formation of
bicarbonate is associated with the hydrolysis of ammonium
carbamate (eq 7). Recent study, however, indicated that the
direct formation of bicarbonate according to eq 6 is more
favorable than the carbamate hydrolysis.441
2(RNH 2) + CO2 ⇋ RNHCO2−RNH3+ (5)

RNH 2 + CO2 + H 2O → RNH3+HCO3− (6)

RNHCO2−RNH3+ + H 2O ↔ RNH 2 + RNH3+HCO3−


(7)
Although ammonium carbamate forms only in the presence of
primary and secondary amines, carbonate and bicarbonate may
form regardless of the type of amine, including tertiary amines, Figure 27. Effect of water content in simulated flue gas on CO2
uptake over 50% PEI-impregnated MCM-41 at 75 °C. Reproduced
provided water is present.163 According to the stoichiometry of
with permission from ref 75. Copyright 2005 American Chemical
the reactions, the maximum amine efficiency, i.e., CO2/N ratio, Society.
in the absence of moisture is 0.5, and 1 in humid conditions.
Therefore, the presence of humidity often promotes CO2
uptake over supported amines through more favorable surface
chemistry.131 In addition to bicarbonate formation, DFT
calculations442 and experimental data 443 indicated the
formation of hydronium carbamate (R2NCOO−H3O+) during
humid CO2 adsorption on supported amines, which also
corresponds to a CO2/N stoichiometry of 1.
As mentioned earlier, the most common methods
for the preparation of supported amine adsorbents are
(i) impregnation of amines on porous materials such as
silica,107,112−114,117,197,201,213,444,445 alumina,446
activated carbon,121,447,448 MOFs,127 zeolites,448,449 and
clays,450 and (ii) surface grafting of amine-containing species
mostly on silica,76,129,136−138 but occasionally on alumina451
and other supports.452 Other procedures such as amine Figure 28. CO2 and water uptake over 2.0 g of 50% PEI-impregnated
tethering 144,145,148−153,205,408−416 and ligand modifica- MCM-41 at 75 °C using 13.55% CO2/3.86% O2/9.87% H2O/N2.
tion154−156,453 have also been used, particularly for MOFs. Reproduced with permission from ref 75. Copyright 2005 American
Accordingly, the effect of water vapor on the CO2 adsorption Chemical Society.
over these classes of materials will be discussed separately.
Note that amine-tethered MOFs have already been discussed
after 30 min seems to indicate the occurrence of a consecutive
in section 3.4.4.
carbamate to bicarbonate reaction according to eq 8. Water
4.1. Amine-Impregnated Adsorbents uptake decreased much more rapidly after 75 min but
Although impregnated polyethylenimine (PEI) CO2 adsorb- remained higher than the corresponding CO2 uptake.
ents were reported in the 1980s,454 they gained prominence
only since 2002, when Song’s group455 reported on CO2 RR′NCOO− + 2H 2O + CO2 ↔ RR′NH 2+ + 2HCO3−
adsorption over PEI-impregnated MCM-41 adsorbent, referred (8)
to as “molecular basket”. As for the effect of water, they found Many other reports confirmed the promoting effect of
that in the presence of 7.5% CO2/N2 with ca. 38% RH at 75 humidity on CO2 capture by PEI-impregnated materials. The
°C, CO2 uptake increased by 50%, demonstrating the extent of such promotion depends on the properties of the
promoting effect of water.113 Furthermore, using 14.9% adsorbent, including the PEI content and molecular weight,
CO2/4.25% O2/N2 with 0−15.5% H2O (up to 40% RH) at the porosity of the support, as well as the RH and CO2
75 °C, it was shown (Figure 27) that even though the CO2 concentration. Using 55 wt % PEI-silica monolith (Mw =
content in the feed gas was reduced gradually as the humidity 10 000) in the presence of 0.5% CO2/N2 at 25 °C, the CO2
level increased, the CO2 uptake increased significantly, uptake increased from 1.33 to 2.28 mmol/g at 90% RH.456
indicating the positive effect of humidity. It should be noted Likewise, Chen et al.201 reported that the uptake of a PEI-
that beyond a CO2/H2O ratio of 1 (ca. 13% H2O, 34% RH), impregnated hierarchical silica monolith in the presence of
further increase in humidity did not result in a noticeable pure CO2 at 60 °C increased from 4.8 to 5.9 mmol/g upon
increase in CO2 uptake.75 addition of 10% H2O (ca. 50% RH) to the feed stream. Using
Moreover, CO2 adsorbed quantitatively at the beginning of PEI on CARiACT silica or polymethylmethacrylate in the
the experiment (Figure 28), whereas water adsorption was not presence of 10% CO2/He at 45 °C, under dry and humid (73%
as effective, indicating that at this stage, CO2 adsorption was RH) conditions, Gray et al.219 found that the presence of
mainly governed by carbamate rather than bicarbonate humidity promoted CO2 uptake by ca. 45%, i.e., from about
formation.75 The significant increase in water adsorption 2.55 to 3.65 mmol/g. CO2 uptake of 50 wt % PEI-impregnated
Y https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

macroporous silica at 40 °C was also found to increase from


2.44 to 3.84 mmol/g when switching from dry to humid gas
mixture (15% CO2/3% H2O/N2, 41% RH).457 The CO2
uptake of PEI-impregnated clays at 45 °C also increased by
20% upon addition of 5% H2O (53% RH) to the 15%CO2/5%
O2/N2 feed gas.450 SBA-15 containing hyperbranched PEI
obtained by in situ polymerization of aziridine, with an amine
loading of 9.78 mmol N/g, also exhibited in the presence of
water-saturated 10% CO2/N2 at 25 °C an impressive CO2
uptake and amine efficiency of 5.55 mmol/g and 56%,
respectively.157,229 All improvements of CO2 uptake in the
presence of humidity were attributed to higher CO2 diffusion,
enhanced amine accessibility, and the partial formation of Figure 30. Effect of specific humidity on CO2 uptake over
bicarbonate alongside carbamate.99,100,201,219,456,457 prehumidified 44.8 wt % branched PEI on hydrophilic and
Other investigations used prehumidified materials to hydrophobic silica in the presence of 10% CO2/N2. Reproduced
delineate the effect of water on CO2 uptake.197,445,458 Using with permission from ref 458. Copyright 2017 Elsevier.
60 wt % PEI-impregnated SiO2 with 3%, 10%, and 15 wt %
preadsorbed water, in the presence of 15% CO2/N2 with 30% prehydrated adsorbent. Water uptake up to 100% RH was
RH at 70 °C, the Abu-Zahra group445 reported moderate found to increase linearly with humidity. A generalized model
improvement of less than 10% compared to the dry material. consisting of a rapid nucleation stage followed by a slow
Zhang et al.458 also investigated the effect of water on CO2 growth period provided the best fit for the adsorption rate of
uptake over a series of linear and branched PEI on water vapor. It is worth mentioning that during exposure to
hydrophobic and hydrophilic silicas at 25−100 °C. Materials moisture, PEI underwent several water-induced phase
were prehumidified in N2 with specific humidity up to 17.2 mg transitions to form hemihydrate, sesquihydrate, dehydrate,
H2O/g of adsorbent, then exposed to 10% CO2/N2 at the and anhydrate, consistent with earlier time-resolved infrared
same humidity. Notice that the authors used specific humidity measurements.459 Similar to water, adsorption of humid CO2
instead of the more common relative humidity. The promoting on prehydrated materials was also described by a two-step
effect of moisture was most pronounced at 25 °C, when the mechanism. The first rapid stage corresponded to surface
water adsorption was the highest. As shown in Figure 29, for adsorption, whereas the second slower step was associated with
bulk diffusion within PEI. Notice that increasing humidity led
to a decrease in the rate of CO2 adsorption for both fast and
slow stages. Further adsorption measurements on materials
prehydrated in humid nitrogen up to 16 wt % moisture, in the
presence of 10% CO2/N2 at 60 °C with the same RH as the
pretreatment nitrogen, showed an increase in CO2 uptake from
3.16 to 5.86 mmol/g.197 This enhancement was attributed to
possible carbamate hydrolysis in addition to direct bicarbonate
formation by reaction of PEI, water, and CO2.
Molecularly imprinted unsupported PEIs referred to as MIP-
PEI-Mw, with Mw = PEI molecular weight, were prepared as
shown in Figure 31 by bubbling CO2 through aqueous PEI

Figure 29. Effect of temperature on CO2 uptake over prehumidified


44.5 wt % linear PEI on hydrophilic silica in the presence of 10%
CO2/N2 with different specific humidities, as indicated. Reproduced
with permission from ref 458. Copyright 2017 Elsevier.

45 wt % PEI on hydrophilic fumed silica, beyond 50 °C,


humidity had little effect on CO2 uptake. Notice that at low
specific humidity, the CO2 uptake versus temperature
exhibited the usual bell shape,117,444 indicating that at low
temperature, adsorption was kinetically controlled because of
excessive diffusion limitation. At high specific humidity, CO2
uptake decreased over the whole temperature range, a clear
indication that water facilitated CO2 diffusion. Figure 30 shows
that the optimum effect corresponded to a specific humidity of
11.5 and 15.3 mg H2O/g for PEI-impregnated hydrophilic and
hydrophobic silicas, respectively.
In a comprehensive investigation, Monazam et al.197 studied Figure 31. Synthesis process for molecularly imprinted solid amine
the uptake and adsorption kinetics of water at 60 °C on PEI- adsorbent. Reproduced with permission from ref 461. Copyright 2018
impregnated CARiACT G10 and humid CO2 on the Royal Society of Chemistry.

Z https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Figure 32. (a) Column-breakthrough curves for CO2 adsorption on nOH-HBPA (n = OH to N ratio) and (b) cumulative CO2 adsorption capacity.
Reproduced with permission from ref 465. Copyright 2018 American Chemical Society.

solution, followed by cross-linking PEI chains using carbamate via an oxonium intermediate. Interestingly, higher
glutaraldehyde and reduction of the imine groups,460 or glycol CO2 capacities were obtained under humid conditions, due to
diglycidyl ether,461 and finally CO2 desorption at 90 °C. In the enhanced amine utilization efficiency (ca. 72%), but the
presence of 10% CO2/N2 at 25 °C, the CO2 uptake over dry opposite order was observed, i.e., 5.26 and 3.28 mmol/g for
MIP-PEI-70000 was only 0.61 mmol/g compared to 8.56 OH-free HBPA and 0.38-OH-HBPA, respectively. The
mmol/g for prehydrated (75 wt %) material.460 In addition to significantly higher CO2 capacity under humid conditions
bicarbonate formation, the remarkable enhancement of CO2 was attributed to the 1:1:1 reaction between amine:CO2:H2O
uptake was attributed to the swelling of MIP-PEIs (Figure 31) compared to 2:1 reaction between amine:CO2 under dry
in contact with water, facilitating the diffusion of CO2 deeper conditions.
into the material, whereas in dry conditions, PEI chains were TEPA-containing adsorbents were also investigated for
extensively coiled, limiting the reaction of CO2 to surface humid CO2 uptake.200,211,225 In the presence of 10% CO2/
amine groups. Koutsianos et al. 462 also used almost N2 with 12% and 37% RH at 60 °C, the CO2 uptake of 50 wt
unsupported PEI (91 wt % PEI/C60) and reported a 10-fold % TEPA-impregnated KIT-6 increased from 2.85 mmol/g to
increase in CO2 uptake after prehydration and exposure to 2.92 and 3.2 mmol/g, respectively, with no further change at
100% CO2 with 21% RH at 25 °C. Notice that in both higher RH.211 Liu et al.200 used 18 wt % TEPA-impregnated
examples, not only the PEI was basically unsupported, but the as-synthesized mesocellular silica foam, i.e., containing both
CO2 adsorption was carried out at 25 °C, explaining the very P123 triblock copolymer (template) and the trimethylbenzene
low uptake under dry conditions and the remarkable pore-expander. Although at 6.7% RH, the CO2 uptake did not
improvement in the presence of moisture.460−462 Likewise, change, increasing RH to 28% enhanced the CO2 uptake from
using a combination of PEI cross-linking in the presence of a 4.5 to 5.2 mmol/g.
diepoxide (bisphenol diglycidyl ether) and increased hydro- Earlier findings on the promoting effect of humidity on DAC
phobicity using epoxides with hydrophobic chains, Hamdy et over amine-impregnated materials were summarized in recent
al.463 found a significant increase in CO2 uptake in the reports.217,466−469 Sayari group reported a novel method for
presence of water, particularly at low temperatures. improving amine efficiency in DAC through higher PEI
Moreover, Qi et al.464 polymerized 2-methyl-2-oxazoline dispersion in a large-pore silica whose internal surface area was
onto propyliodide-grafted silica to afford a tethered linear covered with long carbon chains.217 In the presence of 400
polyethylenimine with high amine content (15.1 mmol N/g). ppm of CO2 with 64% RH, such materials achieved an
At 25 °C in the presence of 8% CO2/N2 with 18% RH, an impressive CO2 uptake of 7.31 mmol/g PEI, 33% higher than
impressive CO2 capacity of 11.8 mmol/g, corresponding to ca. under dry conditions. Using 70 wt % PEI (Mw = 1200) over
14-fold increase compared to dry conditions, alongside an mesoporous cellular silica foam for DAC, it was found that
amine efficiency of 80% was obtained. The observed addition of 2% H2O to 420 ppm of CO2/N2 at 58, 46, and 33
enhancement was attributed to the occurrence of less coiled °C (11−42% RH) led to 20%, 30%, and 53% increase in CO2
polymer chains, increased accessibility of amine sites and the uptake, respectively.470 This enhancement was associated with
formation of bicarbonate under wet conditions, as reported for improved accessibility of amine sites through decreasing
other materials with high polyamine content.460−462 Hydroxy- diffusion resistance and/or to the formation of bicarbonate.
functionalized hyperbranched polyamine adsorbents with Similar enhancement was reported for 72 wt % PEI on
different hydroxyl contents (nOH-HBPA, n = 0−0.38, hierarchical silica with 750 nm macropores and 5 nm
representing the molar ratio of hydroxyl to amino groups) mesopores at 30 °C in the presence of 10% CO2/N2. The
were also investigated for CO2 adsorption in dry and humid CO2 uptake was 2.34 mmol/g under dry condition versus 3.36
conditions.465 In the presence of dry 10% CO2/N2 at 25 °C, in the presence of 19% RH.471 Using 44 wt % PEI-impregnated
the lowest (1.08 mmol/g) and highest (1.60 mmol/g) CO2 nanofibrillated cellulose (NFC), Sehaqui et al.220 found that
uptake was obtained for HBPA and 0.38OH-HBPA, increasing RH from 20% to 80% in the presence of 400 ppm of
respectively (Figure 32), which was attributed to higher CO2 increased the uptake from 0.5 to 2.2 mmol/g, which was
amine utilization efficiency (15% for HBPA vs 35% for attributed to enhanced PEI dispersion and easier accessibility
0.38OH-HBPA) as OH groups acting as Bronsted base, to active amine sites by CO2 at higher RH. Similarly, exposing
assisted a 1:1 reaction of amine (1° or 2°) and CO2 to produce anhydrous cellulose-acetate silica fiber impregnated with 26 wt
AA https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Table 6. CO2 Uptake for Different Polyamine-Impregnated Adsorbents


CO2 uptake
(mmol/g)
support material amine (loading, wt %) CO2 content (%) adsorption temp (°C) RH (%) dry humid ref
Moderate to High CO2 Content
CARiACT G10 silica PEI (40) 1.2 40 28 2.5 2.75 199
1.2 80 5 1.1 1.11 199
10 40 97 2.55 3.65 29

LPEI (40) 10 60 15a 3.16 3.64 197


10 60 55a 3.16 4.58 197
10 60 100a 3.16 5.86 197

MCM-41 PEI (32) 14 40 28 1.7 2.5 205


PEI (50) 15 75 16 2.01 2.43 75
15 75 26 2.01 2.86 75
15 75 39 2.01 2.99 75,113

MCFb PEI (68.9) 15 75 0.6 1.18 0.86 202


15 105 0.8 1.83 2.09 202
50 75 2.6 2.50 2.98 202
50 105 0.8 3.37 3.04 202
50 115 0.6 3.00 2.61 202

LPEI (66.5) 15 105 0.8 0.28 0.28 202


50 75 2.6 3.20 3.34 202
50 105 0.8 1.41 1.23 202
50 115 0.6 0.59 0.57 202

MCFc PEI (71.4) 15 75 0.6 1.26 1.25 202


15 105 0.8 1.87 2.19 202
50 75 2.6 2.66 3.27 202
50 105 0.8 3.43 3.33 202
50 115 0.6 3.02 2.48 202

LPEI (68.4) 15 75 0.6 0.14 0.14 202


15 105 0.8 0.27 0.26 202
50 75 2.6 3.32 3.50 202
50 105 0.8 1.36 1.11 202
50 115 0.6 0.59 0.59 202

mesoporous silica PEI (60) 15 70 30 3.1 3.2d 445


15 70 30 3.1 3.4e 445
15 70 30 3.1 3.0f 445

macroporous silica PEI (50) 15 40 41 2.45 3.84 457


silica monolith PEI (55) 5 25 90 1.33 2.28 456
hierarchical silica monolith PEI (65) 100 75 26 4.80 5.90 201
PMMA PEI (40) 10 40 97 2.40 3.53 219
mesoporous carbon PEI (65) 15 75 80 4.82 5.36 473

C60 PEI (91) 100 25 21 0.04 0.18g 462


100 90 21 0.16 0.18g 462

KIT-6 TEPA (50) 10 70 12 2.85 2.92 211


10 70 37 2.85 3.20 211

SBA-15 and SiO2 gel TEPA (50) 15 75 60 5.05 5.86 225

as-synthesized MCF TEPA (18) 100 75 6.7 4.50 4.5 200


100 75 28 4.50 5.2 200

amorphous silica TEPA (9) 100 50 25 0.86 0.85 443

AB https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Table 6. continued
CO2 uptake
(mmol/g)
support material amine (loading, wt %) CO2 content (%) adsorption temp (°C) RH (%) dry humid ref
Moderate to High CO2 Content
TEPA (20) 100 50 25 1.71 1.73 443
TEPA (35) 100 50 25 2.86 4.03 443
PEI (51) 100 50 25 2.68 3.25 443

atainless steel disk 0.8 μm TEPA film 100 50 25 6.13 11.3 443
3.5 μm TEPA film 100 50 25 5.68 10.2 443
10 μm TEPA film 100 50 25 5.25 9.00 443

SBA-15 HA (7.0)h 10 25 100 NP 3.11 157


HA (3.67)h 10 25 51 NP 2.76 229
HA (7.28)h 10 25 51 NP 4.24 229
HA (9.87)h 10 25 51 NP 5.57 229
HA (2.24)h 10 75 51 NP 0.97 229
HA (3.67)h 10 75 51 NP 1.74 229
HA (7.28)h 10 75 51 NP 2.92 229
HA (9.87)h 10 75 51 NP 4.17 229
HA (3.67)h 10 25 51 NP 2.76 229

mesoporous silica LPEIi 8 25 18 0.85 11.8 464


SBA-15 DEA (20) + TEPA (30) 100 75 100a 3.70 3.18 191

AC DBNj (40) 10 29 100 0 0.86 474


29.3 100 0 0.86 474
38.9 100 0 0.63 474
40.5 100 0 0.62 474
50.3 58 0 0.52 474
Ultralow CO2 Content
mesoporous AC BPEI (55) 4 25 80 2.25 2.58 212
50 25 80 3.34 4.05 212

PE-MCM-41 with CTMA+ PEI (40) 4 25 64 2.20 2.92 217

HP2MGL resin PEI (50) 4 25 10 1.96 3.16 213


4 25 20 1.96 2.88 213
4 25 40 1.96 2.75 213
4 25 60 1.96 2.34 213
50 25 10 2.13 3.15 213
50 25 20 2.13 3.27 213
50 25 40 2.13 3.74 213
50 25 60 2.13 3.51 213

nanofibrillated cellulose PEI (19) 4 25 80 NP 0.25 220


PEI (31) 4 25 80 NP 1.20 220
PEI (44) 4 25 20 NP 0.50 220
PEI (44) 4 25 80 NP 2.22 220
PEI (52) 4 25 80 NP 2.10 220
PEI (62) 4 25 80 NP 1.75 220

cellulose-silica fiber PEI (10) 4 35 47 0.36 0.64 472


fumed silica PEI (33) 4 25 67 1.23 1.77 115

fumed silica PEI (50) 4 25 67 1.70 1.41 115


PAA (30) 50 25 65 0.29 0.27 224
PAAk (30) 50 10 65 NP 0.39 224
a
The sample was prehydrated with the indicated RH. bMesostructured cellular foam (MFC) with average pore diameter and pore volume of 12.6
nm and 2.25 cm3/g, respectively. cMFC with average pore diameter and pore volume of 31.4 nm and 2.50 cm3/g, respectively. dThe material was
prehydrated to 3 wt %. eThe material was prehydrated to 10 wt %. fThe material was prehydrated to 15 wt %. gThe material was prehydrated to 68

AC https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Table 6. continued
wt %. hHyperbranched amine with indicated amine content in mmol N/g. iAmine content of 15.1 mmol N/g. j1,5-Diazabicyclo[4.3.0]non-5-ene.
k
Polyallylamine. PEI = branched PEI. LPEI = linear branched PEI. CTMA+ = cetyltrimethylammonium cations. NP = not provided.

% PEI to dry 395 ppm of CO2/N2 at 35 °C led to CO2 respectively. It was only a decade later that more seminal work
breakthrough capacity of 0.59 mmol/g, compared to 1.6 on amine-grafted CO2 adsorbents was reported.76,132,135,157,214
mmol/g when prehydrated material was exposed to the same Huang et al.132 used propylamine-grafted MCM-48 silica with
feed with 47% RH.472 amine content of 2.3 mmol/g. In the presence of 5% CO2/N2
Wang et al.212 found that the effect of humidity on CO2 at 25 °C, the adsorption capacity was 1.14 mmol/g,
uptake at 25 °C over 55 wt % PEI-impregnated mesoporous corresponding to the quantitative formation of ammonium
carbon using 400 and 5000 ppm of CO2 under dry and humid carbamate (CO2/N = 0.5). Interestingly, under humid
(80% RH) conditions to be rather limited. Humidity increased conditions (100% RH), the CO2 uptake was doubled,
the CO2 uptake from 2.25 to 2.58 mmol/g and from 3.34 to consistent with the quantitative formation of ammonium
4.05 mmol/g for 400 and 5000 ppm, respectively. The role of bicarbonate and/or carbonate (CO2/N = 1). Nonetheless,
water over 50% PEI-impregnated HP2MGL resin was also amine efficiency as high as 0.5 and 1 under dry and humid
found to depend on the CO2 partial pressure, i.e., 400 vs 5000 conditions, respectively, have rarely been reported.76 In the
ppm.213 At 400 ppm of CO2, as RH increased from 0% to 10%, presence of moisture, the CO2 desorption temperature shifted
the uptake increased by more than 60%, from 1.96 to 3.16 ca. 10 °C upward, indicating that the presence of humidity
mmol/g, then declined to reach 2.4 mmol/g at 60% RH. For leads to stronger CO2−amine interactions.132
5000 ppm of CO2, increasing RH up to 40% led to significant Didas et al.477 investigated the effect of water on CO2
enhancement of CO2 uptake from 2.13 to 3.74 mmol/g. adsorption at ca. 1500 ppm over propylamine-grafted SBA-15
Further increase of RH to 60%, however, resulted in a slight with different amine contents. Interestingly, the beneficial
decrease in CO2 uptake, which was attributed to the impact of moisture was found to increase with decreasing
competitive adsorption of H2O over CO2 at high RH. Zerze amine loading, indicating that surface amine coverage beyond a
et al.224 reported that for poly(allylamine) impregnated fumed monolayer hinders the ability of CO2 to interact favorably with
silica, CO2 uptake from air under humid condition was water. Alternatively, the presence of surface silanols on
enhanced by preadsorbed water because it was readily available materials with low amine loading offers an additional route
for bicarbonate formation. In fact, conducting adsorption on a to CO2 adsorption through amine−silanol−water interaction.
dry adsorbent hinders water build-up, which is required for Consistently, FTIR data confirmed the formation of
efficient formation of bicarbonate. Moreover, water inside the bicarbonate only at low amine loading.
pores facilitates the unfolding of the polymer network, thereby The key role of silanol groups in facilitating proton transfer
enhancing the diffusion of CO2.221,224,458,460−462,464 reactions during CO2 adsorption over amine-grafted silica was
While all the above-mentioned studies on DAC reported addressed in a theoretical contribution by Cho et al.478 In
beneficial effect of humidity on CO2 uptake, some workers addition to intermolecular and water-assisted CO2 adsorption
reported that water vapor has an adverse effect.115,213 Using mechanisms, DFT calculations revealed the occurrence of a
PEI-impregnated fumed silica for DAC in dry and humid (67% surface hydroxyl-assisted mechanism or simply surface
RH) conditions, Goeppert et al.115 found that at low PEI mechanism (Figure 33), in which silanol groups act as proton
loading (33 wt %), the CO2 uptake increased from 1.18 to 1.77 transfer species. Under dry conditions, the activation energy
mmol/g when switching from dry to humid gas feed but for the surface mechanism (8.1 kcal/mol) was found to be
decreased from 1.70 to 1.41 mmol/g over a material with 50 wt lower than the intermolecular mechanism (12.7 kcal/mol),
% PEI loading. The explanation was somewhat tedious. leading to the dominance of the former mechanism over the
Considering that the material with higher PEI loading has
lower dispersion, the authors suggested that the detrimental
effect of water vapor on CO2 uptake was due to the water
blocking access to some of the already difficult to reach amine
groups.
As a general observation for amine-impregnated materials,
the presence of humidity improves CO2 uptake, regardless of
the feed CO2 content. Typical explanations involved the
formation of carbonate/bicarbonate in addition to carbamate,
as well as enhanced CO2 diffusion and amine accessibility.
Table 6 shows CO2 adsorption data for different amine-
impregnated adsorbents in dry and humid conditions.
4.2. Amine-Grafted Adsorbents
4.2.1. Amine-Grafted Mesoporous Silicas. Reports on
CO2 adsorption over grafted amines go back to the early
1990s.475,476 Leal et al.475 investigated CO2 adsorption over
propylamine-grafted amorphous silica and described the
formation of carbamate and bicarbonate under dry and Figure 33. Different reaction mechanisms for CO2 adsorption over
humid conditions, respectively. Nonetheless, even in the amine-grafted silica. The numbers indicate the corresponding
presence of pure CO2, the uptake was limited to 0.41 and activation energy in kcal/mol. Reproduced with permission from ref
0.89 mmol/g under dry and humid (100% RH) conditions, 478. Copyright 2018 Royal Society of Chemistry.

AD https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

latter. Nevertheless, during the water-assisted mechanism,


water also acts as a proton acceptor (Brønsted base) and the
corresponding activation energy is lower than for the surface
mechanism, i.e., 6.0 kcal/mol. Therefore, in the presence of
humidity, both the surface and water-assisted intermolecular
mechanisms may take place. Other investigations on CO2−
amine interactions indicated that the nucleophilic attack of
CO2 carbon by amine is assisted or catalyzed not only by
another amine but also by other species such as
water441−443,479 or OH groups,442,478 playing the role of a
Brønsted base to achieve the proton transfer/exchange
required for the formation of ions pairs such as ammonium
carbamate.
Hiyoshi et al.214,281 also used monoamine-, diamine-, and
triamine-grafted SBA-15 for CO2 adsorption at 60 °C in the
presence of 15% CO2/N2 under both dry and humid (60%
RH) conditions. Surprisingly, they found that for grafted Figure 34. Temperature programmed desorption−mass spectrometry
monoamine, moisture had no effect on CO2 uptake, while (TPD-MS) profiles of CO2 after adsorption on propylamine-grafted
PE-MCM-41 in the presence of different levels of humidity.
modest increases were obtained for the other materials, with Reproduced with permission from ref 131. Copyright 2008 American
the highest increase of 14% being associated with triamine- Chemical Society.
grafted SBA-15. In a similar study, Zheng and co-workers
found that using 2% water vapor (64% RH) in 15% CO2/N2 at
25 °C had no effect on CO2 capacity of diamine-grafted SBA- through bicarbonate formation. However, at RH of 74%, CO2
15.480 The authors suggested that contrary to monoamine uptake increased significantly by 60% to reach an amine
adsorbents, where a 1 to 1 reaction between amine and CO2 is efficiency of 0.78. Because water adsorption at 74% RH
possible in the presence of water, formation of stable involved capillary condensation, it was surmised that under
intramolecular ammonium carbamate between CO2 with such conditions, liquid-like water within the pores is more
diamine adsorbents is the preferred pathway with or without effective at promoting CO2 adsorption than at lower pressures.
moisture. Notice that in all adsorption tests under humid conditions, the
As mentioned earlier, tertiary amines do not react with CO2 amine efficiency increased but less than doubled, indicative of
under dry conditions, yet in the presence of moisture, they the limited nature of bicarbonate formation. Figure 34 also
contribute to CO2 adsorption via bicarbonate formation, as shows that at 74% RH, CO2 uptake over prehydrated material
evidenced by FTIR and 15N NMR data.163,481 Interestingly, in was about 10% higher than dry material, possibly because of
the presence of 1% CO2/He at 30 °C, humid CO2 (49% RH) more efficient formation of bicarbonate using water already
uptake over N,N-dimethyl-propylamine grafted SBA-15 was adsorbed. Moreover, CO2 desorption occurred at higher
lower in the presence of high versus low amine loading. This temperatures after adsorption of humid versus dry CO2/N2,
behavior was attributed to the lesser tendency of the former indicating that the presence of humidity results in stronger
material to adsorb water (i.e., 1.5 vs 3.9 mmol water/g), CO2−amine interactions, i.e., higher thermal stability of
necessary for the formation of bicarbonate. Similar surface bicarbonate compared to carbamate, as claimed by
density-dependence of amine−CO2−H2O interaction was also others.132,193
reported for propylamine-grafted silica.477 In a series of contributions,77,78,165,482 Sayari’s group
Sayari group made significant contributions to elucidating investigated the effect of water vapor on CO2 adsorption
the different effects of humidity on CO2 adsorption over over different types of grafted amines and impregnated
amine-functionalized silicas. First, they showed that contrary to polyamines, using TGA, infrared, and solid-state NMR
conventional CO2 adsorbents such as zeolites and many measurements to unravel the surface chemistry in the presence
carbons, amine-containing materials are tolerant to water vapor of water. Figure 35a shows that under cyclic adsorption−
and do not require predried feed gas.162 To do so, they used desorption experiments in dry conditions, in the presence of
diethanolamine (DEA) impregnated pore-expanded MCM-41 grafted monoamine and triamine and PEI-impregnated PE-
(PE-MCM-41) in the presence of dry and humid (27% RH) MCM-41, all adsorbents deactivated gradually at different rates
feeds containing 5% CO2 in N2. Note that at 5% CO2, depending on the severity of the adsorption−desorption
physisorption was found to be negligible, whereas the conditions and the nature of the amine.
chemisorption was almost complete; therefore, any effect on 13
C cross-polarization/magic angle spinning nuclear mag-
CO2/N ratio corresponds to actual changes in amine−CO2 netic resonance (CP MAS NMR) and diffuse reflectance FTIR
interactions.131 In the presence of 27% RH, CO2 uptake at 25 measurements after CO2 adsorption−desorption cycling under
°C increased from 2.65 to 2.94 mmol/g and the amine dry condition provided direct evidence for the formation of
efficiency (CO2/N) from 0.33 to 0.37, indicating that water stable urea linkages through dehydration involving CO2 and
vapor is not only tolerated but it promotes CO2 adsorption.76 two amine groups as shown in Scheme 1. Because the
In a more detailed investigation,131 they measured CO2 uptake formation of such linkages is irreversible under typical
over propylamine-grafted PE-MCM-41 at different RHs (27%, regeneration conditions, urea species accumulate over time at
61%, and 74%) using a combination of TGA and MS the expense of amines, giving rise to increasingly severe
measurements (Figure 34). Increasing RH from 0 to low deactivation. Interestingly, repeating all the experiments shown
(27%) and moderate (61%) levels resulted in modest increase in Figure 35a by bubbling the feed and purge gas in a water
in CO2 uptake by 16% and 22%, respectively, most likely saturator maintained at 20 °C, all other conditions being the
AE https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

from 74% at 25 °C to as low as 0.4% at 105 °C, and yet it was


enough to inhibit the formation of urea linkages and prevent
deactivation. Hence, humidity plays a key role in stabilizing any
supported amine against CO2-induced deactivation.77,78
Notwithstanding the wide-ranging scope of CO2-induced
deactivation, its severity depends on the nature of the amine
groups involved.79,165,482 In general terms, primary amine-
containing adsorbents deactivate more readily than others,
whereas in the absence of primary amines, secondary amines,
as in LPEI, deactivate more slowly, with single secondary
amines such as grafted N-methyl propylamine being excep-
tionally resistant to CO 2 -induced deactivation in dry
conditions. It was inferred that in the presence of primary
amines, the loss of water takes place via dehydration of
carbamic acid into isocyanate, which then reacts with a
neighboring primary or secondary amine to form open chain or
cyclic urea as shown in Scheme 1, route 1 in mechanisms A
and B, respectively. Although the loss of water from the
primary amine-derived ammonium carbamate (Scheme 1,
mechanism A, route 2) cannot be completely ruled out, DFT
calculations showed that the isocyanate mechanism provides
the lowest energy route to urea formation.483 As for single
secondary amines, based on their high stability under dry
conditions, and the fact that no isocyanate can be formed from
the corresponding carbamic acid, it was surmised that the
associated ammonium carbamate does not dehydrate under
typical dry CO2 adsorption conditions. Nonetheless, although
single secondary amines are highly stable against CO2-induced
deactivation, LPEI, which contain ethylenediamine units with
Figure 35. CO2 uptake of different adsorbents during cyclic only secondary amines, deactivates under dry condition, albeit
adsorption−desorption operations under (a) dry and (b) humid at a much lower rate than BPEI. It is thus inferred that the
conditions (20 °C dew point). Mono and TRI stand for grafted corresponding intramolecular ammonium carbamate dehy-
monoamine and triamine and PEI for impregnated PEI. X/Y-d and drates into tetrasubstituted imidazolidinone (Scheme 1,
X/Y-h indicate the adsorption and desorption temperatures under dry mechanism B, route 2).165,482
and humid conditions, respectively. Reproduced with permission from 4.2.2. Amine-Grafted Nanofibrillated Cellulose. Cli-
ref 77. Copyright 2010 American Chemical Society. meworks developed CO2 capture from ambient air using
amine-grafted NFC.221,484−486 Their first reported material,
same, showed that none of the materials deactivated under any diamine-grafted NFC exhibited a CO2 uptake and working
of the conditions used (Figure 35b). Notice that the RH varied capacity of 1.39 and ca. 0.70 mmol/g in the presence of 506

Scheme 1. Proposed Mechanisms for CO2-Induced Degradation of Supported Aminesa

a
Reproduced with permission from ref 165. Copyright 2012 American Chemical Society.

AF https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

ppm of CO2/Ar with 40% RH at 25 °C and desorption at 95


°C.484 The CO2 capture of monoamine-grafted NFC at 400
ppm of CO2 also increased in the presence of water (91% RH)
from 1.11 to 2.13 mmol/g at 23 °C and from 0.11 to 1.42 at 50
°C, with the effect of water decreasing at higher CO2
concentrations.221 This finding was attributed to the presence
of two different amine sites on the material, i.e., the ones that
are readily accessible to CO2 and those that are buried within
the bulk of the adsorbent, which can be reached only via
diffusion. In dilute CO2 streams, the amine sites within the
bulk were accessible only under humid conditions, with water
facilitating the diffusion of CO2 toward such sites, whereas, at
high CO2 pressure, both types of sites were accessible under
dry and humid conditions. Moreover, water was thought to Figure 36. CO2 and water uptake on Lewatit versus time under
contribute to the formation of diffusive intermediates as ambient air. Reproduced with permission from ref 203. Copyright
reported by others,442,462,487 thereby enhancing CO2 diffusion 2014 Elsevier.
and the accesibility of amine sites. This interpretation is
consistent with the oberved increase in amine efficiency at 25 CO2, (ii) amine catalyzed formation of carbamic acid, (iii)
°C from 0.26 under dry conditions to 0.51 at 91% RH. Under direct amine−water catalyzed formation of carbamic acid, and
otherwise the same conditions, the amine efficiency at 1 atm (iv) amine catalyzed formation of carbonic acid. DFT
CO2 and 25 °C increased only marginally from 0.54 to 0.60 calculations showed that reaction (ii) which does not involve
due to the much lower H2O/CO2 ratio. Using a similar water and reaction (iii) of water and CO2 catalyzed by amine,
material under temperature-vacuum swing (TVS) conditions, are both responsible for CO2 capture on Lewatit.
increasing RH from 20% to 80% was also found to increase The Lackner group468,491−493 developed moisture-swing
CO2 uptake at 20 °C from 0.39 to 0.65 mmol/g.485 At a given DAC materials, with water playing a key role in the CO2
RH, temperature had limited effect on CO2 and water uptakes. adsorption−desorption process. Such adsorbents consist of an
For example, CO2 (H2O) uptake at 40% RH was 0.42 (1.67), anion-exchange resin containing quaternary ammonium
0.44 (1.69), and 0.38 (1.74) mmol/g at 10, 20, and 30 °C, cations attached to a polymer network with hydroxide or
respectively. Column breakthrough data at 20 °C showed that carbonate anions as mobile counterions. As illustrated in
while water adsorption was complete between 0.5 and 4 h at Figure 37, for a fresh wet material (state 1), the more stable
20−80% RH, CO2 uptake in the presence of 455 ppm of CO2/
N2 did not reach equilibrium after 5 h at all RHs because of the
very low driving force.
4.2.3. Amine and Ammonium-Containing Polymers.
Amine-containing resins, including the highly hydrophilic
benzylamine-containing polymer Lewatit VP OC 1065 (or
Lewatit), were investigated for water and CO2 coadsorption.
Alesi Jr. and Kitchin474 found that exposure of prehydrated
(2.7 wt %) Lewatit to 11.1% CO2/N2 with 75% RH and 50 °C,
led to a CO2 capacity of 2.39 mmol/g, which was 30% higher
compared to dry conditions. The Brilman group carried out
extensive DAC investigations on Lewatit.203,226,227,488 Meas-
urements at 15−35 °C using 400 ppm of CO2/N2 with up to
60% RH showed that, while water adsorption was not affected
by CO2, higher CO2 uptake was recorded at increasing
humidity, reaching 47% increase at 15 °C and 60% RH.227
This was attributed to the formation of
bicarbonate, as reported for other amine-containing
materials.99,100,197,224,456,457,460,461,470,489 It was also reported
Figure 37. Moisture-swing sorbent for CO2 capture from ambient air
that the CO2 capacity of Lewatit in the presence of 500 ppm of over ion-exchanged resin. Reproduced with permission from ref 468.
CO2 at 20 °C increased by 216% at 79% RH compared to dry Copyright 2020 John Wiley and Sons.
conditions (0.6 vesus 1.3 mmol/g).488 Smal et al.203 exposed
Lewatit to ambient air for different periods of time up to 15 h
and monitored the effluent by TGA and IR spectroscopy (TG- carbonate anions are the dominant counterions. As the resin
FTIR) during regeneration of the material up to 130 °C. Water was dried, the carbonate evolved into bicarbonate and hydroxyl
adsorption was found to be dominant over CO2 in the early anion through reaction with water (state 2). Because of the
stages, mostly because of the high water to CO2 ratio (Figure occurrence of hydroxyl ions, state 2 had a strong affinity
36). After 450 min of air exposure, the water uptake began to toward CO2, even at very low concentrations. This led to state
decrease in favor of CO2, presumably because water is 3, with bicarbonate species as the dominant compensating
displaced by CO2 over time. As for the DAC mechanism anions. Wetting the material to full hydration led to state 4,
under humid condition over Lewatit, Buijs and de Flart490 used which evolved into state 1 with the release of CO2.
molecular simulations to investigate the following scenarios: Similar adsorbents were prepared by impregnating an amine-
(i) uncatalyzed formation of carbamic acid from an amine and containing polymer on carbon black and further quaternization
AG https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Table 7. CO2 Uptake for Different Amine-Grafted Adsorbents


CO2 uptake
(mmol/g)
support material amine amine loading (mmol N/g) CO2 content (%) adsorption temperature (°C) RH (%) dry humid ref
Moderate CO2 Content
SBA-15 APTESa 2.60 15 60 60 0.66 0.65 214
MAPSb 1.92 15 60 60 0.25 0.25 214

AEAPSc 2.50 10 25 51 NP 0.71 157


4.14 15 60 60 0.87 0.98 281
2.25 15 60 60 0.26 0.34 214
3.75 15 60 60 0.87 0.90 214
4.62 15 60 60 1.36 1.51 214
2.65 15 25 64 0.45 0.45 480

TRId 5.00 15 60 60 1.10 1.28 281


2.73 15 60 60 0.35 0.39 214
4.84 15 60 60 1.10 1.21 214
5.79 15 60 60 1.58 1.80 214

DMAPSe 1.90 1 30 49 0.01 0.10 163


2.50 1 30 49 0.01 0.06 163

MCM-41 APTES 4.2 18 40 30 1.7 1.2 205


MAPS 3.7 18 40 27 1.4 1.5 205

TRI 5.78 5 25 27 1.04 1.12 76


5.97 5 25 27 1.08 1.18 76
5.84 5 25 27 1.01 1.09 76
5.70 5 25 27 0.97 1.01 135

PE-MCM-41 TRI 6.06 5 25 27 1.51 1.60 76


6.10 5 25 27 1.55 1.66 76
6.02 5 25 27 1.47 1.58 76
5.97 5 25 27 1.41 1.52 135
7.74 5 25 27 2.33 2.58 76
7.98 5 25 27 2.65 2.94 76
6.75 5 25 27 1.89 2.08 76
6.65 5 25 27 1.69 1.85 76
4.18 5 25 27 2.05 2.37f 131
4.18 5 25 61 2.05 1.31g 131
4.18 5 25 74 2.05 1.90g 131
4.18 5 25 27 2.05 2.46 131
4.18 5 25 61 2.05 2.64h 131
4.18 5 25 74 NP 3.67h 131

MCM-48 APTES 2.30 5 25 64 1.14 2.30 132


2.74 15 60 60 0.52 0.56 281
1.11 15 60 60 0.15 0.10 214
2.58 15 60 60 0.52 0.50 281

NFC APDES 3.21i 0.1 23 91 1.72 2.29 221


3.21i 1 23 91 2.03 2.37 221
3.21i 10 23 91 2.26 2.54 221
3.21i 0.1 50 21 1.51 1.97 221
3.21i 1 50 21 1.81 2.29 221
3.21i 10 50 21 2.05 2.37 221

HMS APTES 1.92 90 20 30 0.86 0.91 193


AEAPS 3.08 90 20 30 0.89 0.45 193
TRI 3.85 90 20 30 1.20 0.98 194

AH https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Table 7. continued
CO2 uptake
(mmol/g)
support material amine amine loading (mmol N/g) CO2 content (%) adsorption temperature (°C) RH (%) dry humid ref
Ultralow CO2 Content
SBA-15 APTMSj 1.00 4 25 80 0.09 0.13 273

NFC APDES 3.21i 4 23 91 1.11 2.13 221


3.21i 4 50 21 0.11 1.42 221
3.86 4 10 20 NP 0.36 485
3.86 4 10 80 NP 0.63 485
3.86 5 20 20 NP 0.39 485
3.86 5 20 80 NP 0.65 485
3.86 5 30 20 NP 0.32 485
3.86 5 30 60 NP 0.50 485
a
(3-Aminopropyl)triethoxysilane. bN-Methylaminopropyltrimethoxysilane. cN-(2-Aminoethyl)-3-aminopropyltriethoxysilane. d(3-Trimethoxysilyl-
propyl)-diethylenetriamine. eN,N-Dimethyl-3-aminopropyltrimethoxysilane. fHumid CO2 adsorption was conducted on a dry material. gUptake is
in kmol/m3. hHumid CO2 adsorption was conducted on a prehydrated material. iTheoretical amine content. j3-aminopropyltrimethoxysilane.
HMS = hexagonal mesoporous silica. NFC = nanofibrillated cellulose. NP = not provided.

and ion-exchange of counterions with hydroxide.494 A cheaper


moisture-swing adsorbent was also developed from biomass-
derived amine-containing chitosan as starting material.495 The
biomass was cross-linked by poly(vinyl alcohol) in a three-
dimensional network to address the water solubility of chitosan
and increase the porosity of the material. The CO2 uptake
from air at 3% RH was 0.18 mmol/g at 20 °C. Exposing the
CO2-loaded material to 95% RH led to complete CO2
desorption, while decreasing the RH back to 3% led to a
100% CO2 readsorption. Table 7 shows selected CO2 uptakes
for different amine-grafted adsorbents in dry and humid
conditions.
4.2.4. Amine-Functionalized Zeolites. Although the
rational of combining zeolites and amines for the purpose of
CO 2 capture is questionable, several groups did
so. 123,279,496−501 Su et al. 123 conducted 15% CO 2 /N 2
adsorption on TEPA-modified Y zeolite in the presence of Figure 38. TPD-MS profiles of CO2 after adsorption on monoamine
different humidity levels at 60 °C. Increasing water content up impregnated zeolite β in dry and humid conditions. Reproduced with
to 17.4% (88% RH) led to an increase in CO2 uptake from permission from ref 496. Copyright 2016 Elsevier.
2.56 to 3.67 mmol/g with an optimum of 4.32 mmol/g in the
presence of 7% H2O (36% RH). Using APTES-impregnated
mmol/g over propylamine-functionalized hierarchical LTA
beta zeolite, Madden and Curtin496 found that addition of 2%
zeolite in the presence of 15% CO2/N2 at 60 °C decreased by
water to 15% CO2/He at 35 °C (36% RH) led to moderate
65% under humid conditions, which was attributed to the high
increase in CO2 uptake from 4.09 to 4.89 mmol/g. TPD-MS
affinity of LTA for water.
profiles of CO2-loaded adsorbent under humid condition
showed an increased amount of CO2 desorbing at 55, 69, and 4.3. Hydrophobic Amine-Containing Adsorbents
86 °C and a comparable amount at 98 °C (Figure 38).496 It As discussed earlier, water vapor promotes CO2 adsorption
was surmised that the carbamate decomposes at higher over amine-containing adsorbents and inhibits their CO2-
temperatures (98 °C) than bicarbonate. This finding is at induced degradation. Understandably, everything else being
odds with earlier reports on amine-containing mesoporous equal, adsorbed water increases the adsorbent regeneration
silicas, which indicated higher thermal stability of bicarbonate energy (vide infra). However, whether vapor or liquid, water
compared to carbamate.131,132,192 may have additional drawbacks, including decreasing amine
On the contrary, other groups found that CO2 adsorption on content and CO2 uptake. Because amines and polyamines are
amine-modified zeolites was adversely affected by moisture. mostly soluble in water, exposure of impregnated amine
Kim et al.279 reported that diamine-grafted Y zeolite exhibited adsorbents to liquid water, even at room temperature, leads to
a CO2 uptake of 1.9 mmol/g under dry flue gas (15% CO2/2% either amine leaching or surface segregation with increased
Ar/N2) compared to 1.6 mmol under humid conditions (41% diffusion resistance and pore blockage.457,502 Amine-grafted
RH), which was attributed to preferential water adsorption on adsorbents also undergo amine leaching or rearrangement via
the support over CO2. Interestingly, FTIR data showed that hydrolysis of the siloxane bridges, followed by dissolution or
due to strongly coadsorbed water on the zeolite support, not all regrafting, depending on whether liquid water is separated or
water was removed during desorption, which proved to be evaporated.138,503−505 In addition to leaching of active sites,
critical in obviating urea formation during cyclic operations. exposure to steam may lead to structural degradation of the
Nguyen et al.497 also reported that the CO2 uptake of 2.3 material, thereby inhibiting access to amine sites.506−508
AI https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Developing adsorbents with hydrophobic character is a than the added benefit associated with enhanced CO2 working
promising option to overcome these shortcomings. Such capacity.445,517,519 In practice, the net regeneration heat should
materials would reduce water uptake/working capacity, while take into account the thermal energy recovered from different
improving the stability of amine sites toward leaching and sources, such as the sensible heat from the adsorbent and
surface mobility. Several adsorbents, consisting of either a desorbed CO2 and the latent heat in moisture. Heat
hydrophobic support509,510 or hydrophobic organic moi- management is a major component in the design and
eties511,512 with improved behavior in the presence of optimization of adsorption processes.
moisture, were reported in the literature. Guo et al.509 used Several groups used eq 9 or closely related equations to
an amine-tethered polymer synthesized by copolymerization of calculate the regeneration heat, Qr (kJ/kgCO2) of amine-
glycidyl methacrylate, tert-butyl methacrylate, and trimethylol- containing adsorbents without considering water adsorption or
propane triacrylate into microporous hydrophobic beads, any heat recovery.
followed by covalent immobilization of TEPA via epoxy 1
groups. Primary amine groups on the resultant material were Qr = Cp,sΔT + ΔHa
converted into secondary amines by reaction with propylene qw (9)
oxide. The material showed an uptake of 2.03 mmol/g in the
where qw (kg CO2 per kg adsorbent) is the CO2 working
presence of pure CO2 at 60 °C, while under humid conditions
capacity, Cp,s (kJ/kg.K) is the adsorbent specific heat capacity,
(6% O2/8% H2O/30 ppm of SO2/CO2; 41% RH), the CO2 ΔT (K) is the difference between the adsorption and
uptake increased to 2.15 mmol/g (1.94 mmol/g after 1000 desorption temperature, and ΔHa (kJ/kg) is the absolute
cycles) despite a lower CO2 concentration in the feed. value of the heat of adsorption. The first term on the right-
Moreover, the CO2 uptake under dry condition was unaffected hand side of eq 9 is the sensible heat required to heat the
by steam treatment at 120 °C for 360 h. In contrast, upon adsorbent over ΔT, and the second term is the latent heat to
exposing triamime-grafted hydrophilic silica (CARiACT, P10) desorb CO2. Another less important term, namely the sensible
to steam under similar conditions, Fayaz et al.508 reported 83%, heat for CO2, is often ignored. Sjostrom and Krutka30 reported
61%, and 26% decrease in CO2 uptake at 25, 50, and 75 °C, values mostly between 2 and 3 GJ/tCO2 for different
respectively. The superior performance of the hydrophobic supported amine adsorbents. Similarly, applying eq 9 to PEI/
copolymer framework was attributed to its ability to nanosilica adsorbents, Li et al.520 found regeneration heats in
significantly limit water uptake, while its macropores obviated the range of 1.7−1.8 GJ/tCO2. Veneman et al.521 calculated
pore blocking. Although incorporating long alkyl chains is a the regeneration heat of TEPA-impregnated HP2MG as the
viable strategy to introduce hydrophobic character to amine- sum of the adsorption heat, considered to be constant at 1.5
containing adsorbents, the resultant material typically adsorb GJ/tCO2, and the sensible heat as a function of CO2 working
less CO2 compared to their unmodified counterpart. capacity. At a working capacity in the range of 2.6− 3.1 mmol/
4.4. Effect of Water on the Regeneration Heat g, the regeneration heat was found to be ca. 1.8 GJ/tCO2. In
The viability of CO2 capture processes depends on the overall direct air capture applications, the regeneration heat is
energy cost per tonne of CO2 removed. Regarding CO2 necessarily higher because of the inherently lower CO2
capture from typical humid industrial gases over solid working capacity. Using dry air in the presence of an amine-
adsorbents, the key issue is the regeneration heat. To be containing adsorbent at 25 °C, Elfving et al.522 found
regeneration heats of 4.2 and 6.4 GJ/tCO2 for CO2 adsorption
competitive, CO2 adsorption processes should achieve
at 60 and 100 °C, respectively, with working capacities of 0.47
significant energy reduction compared to ca. 4 GJ/tCO2
and 0.55 mmol/g.
required for regeneration of the benchmark 30% MEA
The presence of humidity in the gas has two opposing effects
solution.513
on the regeneration heat, including a positive contribution
The regeneration energy includes many terms, depending on
associated with the often observed increase in CO2 capacity,
the adsorption process, whether a pressure (PSA), vacuum
indicating that a smaller amount of adsorbent is required to
(VSA), or temperature (TSA) swing adsorption, and the gas−
capture a given amount of CO2, under otherwise the same
solid contactor, such as fixed bed, structured bed, or fluidized
conditions. The negative contribution is associated with the
bed.514−516 In addition to the energy required to regenerate
evaporation of adsorbed water445,517−519 and the higher
the solid adsorbent as such, other terms associated, for
binding strength of CO2.131,132 Siegelman et al.411 found a
example, with heating the equipment, pumping, and steam regeneration heat of 2.8 GJ/tCO2 for adsorption of 4% CO2/
generation may be included. Whenever possible, this N2 at 40 °C over 2-ampd-Mg2(dobpdc) and desorption in
discussion will be limited to the regeneration heat to bring 100% CO2 at 140 °C. In the presence of water, at a rate of one
the adsorbent from the end to the beginning of the adsorption H2O and one CO2 per Mg2+ cation, the regeneration heat,
step, typically in a TSA operation, and the adsorption will be without heat recovery, increased by ca. 54% to 4.3 Gt/tCO2.
limited to CO2 and H2O. The regeneration heat for adsorbents Zhang et al.517 proposed a more elaborate energy equation
includes (i) the latent heat of desorption of CO2 and water, to calculate Qr while including the latent heat for water
(ii) the sensible heat to heat the material from the adsorption vaporization and taking into account the recovered heat (eq
to the desorption temperature, and (iii) the recovered heat, if 10)
any.445,517,518 The most straightforward options to decrease the
regeneration heat include lower heat of adsorption, higher CO2 1−α Q f
working capacity, lower water working capacity, and smaller Qr = Cp,s(ΔT ) + (1 − β)ΔHa + (1 − γ ) v H2O
qw qw
temperature swing between adsorption and desorption. Hence,
for amine-containing adsorbents, although water plays a pivotal (10)
role as CO2 adsorption promoter and material stabilizer, its where α, β, and γ are the fractions of heat recovery from
removal during adsorbent regeneration may be more costly sensible heat, latent heat of adsorption, and water vaporization
AJ https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

heat, respectively, Qv (kJ/kg) is the vaporization heat of water, other, shows that the heat of regeneration is much more
and f H2O is the moisture uptake, which is evaporated during sensitive to f H2O than qw or ΔHa, indicating the adverse effect
regeneration. The authors determined many of the parameters of water desorption on the energy requirement of the process.
in eq 10 for cyclic adsorption−desorption of 15% CO2/N2 at To investigate the effect of adsorbed water on the
70 °C over 40 wt % PEI/silica adsorbent, using a combination regeneration heat for supported PEI, Min et al.519 used a
of fluidized bed adsorption data, TGA adsorption measure- series of adsorbents consisting of silica impregnated with PEI
ments for CO 2 and water, and differential scanning and PEI functionalized with ethylene oxide or 1,2-epoxides
calorimetry. The recovery ratio α of sensible heat was assumed with C1−C4 alkyl chains. The CO2 and water working
to be 0.75, based on Veneman et al.,521 whereas β and γ were capacities shown in Figure 40a were determined by TGA-MS
estimated to be 0.25 and 0.5, based on other considerations. based on adsorption of 15% CO2/3%O2/10% H2O/2% Ar/N2
For ΔT = 60 °C, ΔHa = 95 kJ/mol, Cp,s = 1.81 kJ/kg·K, qw = (51% RH) at 60 °C and desorption in the presence of 100%
1.35 mol/kg, and f H2O = 2 wt %, the heat of regeneration was CO2 at 110 °C. Although the CO2 working capacity decreased
found to be 2.46 GJ/tCO2. Comparison of the different as the PEI was modified by heavier epoxides, the water
contributions to Qr for this adsorption process and the working capacity decreased more strongly, particularly up to
benchmark 30% MEA system (Figure 39) shows that the main 1,2-epoxybutane (C2-E-PEI/SiO2). The heat of CO2 adsorp-
tion decreased from 1.83 GJ/tCO2 for PEI/SiO2 to about 1.48
GJ/tCO2 for C4-E-PEI/SiO2 with 1,2-epoxyhexane-modified
polymer. The heat of adsorption of water and the specific heat
capacity of the adsorbent also decreased from 2.62 GJ/tH2O
and 1.41 J/g·K for PEI/SiO2 to 2.22 GJ/tH2O and ca. 1.31 J/g·
K for C4-E-PEI/SiO2. The regeneration heat was calculated
using an equation similar to eq 10 but without any heat
recovery. Figure 40b shows that while the combined
contribution of the sensible and latent heat for CO2 desorption
was similar for all adsorbents, the regeneration heat depended
strongly on the latent heat for water desorption. These findings
are in general agreement with Quang et al.,445 who used a
similar relationship as eq 10, also without heat recovery. They
assumed that the latent heat of CO2 desorption does not
depend on the adsorbent moisture content, but the sensitive
Figure 39. Comparison of regeneration heat requirement for CO2 heat increases somewhat with moisture content as a result of
capture over PEI/silica with the benchmark 30% MEA system. increasing specific heat capacity. However, the most moisture-
Reproduced with permission from ref 517. Copyright 2016 Elsevier sensitive contribution was the vaporization heat of water. In
Ltd. summary, these three investigations showed that while the
CO 2 working capacity is important, managing water
advantage of adsorption resides in the much smaller sensible adsorption−desorption is critical.445,517,519
heat corresponding to the difference between heating a solid It is clear that all other things being equal, the regeneration
material and an aqueous solution. Notice that in this particular heat decreases as the CO2 working capacity increases and/or
example, qw and f H2O were somewhat low, and ΔHa was the water working capacity decreases. Because of their
particularly high. Sensitivity analysis of the influence of each hydrophilic nature, most amine-containing materials adsorb
parameter, assuming that they are independent from each significant amounts of water,75,135,162,227 which may increase

Figure 40. (a) CO2 and H2O working capacities of the amine/silica composite adsorbents. The black solid line indicates the ratio of the CO2 and
H2O working capacity. (b) Heat required for adsorbent regeneration. Reproduced with permission from ref 519. Copyright 2018 American
Chemical Society.

AK https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

the energy penalty to unacceptable levels. The penalty is than CO2 adsorption. (iii) Limiting the desorption of water
particularly severe when CO2 capture is conducted at low during regeneration: this technique would be applicable if the
concentration and temperature and high RH as in DAC and in purge medium has high humidity, such as steam. (iv)
air revitalization in confined spaces, with pCO2 of 400 to ca. Preconditioning the flue gas by decreasing its water content
5000 ppm.485,523 Bos et al.488 reported detailed energy via partial condensation. It was found that lowering the dew
calculations for low concentration CO2 capture over Lewatit. point of the flue gas from 47 to 30 °C resulted in reduction of
Figure 41 shows the regeneration heat for CO2 adsorption at water uptake from 0.85 to 0.25 mol H2O per mole of captured
CO2 from flue gas. This lowered the regeneration heat from 3
to 2.6 GJ/tCO2. Although this strategy seems to be the best
solution, adding a drying unit prior to CO2 capture can
significantly increase the operation costs. Instead of adjusting
the process parameters, Yogo and co-workers526−528 intro-
duced terminal alkyl groups to impregnated TEPA to facilitate
CO2 desorption and reduce the energy penalty associated with
regeneration. Steam-aided vacuum swing desorption (SA-VSA)
over isopropyl-substituted TEPA solid sorbents following
exposure to ca. 10.5% CO2/N2 with 50% RH required a
regeneration heat of 1.65 GJ/tCO2 to achieve more than 90%
CO2 recovery, compared to 80% for unsubstituted TEPA-
based adsorbents.526 Choi et al.529 reduced the regeneration
heat of PEI-impregnated silica by removing primary amine
groups in favor of secondary amines, through treatment with
1,2-epoxybutane. Not only epoxidation decreased the heat of
adsorption of both CO2 and H2O, but more importantly, it
inhibited water adsorption to a large extent without affecting
CO2 uptake significantly. As a result, the regeneration heat
Figure 41. Regeneration heat for DAC over Lewatit. For conditions, calculated for a TSA operation with adsorption of 15% CO2,
see text. The H2O to CO2 working capacity ratio is indicated on the 10% H2O, 2% Ar in N2 balance at 60 °C (50% RH), and
inner abscissa. Reproduced with permission from ref 488. Copyright desorption at 110 °C in pure CO2, decreased from ca. 4 to 2.67
2019 Elsevier. GJ/tCO2 upon PEI epoxidation. By far, the main difference
was the latent heat of water desorption.
two different concentration versus RH. The conditions were as
follows: adsorption at 20 °C for 15 min, desorption at 116 °C 5. MEMBRANES
and 0.5 bar, and RH up to 86%. The CO2 working capacity was Membrane CO2 capture is beyond the scope of this
calculated based on the amount of CO2 collected at 900 s of contribution. Nonetheless, because membranes often contain
desorption. It is seen that at 400 ppm, under dry conditions, actual CO2 adsorbents, such as amines,171,174,530 ionic
the regeneration heat is as high as 28 GJ/tCO2, mostly because liquids, 182,183 enzymes 184 or mimic enzymes, 185,186
of the low qw, thus the high sensible heat required to heat the MOFs,176−178 carbon,181 and zeolites,177,179,180 the water
large amount of adsorbent. Here, the presence of water in the benefits or drawbacks on CO2 membrane separation will be
feed actually decreased the regeneration heat because the briefly discussed, with particular emphasis on water-facilitated
additional energy required to evaporate adsorbed water was transport in polymeric membranes.171,188,189,530−532 Compre-
overcompensated by the higher CO2 working capacity. At 5000 hensive reviews on membrane CO2 removal from large point
ppm, the overall regeneration heat was dictated by the sources are readily available else-
opposing effect of the water and CO2 adsorption capacities. where.174,176−181,187,190,530,533,534
Similarly, Wurzbacher et al.485 reported a regeneration heat of In a typical membrane gas separation process, the feed gas
11−14.5 GJ/tCO2 for CO2 capture over APDES/NFC at 10− mixture is driven by a pressure (concentration) difference
30 °C in the presence of up to 80% RH. Kulkani and Sholl524 across the membrane and separated into one or more gaseous
estimated the regeneration heat to be 6.7 GJ/tCO2 for DAC components by selective permeation, thus generating a specific
over triamine-grafted PE-MCM-41 under dry conditions. Sinha gas-enriched permeate or retentate. Permeability and selectiv-
et al.525 reported a surprisingly low value of 3.3 GJ/tCO2 for ity are the most important parameters for evaluating the
DAC with 25% RH over mmen-Mg2(dopdc) in a TVSA performance of membranes.535 Note that there is often a
process using a structured contactor. permeability/selectivity trade-off with an upper bound.536 An
Veneman et al.227 proposed several strategies to curb the ideal membrane for CO2 separation should have an optimum
energy requirement for water desorption, either through the combination of high CO2 permeability and selectivity, in
adsorption process or the material design. These are as follows: addition to other attributes, depending on the feed
(i) Conducting adsorption at higher temperature with the composition, temperature, and pressure.537 CO2 separation
expectation that the relative decrease in the uptake of membranes are comprised of three major groups, namely
physisorbed water would overcompensate the smaller decrease inorganic, polymeric (or organic), and mixed matrix
in chemisorbed CO2, provided that the CO2 working capacity membranes.
remains within acceptable range, i.e., higher than ca. 2 mmol/g. Inorganic membranes are based on materials such as
(ii) Using the difference in the adsorption kinetics of water and MOFs,176−178 zeolites,177,179,180 carbon,181,538 and gra-
CO2: the effectiveness of this strategy is contingent upon using phene.539,540 They are typically porous, with high separation
an adsorbent whose water adsorption is significantly slower factor due to well-defined pore sizes and molecular sieving
AL https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

properties. Similar to their adsorbent counterparts discussed in


section 3, the CO2 separation performance of inorganic
membranes is often adversely affected in the presence of
water, particularly at low temperatures.541−544 Representative
behavior was reported by Gu et al.,542 using Y zeolite-based
membranes for separation of CO2/N2 mixtures under dry and
moist conditions in the temperature range of 23−200 °C. In
the presence of equimolar CO2/N2 at 23 °C and atmospheric
pressure, the CO2 selectivity in the presence of 94% RH (PH2O
= 2.64 kPa) was less than 1 versus 31.2 under dry conditions,
mostly because of decreasing CO2 permeance due to the
membrane pores being blocked by water. The N2 permeance
was low and hardly changed with humidity because the zeolite
pores were filled with either CO2 or H2O under dry and humid
condition, respectively. At increasing temperature up to 200 °C
(0.17% RH) using the same humid gas composition, the
permeance of all three components increased at different rates.
Below 80 °C, because of the decreasing affinity of water to Y Figure 42. CO2 permeability versus water content in hydrogel
zeolite, the CO2 permeance increased slowly, whereas beyond membranes at 23 °C and 5.15 bar (AGA, alginic acid; CE, cellulose;
80 °C, water desorbed significantly, and the CO2 permeance CEP, cellophane; CMC, carbonyl methyl cellulose; CS, chitosan;
increased rapidly. As a result, the CO2/N2 selectivity for the PVA, poly(vinyl alcohol); PVAm-X, polyvinylamine with X% PVA).
moist system increased and then decreased slightly with Reproduced with permission from ref 546. Copyright 2008 Elsevier.
temperature after reaching a maximum at ca. 140 °C. Above
110 °C, the CO2 versus N2 selectivity of the moist mixture was ment of the membrane polymer chains, similar to unsupported
significantly higher than that of the dry mixture, but their polyamine adsorbents.460−463
permeances were lower under humid condition over the entire Dai et al.547 investigated the CO2 separation properties of
range of temperature. Additional measurements carried out at sulfonated multiblock polymer membranes, referred to as
200 °C in the presence of equimolar CO2/N2 with up to 48 TESET, before and after submersion in water for 24 h and
kPa H 2 O (up to 3.1% RH), showed decreasing CO 2 drying under vacuum at room temperature as a function of RH.
permeance with increasing water partial pressure, whereas The membranes were casted from different solvents, including
the N2 permeance decreased up to 12.3 kPa of water (0.8% 50/50 v/v cyclohexane/heptane (CH), 85/15 w/w toluene/
RH) then remained practically constant. As a result, the CO2 isopropyl alcohol (TIPA), chloroform (CF), and tetrahydro-
selectivity reached a maximum value of 4.6 at 0.8% RH. furan (THF). Under dry condition, all membranes exhibited
Similarly, chabazite membrane grown on tubular alumina particularly low CO2 permeability (<30 Barrer) and CO2/N2
support showed limited decreases in CO2/N2 and CO2/CH4 selectivity (<30). With the exception of the THF-derived
selectivity of 11 and 15%, and CO2 permeance of 18% and membrane, before exposure to liquid water, all samples showed
similar separation properties with CO2 permeability in the
20% at 200 °C, in the presence of equimolar humid CO2/N2
range of 20−80 Barrer and CO2/N2 selectivity of 8 to ca. 60
and CO2/CH4 mixtures (8% H2O, 0.5% RH), respectively,
(Figure 43). The THF-cast membrane showed a CO 2
compared to dry condition.543 Bernal et al.544 also reported a
permeation of 300 Barrer at the highest RH of 95%. After
decrease in CO2 permeability through stainless steel-supported
submersion in water, all membranes showed significantly
boron-substituted ZSM5 membrane. However, the CO2 higher CO2 permeabilities, reaching 482 Barrer for the CF-
selectivity was not significantly affected by water, even at low derived membrane at 95% RH. All such improvements of the
temperatures, presumably because of the limited hydrophilicity membrane separation properties in the presence of water
of the B-ZSM-5 membrane.
With regard to polymeric membranes, the most straightfor-
ward effect of water on CO2 separation is associated with
polymer swelling. Water-swellable polymer membranes,
including nonreactive hydrogels and membranes containing
CO2 carriers (vide infra), such as polyamines, polyamides,
mixed matrix polymers, and polymer electrolytes, exhibit much
higher mass transfer in the presence of water due to diffusion
enhancement and other water-facilitated CO2 capture mech-
anisms. This is in contrast to the often prevailing solution-
diffusion mechanism545 under dry conditions, which is
governed by the permeability/selectivity trade-off mentioned
above.536 Swelling induced by water drastically increases the
polymer free volume, which is directly responsible for diffusion
enhancement, leading to higher CO2 permeance. Figure 42 Figure 43. CO2 separation performance of TESET membranes cast
shows the dramatic effect of water content in hydrogel from different solvents, as indicated, before and after submersion in
membranes, including polyvinylamine-poly(vinyl alcohol) liquid water (filled and open symbols, respectively). The dashed line
(PVAm-PAA) blends, on their CO2 permeability.546 In represents the Robeson upper bound. Reproduced with permission
addition, water swelling is often accompanied by disentangle- from ref 547. Copyright 2019 Springer Nature.

AM https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

vapor, and upon immersion in liquid water were associated CO 2 permeability increased significantly, the CO 2 /N 2
with morphological changes due to polymer swelling and the decreased.
formation of nanoscale channels promoting CO2 diffusion. Reactive polymeric membranes may contain fixed-site and/
While optimum humidity promotes both CO2 permeability or mobile carriers that react reversibly with CO2. Examples of
and selectivity, often exceeding the upper bound Robeson CO 2 carriers include amines, 171,174,188,530,549,550 poly-
limit, excessive humidity may have an adverse effect on amines,171,174,530,540,550,551 ionic liquids (ILs),182,183,552 en-
selectivity through relaxation of the polymer network.548 Using zyme,184 and mimic enzyme.185,186 As shown schematically in
a nanofibrillated cellulose (NFC)-based membrane doped with Figure 45, in polymeric membranes with fixed-site carriers,
1-ethyl-3-methylimidazolium acetate ionic liquid (IL), as
shown in Figure 44, Janakiram et al.182 identified three

Figure 45. Schematic diagram of transport mechanisms across


membranes.

CO2 diffuses across the membrane by hoping from one site to


another, from the high-pressure side to the permeate interface,
whereas for membranes with mobile carriers, CO2 reacts with
the carrier forming an intermediate product that migrates
through the membrane toward the low-pressure side, under the
concentration gradient driving force. In both cases, CO2-carrier
product dissociates at the permeate interface, leading to the
Figure 44. Mixed-gas CO2 permeability and CO2/N2 separation separation of CO2 and regeneration of the carrier. Such a
factor for two hybrid NFC/IL membranes versus RH at 35 °C and 1.7 mechanism is referred to as a facilitated transport mecha-
bar. R-I, R-II, and R-III indicate the three regimes of gas permeation nism.174,530−532,537 Occurrence of this mechanism is often
with the corresponding schematic morphologies. Reproduced with accompanied by enhancement of both permeability and
permission from ref 182. Copyright 2020 Royal Society of Chemistry.
selectivity for the target species, thus providing a method to
overcome the above-mentioned permeability/selectivity trade-
off limitation.537 Moreover, in many instances, the CO2-carrier
permeation regimes, depending on relative humidity. The first reaction is either more efficient in the presence of water, or
regime (R-I) is dominated by low permeation of semicrystal- requires water. Figure 46 illustrates the facilitated transport of
line NFC nanofibrils because the humidity level is not CO2 in the presence of carbonic anhydrase-inspired imidazole-
sufficient to swell the membrane and create continuous Zn(II).171 In the presence of this carrier, CO2 diffuses through
water channels. In this regime, CO2 permeation occurs the membrane as bicarbonate, whose formation involves water.
through chemisorption on IL and diffusion along the NFC/ In the presence of amine carrier sites, the reaction-
IL interface, explaining its strong sensitivity to IL content. The decomposition cycle takes place with the intermediacy of
next regime (R-II) occurs at RHs sufficient for NFC swelling ammonium carbamate (eq 5) or bicarbonate (eq 11),
with decreased viscosity. Major morphological changes occur, depending on the humidity of the feed gas and/or the
including physical separation of nanofibrils due to disruption of membrane. Interestingly, using tertiary amines or sterically
interfibrillar hydrogen bonding by water and formation of hindered amines to inhibit the formation of ammonium
contiguous IL/water channels along the fiber surfaces. Hence, carbamate in favor of ammonium bicarbonate with higher
CO2 permeation is promoted due to higher solubility and stoichiometric CO2/amine ratio proofed to be an excellent
faster diffusion compared to N2, leading to much higher CO2/ strategy to achieve high CO2 permeability and selectivity in the
N2 selectivity for the 35% IL-containing membrane. As for the presence of water.175,549 Using poly-N-isopropylpolyallylamine
membrane containing 50 wt % IL, the selectivity hardly as a fixed-site carrier with hindered amine groups, at 110 °C
changed, presumably because of excessive swelling. In the third and 58% RH, Zhao and Ho549 achieved an impressive CO2
regime, further swelling takes place, with the formation of permeability of 6600 Barrer and CO2/H2 selectivity of 340,
continuous pathways bulk water channels containing IL, compared to 3950 Barrer and 160, respectively, for unmodified
allowing fast diffusion of both CO2 and N2. Hence, while polyallylamine.
AN https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

consisting typically of a polymeric membrane with a dispersed


phase, which could be an inorganic filler such as MOFs, ILs,
zeolites, carbons, and metal oxides.554,555 or organic filler such
as polyaniline particles.556 Depending on the MMM design,
the polymer and the filler may play different roles in the CO2
separation process. In membranes consisting of an inert
polymer and an active dispersed phase, the filler adsorbs
significant amount of water, creating efficient CO2 transport
pathways. Liu et al.557 used a hydrophobic polyimide
membrane combined with carbon nanotubes functionalized
by polyzwitterions containing quaternary ammonium sulfonate
(Figure 48). Under humid conditions, the MMM showed

Figure 46. CO2 hydration and deprotonation catalyzed by carbonic


anhydrase-inspired imidazole-Zn(II). Reproduced with permission
from ref 171. Copyright 2017 Royal Society of Chemistry.

R3N + CO2 + H 2O → R3NH+HCO3− (11)

Kim et al.489 demonstrated that hollow fiber-based membrane


containing hyperbranched aminosilica (HAS) obtained by in
situ polymerization of aziridine exhibit distinctly different Figure 48. CO2 transport pathways within the hydration layer at the
behavior in humid versus dry conditions, switching from CO2- polymer−filler interface. Reproduced with permission from ref 557.
to N2-selective membrane, respectively. These findings were Copyright 2014 American Chemical Society.
associated with different interaction mechanisms between CO2
and amine, depending on the actual conditions (Figure 47). In
dry condition, CO2−amine interactions triggered polymer significantly enhanced CO 2 permeability compared to
chain cross-linking, which significantly decreased the CO2 unmodified polyimide membrane. It was argued that because
permeability. However, in humid conditions, amine cross- of the high hydrophilicity of the zwitterion, the filler adsorbs
linking was limited, while the CO2 solubility was enhanced, significant amount of water. Moreover, the particles were
leading to improved CO2 permeation and near-quantitative found to form an interconnected network morphology with
amine−water−CO2 reaction to form bicarbonate. Additional preferred CO2 transport pathways within the continuous
data regarding enhancement of CO2 permeability and hydration layer. Further enhancement of CO2 permeability
selectivity in the presence of a variety of carriers and water under humid conditions was achieved by incorporating highly
may be found elsewhere.171,537 hydrophilic nanohydrogel particles within the polyimide
While inorganic membranes are the most robust because of membrane.558 Water-facilitated CO2 capture also takes place
their superior mechanical, thermal and chemical stability, and in active membranes such as PVAm combined with fillers such
offer higher selectivity and permeability, they are very brittle as polyaniline nanosheets556 or hydrotalcite particles,559 which
and significantly more expensive. In contrast, polymeric disrupt the polymer chain packing, creating transport channels.
membranes are relatively cheaper and easier to process, Another important effect of water on CO2 separation by
however, they are less stable, susceptible to physical aging membranes is the salting out affect introduced by Zhang and
and CO2-induced plasticization, which may lead to loss of Wang.560 Addition of aqueous solutions of salts to polymeric
selectivity and mechanical strength.553 An excellent trade-off membranes was found to decrease the solubility of nonreactive
between membrane permeability, selectivity, and processability gases, much more than the solubility of CO2, leading to
can be achieved by mixed matrix membranes (MMM), significantly higher selectivity for CO2 separation, albeit at

Figure 47. CO2 adsorption on grafted-polyethylenimine HAS membrane under dry condition. Reproduced with permission from ref 489.
Copyright 2015 American Chemical Society.

AO https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Figure 49. (a) Permeability vs selectivity plot of Pebax electrolyte membranes with and without added salt (Ca1.5PW, calcium phosphotungstate;
CaLS, calcium lignosulfonate). (b) Corresponding CO2 permeability vs free water content and CO2/N2 selectivity vs bound water content.
Reproduced with permission from ref 171. Copyright 2017 Royal Society of Chemistry.

slightly lower permeability. For example, immersion of water-


swollen PVA membrane into 0.5 M NaF increased the CO2/
N2 separation factor from 79 to 500, while it decreased the
CO2 permeance from 0.79 GPU to only 0.53 GPU.560
Similarly, upon dipping a polyallylamine membrane in 0.5 M
of Na3[PW12O40], the CO2/N2 separation factor increased
from 50 to 1700, whereas the permeance decreased
considerably from 8.4 to 1.7 GPU. It was surmised that the
salting-out effect could be attributed to the preferential
attraction between the salt ions and the water-swollen polymer,
which would decrease the amount of “free water” and PVA
macromolecules to dissolve nonreactive gases, such as N2, as
opposed to the favorable water−CO2 interactions in the water-
swollen membrane.560 Li et al.561 found that not only
humidified poly(ether-b-amide) (Pebax) doped with calcium
chloride, calcium phosphotungstate, and calcium lignosulfo-
nate exhibits significantly higher CO2 permeability and
selectivity than their dry counterparts (Figure 49a), with a
ratio exceeding the traditional upper bound limit,536 but the
CO2 permeability and selectivity increase linearly with the total
water and bound water content, respectively (Figure 49b).
Consistent with Zhang and Wang’s proposal,560 they
hypothesized that bound water surrounding metal ions does
not have the ability to dissolve noncondensable gases and Figure 50. Schematic illustration of the effect of bound water on
disrupts the continuity of their transport routes. However, CO2/N2 separation for Pebax-containing calcium salts. Reproduced
with permission from ref 562. Copyright 2015 Royal Society of
water bound to the polar groups of Pebax may facilitate the Chemistry.
hydration of CO2 into bicarbonate anion and its permeation
through the membrane. Hence, as shown schematically in
Figure 50, the adverse effect of salt additives on the
permeability of noncondensable gases would be much more including increased CO2 solubility,565 enhanced CO2-complex
severe, in comparison to CO2 permeability, with calcium transport across the membrane,489,552 and occurrence of
phosphotungstate showing the highest amount of trapped preferential transport pathways.557,558,562 As the CO2 per-
water and the strongest salting-out effect.562
meation increases with higher membrane wetting, the
To sum up this section, while moisture has a limited or even
adverse effect on CO2 separation by inorganic membranes, it selectivity may go through a maximum value.546 In contrast,
plays a key role in facilitating CO2 transport across flexible in the presence of rigid polymers such as polymers of intrinsic
polymer membranes, through a multitude of mechanisms. microporosity567 and thermally rearranged polymeric mem-
Water is necessary for the reversible CO2 hydration.562,563 branes568 or in the presence of hollow fiber membrane
Moreover, water can swell the majority of polymer membranes, contactors,569 water may adversely affect CO2 permeability,
leading to enhanced CO2 permeability through decreased mostly because of competitive adsorption. Moreover, for flue
viscosity564,565 and high fractional free volume.546,566 Depend- gas containing NOx and SOx, the presence of water may lead to
ing on the nature and structure of the membrane, other water- permanent degradation of the membrane due to acid-catalyzed
facilitated CO2 transport mechanisms may also apply, hydrolysis of specific functional groups, e.g., nitrile.567
AP https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

6. REACTIVE ADSORBENTS moisture content of 40−90 wt % at 30 °C had CO2 uptakes


Because of the basic nature of some metal oxides, hydroxides below 0.74 mmol/g. Decreasing water content to 35 wt % led
and carbonates, they react with acidic CO2, often at high to a dramatic increase in CO2 uptake to 3.55 mmol/g, in
temperature, while demonstrating high CO2 uptake and addition to enhanced adsorption rate. At optimum water
selectivity.62,570,571 Contrary to typical chemical adsorbents, content of 30 wt %, the material exhibited a CO2 uptake as
such as amine-functionalized materials, they interact with CO2 high as 6.41 mmol/g and the fastest kinetics. Further decrease
through a bulk as opposed to surface reaction. Alkali metal of the water content to 25 wt % slightly increased CO2 uptake
carbonates 74,169,572−574 and alkaline earth metal ox- to 6.50 mmol/g, but the adsorption rate was about four times
ides74,575−578 are two important groups of metal-based lower. Figure 51 shows the efficiency of CO2 removal over
adsorbents, widely used for CO2 capture and sequestration.
6.1. Alkali Metal-Based Adsorbents
6.1.1. Sodium Carbonate. Alkali metal carbonates react
with CO2, but only in the presence of water to form
bicarbonates (eq 12).74,573,579,580 Sodium carbonate is a
common adsorbent that reacts with CO2 under humid
condition generating NaHCO3.580,581 Above 70 °C, the
material reacts with CO2 and water to produce preferentially
Wegscheider’s salt, according to eq 13, instead of bicarbon-
ate.582
M 2CO3(s) + H 2O(l) + CO2(g) ⇋ 2MHCO3(s)
(M = Na, K) (12)
Figure 51. CO2 removal efficiency at 50 °C for Na2CO3 pre-exposed
to water vapor (100% RH) for different durations. Reproduced with
Na 2CO3(s) + 0.6CO2(g) + 0.6H 2O(g) permission from ref 580. Copyright 2007 Elsevier Ltd.
⇋ 0.4[Na 2CO3 · 3NaHCO3](s) (13) Na2CO3 in the presence of 10% CO2/12.2% H2O/N2 (100%
Dong et al.583
found that at 60 °C in the presence of 15% RH) at 50 °C as a function of pretreatment time in water-
CO2/15% H2O/N2 (76% RH), unsupported Na2CO3 gave rise saturated nitrogen.580 It is clear that CO2 removal increased as
mostly to Wegscheider’s salt, whereas alumina-supported the duration of water pretreatment increased and leveled off at
Na2CO3 formed bicarbonate. Under specific conditions, ca. 15 min. Overexposure of the material to water vapor,
sodium sesquicarbonate, Na2CO3·NaHCO3·2H2O, can also however, could result in agglomeration, hindering proper
be produced in the presence of CO2 and moisture.582 While fluidization of the adsorbent.
humid CO2 adsorption at high concentration takes place on 6.1.2. Potassium Carbonate. Potassium carbonate is
the basis of eqs 12 and 13, PXRD data584 showed that at 30 °C recognized as a more effective adsorbent than sodium
in the presence of 1−5% CO2/N2 with 74% RH, only Na2CO3· carbonate.573,586 Hayashi et al.586 impregnated K2CO3 on
H2O (eq 14) was detected due to unfavorable conditions for AC and compared its CO2 capture performance in the
the forward reactions of eqs 12 and 13 to take place at low presence of 10% water to Li2CO3- and Na2CO3-based
CO2 pressure. Increasing the CO2 concentration up to 10% adsorbents. As illustrated in Figure 52, below 80 °C, K2CO3
gave rise to Na5H3(CO3)4 (eq 15), eventually leading to the
formation of bicarbonate, according to eq 16.
Na 2CO3(s) + H 2O(g) ⇋ Na 2CO3 ·H 2O(s) (14)

5Na 2CO3 ·H 2O(s) + 3CO2(g)


⇋ 2Na5H3(CO3)4(s) + 2H 2O(g) (15)

Na5H3(CO3)4(s) + CO2(g) + H 2O(g) ⇋ 5NaHCO3(s)


(16) Figure 52. Effect of temperature on CO2 uptake of alkali metal
The beneficial effect of the feed humidity on CO2 adsorption carbonates in the presence of 13.8% CO2/10% H2O/He. Reproduced
kinetics and uptake over Na2CO3 was also reported by Luo with permission from ref 586. Copyright 1998 American Chemical
Society.
and Kanoh,584 as increasing RH from 38% to 74% reduced the
equilibrium time from 300 to 100 min, along with a slight
increase in CO2 uptake in the presence of 30% CO2/N2, i.e.,
from 7.21 to 7.39 mmol/g. In parallel, the intensity of the reacted quantitatively with CO2 on the basis of eq 12.
diffraction peaks assigned to Na5H3(CO3)4 increased, indicat- Although CO2 uptake of K2CO3 decreased at higher
ing the favorable impact of water on the basis of eq 15. temperatures, it was still 82% of the maximum uptake at 100
Beside water vapor being an essential reactant for CO2 °C. Similarly, CO2 uptake over Na2CO3 decreased with
adsorption over Na2CO3, controlling the humidity of the temperature; however, the resulting NaHCO3 decomposed at
adsorbent was found to be of critical importance.585 Upon lower temperature compared to KHCO3. Moreover, CO2
exposure to pure CO2 at ambient pressure, materials with uptake over Li2CO3 was very poor, irrespective of temperature.
AQ https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Adsorption of CO2 over K2CO3 between 40 and 80 °C is of KHCO3 and K2CO3 only, without considering the possible
believed to take place according to eq 12, in two steps occurrence of hydrates.
involving hydration into an active adsorption intermediate, 6.2. Alkaline Earth Metal-Based Adsorbents
K2CO3·(1.5H2O), as shown in eq 17, followed by reaction with
Chemical looping involving the carbonation of alkaline earth
CO2 (eq 17).573,586−588
metal-containing compounds, particularly Ca and Mg-rich
K 2CO3 · (1.5H 2O)(s) + CO2(g) ⇋ 2KHCO3(s) + 0.5H 2O(g) materials such as their oxides, hydroxides and mixed oxides,
(17) and wollastonite (CaSiO3), was extensively investigated as a
means of CO 2 separation577 or permanent sequestra-
Nevertheless, PXRD data showed in addition to K2CO3· tion.18,592,593 Contrary to alkali metal-based materials,
(1.5H2O), the occurrence of K4H2(CO3)3·(1.5H2O) according adsorbents containing alkaline earth metal oxides or hydrox-
to eq 18, which ultimately gave rise to KHCO3 (eq 19).589 ides can react with CO2 in the absence of water, according to
2K 2CO3(s) + 2.5H 2O(g) + CO2(g) eqs 20 and 21.62,594

⇋ K4H 2(CO3)3 ·(1.5H 2O)(s) (18)


MO(s) + CO2(g) → MCO3(s) (M = Mg, Ca) (20)

K4H 2(CO3)3 ·(1.5H 2O)(s) + CO2(g) M(OH)2(s) + CO2(g) → MCO3(s) + H 2O(g)

⇋ 4KHCO3(s) + 0.5H 2O(g) (M = Mg, Ca) (21)


(19)
Lee et al.74,169,573,588 carried out extensive work on CO2 6.2.1. Calcium Oxide and Hydroxide. Among the oxides
adsorption over supported K 2 CO 3, demonstrating the studied for CO2 adsorption, CaO has received much attention
importance of the humidity of the gas feed and prehumidifi- because it is readily available, cheap, and nontoxic, with
cation of the adsorbent. Increasing the pretreatment RH from favorable thermodynamic properties and theoretical capacity of
6% to 75% increased the CO2 uptake at 60 °C in the presence 0.78 g/g CaO.595−598 Calcium looping process (CLP) is a
of 1% CO2/N2 with 6% RH from 0.58 to 2.39 mmol/g while promising technology for CO2 capture and release, involving
equally enhancing the adsorption rate. The beneficial effect of carbonation, typically at 600−700 °C, and calcination at ca.
prehydration was attributed to higher conversion of K2CO3 to 900 °C (Figure 53).599 This process uses high temperature
the active K2CO3·(1.5H2O), as confirmed by PXRD measure-
ments. Moreover, at low humidity, the K2CO3·(1.5H2O)
intermediate phase could be easily converted into K2CO3
during thermal regeneration. Therefore, the RH should be
maintained sufficiently high to prevent the active phase from
transforming back into K2CO3.588
Unlike K2CO3/AC, K2CO3/Al2O3 deactivated over cyclic
operation with adsorption at 60 °C and regeneration at 150 °C
due to the formation of inactive mixed metal hydroxycar-
bonates, KAl(CO3)2(OH)2.573 In the presence of K2CO3/ Figure 53. Schematic of calcium looping process (CLP). Reproduced
MgO, CO2 adsorption took place between 200 and 320 °C and with permission from ref 599. Copyright 2017 Elsevier Ltd.
above 600 °C, whereas under humid conditions, CO2 adsorbed
at temperature as low as 60 °C, up to 140 °C.74 Using 1% streams with the potential of reducing the energy penalty
CO2/N2 at 60 °C in the presence of 7%, 9%, and 11% H2O associated with CO2 capture.599,600 The merits and drawbacks
(35%, 46%, and 56% RH), the CO2 uptake over 30% K2CO3- of water/steam on hydration of adsorbent, carbonation, and
impregnated MgO increased with humidity, reaching ca. 2.95 calcination reactions were summarized by Zhang et al.601
mmol/g at 56% RH. At 45% RH, the CO2 uptake over Nonetheless, there is ample evidence that water has beneficial
K2CO3/MgO was 30% higher than the maximum uptake based effects on both carbonation and regeneration and even
on CO2/K2CO3 = 1, indicating that the support contributed to reactivation of spent CaO.602−604
the overall CO2 adsorption. In addition to direct adsorption of Carbonation of Ca−Al mixed oxide (Ca/Al = 2.5) at 600 °C
CO2 over MgO as discussed in section 6.2.2, MgO promoted in the presence of 50% CO2/N2 showed that the initial CaO
CO2 uptake of K2CO3 through the formation of mixed metal carbonation increased with moisture, from 12.7 mmol/g under
carbonates, namely K2Mg(CO3)2 or K2Mg(CO3)2·4(H2O). dry conditions to a maximum of 16.1 mmol/g in the presence
Similar to K2CO3/Al2O3, the decrease in CO2 capacity of of 1.6% water vapor, while further increase in water led to the
K2CO3/MgO over time was attributed to the formation of opposite effect.604 Moreover, the stability of the material
inactive mixed metal carbonates.74,169 during multiple cycles of carbonation at 600 °C and calcination
In contrast to the reports discussed above,573,580,586−588 the in N2 at 700 °C in the presence of 1.6% water was higher than
Mahinpey group590,591 used a combination of TGA and MS to cycling in dry condition. As for the mechanism of the water
simultaneously monitor the consumption of both water vapor effect, water may promote the initial reaction-controlled, fast
and CO2 on several potassium-based materials and demon- stage of CaO carbonation or the subsequent slow stage
strated that K2CO3 hydration (eq 17) and carbonation (eq 12) controlled by diffusion of CO2 through the formed layer of
are actually competitive reactions that take place in parallel, carbonate. It was also hypothesized that water may catalyze the
with no direct conversion of K2CO3·(1.5H2O) into KHCO3 reaction, with the formation of a more active Ca(OH)2
(eq 17). Their main objection was that the two-step intermediate.605 Literature data provide strong evidence that
carbonation mechanism was based solely on TGA data water enhances diffusion in the product layer.604−606 For
assuming that after carbonation, the material was composed example, the morphology of spent Ca−Al mixed oxide after 17
AR https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

cycles of carbonation at 600 °C and calcination at 700 °C with 6.2.2. Magnesium Oxide. Magnesium oxide received
up to 1.6% water vapor consisted of loose and porous particles much attention as a sorbent or additive for CO2 adsorption at
compared to compact particles under dry conditions or in the different temperatures.169,615,616 Although water is not required
presence of more water, suggesting enhanced mass transfer for CO2 to react with MgO (eq 20), adsorption is strongly
during the diffusion-controlled stage of carbonation under promoted in the presence of humidity.74,167,594 For example, in
optimized moisture conditions. Interestingly, under dry the presence of 15% CO2/N2 at 50 °C, increasing RH from 0%
conditions, the CO2 uptake decreased from 12.5 to 8 mmol/ to 50% and 70% enhanced CO2 uptake over MgO from 0.66 to
g after 17 cycles but the adsorbent could be reactivated by 3.01 and 3.42 mmol/g, respectively, without the formation of
introduction of 1% steam from the 18th cycle, leading to a intermediate species as indicated by PXRD and FTIR data.167
steady increase in CO2 uptake up to 13.6 mmol/g in the 28th Nonetheless, capillary condensation of water should be
cycle.604 avoided, as prehydrating MgO at 50 °C with moist N2 (70%
Zhang et al.607 reported that compared to dry conditions RH) led to 34% decrease in CO2 uptake compared to the
(15% CO2/N2), carbonation at 650 °C followed by calcination nonhydrated material. It was inferred that the presence of
at 950 °C of 0.74 mol % Mn-doped CaO was ca. 100% higher capillary condensed water in the pores of MgO hindered the
in the presence of steam (15% CO2/15% H2O/N2) over diffusion of CO2 toward active sites. Moisture had no
multiple cycles. While steam brought about enhanced porosity, noticeable effect on the initial rapid uptake of CO2 because
Mn was thought to significantly reduce sintering of CaO it is a surface-limited reaction. However, similar to CaO,604−606
particles during calcination.604,606 Following CO2 adsorption at the rate of the diffusion-controlled stage was enhanced by
650 °C over CaO derived from limestone, calcination in humid moisture.167 PXRD and FT-IR data provided evidence that
nitrogen at 850 °C led to higher CO2 uptake in subsequent water vapor not only increases CO2 uptake, but also the
adsorption cycles compared to calcination under dry adsorption kinetics, presumably through the formation of
conditions.608 For instance, consistent with Guo et al.,604 MgO·H2O reactive intermediate.594 Zarghami et al.617 claimed
CaO conversion to CaCO3 after 20 cycles was ca. 16%, 23.4%, that the promoting effect of steam on the carbonation of MgO
and 25.5% when calcination was carried out in N2 with 0%, was due to the formation of MgO·H2O/Mg(OH)2 inter-
30%, and 80% RH, respectively, with RH calculated at room mediates, which are more reactive than MgO, and to the fact
temperature. This was attributed to the formation of larger that these species have larger molar volume than MgO, leading
pores in the presence of water vapor, enabling CO2 to diffuse to pore expansion, thereby alleviating diffusion limitations.
deeper into the bulk material and enhanced carbonation due to CO2 uptake at 60 °C over mesoporous MgO/Al2O3 (Mg/Al =
steam retained in the pores and interparticle voids after 1) was found to increase from 0.74 mmol/g under dry
calcination. condition (10% CO2/N2) to 1.74 and 2.0 mmol/g at 5% and
CaO-based materials were also investigated for CO2 capture 10% RH, respectively, which was attributed to the formation of
at low to moderate temperatures.166,609,610 Azmi et al.166 more reactive Mg(OH)2 phase.618 However, further increase in
compared commercial CaO and a CaO derived from waste RH led to a decrease in CO2 uptake, presumably due to
seashell cockle for CO2 adsorption at 2 °C and pressures up to diffusion resistance for CO2 owing to high water uptake. With
36 bar. Moisturizing the commercial-grade and synthesized regard to the carbonation of Mg(OH)2, it was found to be
materials to optimum water content of 32% and 67 wt % much faster than the carbonation of MgO and less sensitive to
increased CO2 uptake ca. 4 and 8 times, respectively. Beyond steam, as water is already obtained intrinsically by
these optimum water contents, lower CO2 uptake was dehydration.594
obtained, presumably due to the blockage of adsorbent pores Calcination of layered double hydroxides (LDH), also
by water. known as hydrotalcite, often leads to highly dispersed mixed-
Theoretical studies to elucidate the promoting effect of metal oxides with remnant lamellar structures. This synthesis
water on carbonation611−613 and calcination614 were also methodology was used to design Mg-containing oxides for
reported. DFT calculations showed that CO2 adsorption on CO2 adsorption. Ram Reddy et al.619,620 used Mg-Al oxides
CaO (001) is facilitated by the presence of water, suggesting nanoplatelets obtained by calcination of the corresponding
an Eley−Rideal type mechanism in which H2O preferentially LDH at 200 °C. They found that at 200 °C, switching from
adsorbs on the CaO (001) surface and activates adjacent O dry (14% CO2, 4% O2 in N2) to humid (12% H2O, 0.8% RH)
atoms, which are more reactive with CO2 compared to water- conditions increased CO2 uptake from 0.62 to 0.71 mmol/g.
free CaO surfaces.611,612 Consistent with TGA data, molecular Similarly, Hanif et al.621 carried out pure CO2 adsorption at
dynamic simulations using ReaxFF reactive force field showed 200 °C and 1.2 bar over Ni-Al-LDH nanosheets, prehydrated
that water vapor improved CaO carbonation mainly through at the same temperature. The CO2 uptake was found to be
enhanced diffusion, while the beneficial impact of water on the 1.22 mmol/g compared 0.66 mmol/g under dry conditions,
kinetics of carbonation was negligible.613 As often reported, He which was attributed to the formation of bicarbonate and
et al. 614 observed an increase in the rate of CaCO 3 carbonate in the presence of water.621
decomposition and a decrease in the onset decomposition In a series of studies, Coenen et al.622−626 investigated the
temperature during calcination of CaCO3 with increasing adsorption−desorption of CO2 under humid conditions over
steam content, suggesting a reduction in activation energy potassium promoted hydrotalcite-derived mixed oxides,
barrier, with water acting as a catalyst. Their DFT calculations consisting of ca. 20 wt % K and Mg/Al molar ratio of 0.54
revealed that water splits into OH− and H+, then the proton (KMG30). Adsorption data in the presence of humid CO2 by
reacts with CO32− of CaCO3 to form HCO3−, which then TGA-MS (400 °C, PH2O = 0.34 bar, PCO2 = 0.66 bar) and
decomposes to CO2 and OH−. The activation energy for column-breakthrough technique (400 °C, PH2O = 0.1 bar, PCO2
CaCO3 decomposition was determined to be 342 kJ/mol, = 0.025 bar) were interpreted based on the occurrence of
compared to 263 kJ/mol for HCO3−, i.e., in the presence of multiple adsorption sites (A, B, C, and D) with different
steam. affinities toward CO2 and water (Figure 54).623,624 In the
AS https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Figure 56. Conversion of bidentate carbonate into bicarbonate along


Figure 54. (a) Proposed model and (b) regeneration of adsorption with the corresponding wavenumbers. Reproduced with permission
sites for CO2 and H2O on hydrotalcite-based Mg-Al mixed oxide from ref 625. Copyright 2018 Creative Commons Attribution License.
adsorbent at 400 °C. Reproduced with permission from refs 623 and
624, respectively. Copyright 2017 and 2018 Creative Commons
Attribution License.

presence of pure water and pure CO2, weight gains of 0.44 and
0.27 mmol/g were obtained, corresponding to adsorption on
sites A and B, respectively. Upon exposure of CO2-loaded
material to dry N2, not all CO2 desorbed. However, the use of
moist N2 led to complete removal of the remaining CO2, which
was associated with adsorption site C (Figure 55). Further
exposure to N2 led to complete H2O desorption, but when
CO2 was reintroduced, the CO2 capacity exceeded 0.27 mmol/
g. This increase was attributed to the displacement of adsorbed
H2O by CO2.
Figure 57. Infrared bands associated with CO2 and/or water
Site D can adsorb CO2 only in the presence of water and can
adsorption on different sites of KMG30. Reproduced with permission
be regenerated completely by using water or partly by using from ref 625. Copyright 2018 Creative Commons Attribution License.
nitrogen.623 Moreover, the presence of humidity during
desorption increased CO2 working capacity from 0.3 to 0.53
mmol/g, suggesting that water facilitated the recovery of CO2 and water on the same sites may be different for different
strongly bound CO2 species, particularly from site C. FTIR materials. Increasing MgO content was found to enhance the
data provided direct evidence that the beneficial impact of CO2 working capacity on site B, a basic site that can be
water on CO2 working capacity was due to the transformation regenerated under flowing nitrogen. Furthermore, the working
of strongly bound bidentate carbonate species into bicarbonate capacity for water on site A was slightly higher for KMG
as shown in Figure 56.625 sorbents relative to alumina hydrotalcite. Figure 58 shows
The proposed mechanism was further investigated by in situ comparative data for the three adsorbents as measured by
FTIR during CO2 and water coadsorption on KMG30.625 TGA. The total working capacity of CO2 and water depends
Figure 57 indicates the main IR bands associated with CO2 primarily on the effect of adsorbate partial pressure on
and/or water adsorption over the proposed four sites in adsorption over site C. Decreasing temperature led to an
KMG30. increase in the cyclic capacity of water on site A and a decrease
Further work on other LDH-derived materials such as in CO2 cyclic capacity on site B for all materials. Moreover, site
KMG70 (with Mg/Al ratio of 2.95) and alumina hydrotalcite C in KMG70 exhibited the highest CO2 working capacity at
(KSORB),624 indicated that the cyclic working capacity for 300 °C.624

Figure 55. (a) TGA profile for CO2-loaded Mg−Al mixed oxide exposed to four gases in sequence at 400 °C with PH2O = 0.34 bar PCO2 = 0.66 bar.
(b) Corresponding fixed bed reactor experiment at 400 °C with PH2O = 0.1 bar and PCO2 = 0.025 bar. Reproduced with permission from ref 623.
Copyright 2017 Creative Commons Attribution License.

AT https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

solvated carbonate. In line with the finding of Baltrusaitis et


al.,627 CO2 adsorption in the presence of humidity was
suggested to take place through the intermediacy of carbonic
acid, which may undergo deprotonation and further react with
surface hydroxyl groups to generate adsorbed carbonate (eqs
23−24, where (a) indicates adsorbed species). Therefore,
consumption of hydroxyl groups was observed in both dry and
humid conditions.
MOH + H+ ⇋ M − OH 2+ (22)

MOH + H 2CO3(a) ⇋ M − OH 2+ + HCO3(a)− (23)

MOH + HCO3(a)− ⇋ MOCO2(a)− + H 2O(a) (24)


Reactions 22−24 occur at low RHs, where surface coverage of
water is limited. Higher RHs gave rise to competition between
Figure 58. CO2 and water working capacity for different materials as
CO2 and intermediate species (H2CO3 and HCO3−) for
measured by TGA at PCO2 = 0.66 bar and PH2O = 0.34 bar. EQ =
equilibrium. Reproduced with permission from ref 624. Copyright adsorption sites, as evidenced by increased conversion of
2018 Creative Commons Attribution License. surface-adsorbed carbonate to solvated carbonate. With
increasing RH, the higher water to CO2 molar ratio in the
gas phase increased the likelihood of water to interact with the
TGA data for CO2 and H2O coadsorption over KMG oxide surface and form a barrier for direct CO2 adsorption to
sorbents in the range of 300−500 °C were adequately form bicarbonate and carbonate species. Hence, at high RHs,
described by a modified Elovich equation.626 The model CO2 was adsorbed mainly in the form of solvated carbonate.168
consisted of two chemisorption sites for CO2 and H2O, with a In summary, the presence of water is essential for CO2
third site for interactions between CO2, H2O, and the adsorption on alkali metal carbonates into bicarbonates.
adsorbent. Unlike the two former sites that can be regenerated Depending on the adsorbent and adsorption conditions,
by flowing nitrogen, regeneration of the latter site was by humid CO2 adsorption over Na2CO3 may give rise to several
displacement adsorption of CO2 or water, depending of the intermediate compounds, including Wegscheider’s salt
partial pressure of each adsorbate. (Na2CO3·3NaHCO3), sodium sesquicarbonate (Na2CO3·
6.3. Other Metal Oxides NaHCO3·2H2O), Na2CO3·H2O, and Na5H3(CO3)4. The
Water also plays a determining role in the CO2 adsorption presence of humidity equally promotes CO2 adsorption on
mechanism and structural evolution of other metal oxide K2CO3, presumably through the formation of K2CO3·
adsorbents. In the absence of moisture, CO2 reacts with the (1.5H2O), considered as a highly active phase for CO2
hydroxylated surface of iron oxide to form bicarbonate, adsorption. For both adsorbents, pre-exposure of the material
whereas in the presence of water, CO2 reacts in a stepwise to water vapor and increasing humidity during CO2 adsorption
process, giving rise to surface carbonate.627 Occurrence of a enhanced CO2 uptake. The presence of water during CO2
thin layer of adsorbed water on oxide surfaces can impact their capture over CaO and MgO-based materials is beneficial to
interaction with CO2. Baltrusaitis et al.628 used FTIR to carbonation by enhancing the diffusion of CO2 through the
investigate dry and humid CO2 adsorption at 23 °C on carbonate layer as a result of increased porosity. It also
hydroxylated Fe2O3, γ-Al2O3, and TiO2 nanoparticles. The promotes the calcination of CaCO3 by reducing the temper-
main adsorption products were bicarbonate and monodentate ature for CaCO3 decomposition while increasing its rate. For
and bidentate carbonates. In dry conditions, CO2 adsorption potassium-promoted Mg-Al mixed oxide derived from hydro-
led to the formation of different surface species, depending on talcite, multiple adsorption sites with different affinity with
the basicity of the adsorption sites. For Fe2O3 and γ-Al2O3, water and/or CO2 were proposed. Moreover, FTIR data
CO2 adsorbed primarily in the form of bicarbonate, whereas showed the beneficial impact of humidity in increasing the
for TiO2, mostly carbonate was formed in addition to traces of CO2 working capacity of the material by converting the
carboxylate species. In the presence of moisture (40% RH), strongly bound bidentate carbonate species into bicarbonate
CO2 reacted with a layer of adsorbed water to form solvated during CO2 desorption. As for metal oxides with surface
hydronium carbonate. The CO2 initial adsorption rate was hydroxyl groups, CO2 adsorbs in the form of bicarbonate and/
higher in dry versus humid condition, suggesting rapid water or carbonate in dry conditions. Although water preferentially
uptake followed by slow CO2 diffusion to form solvated adsorbs on these surfaces, CO2 adsorption in the form of
carbonates. Irrespective of the presence of water or the lack solvated carbonate is enhanced in the presence of partially
thereof, the initial rate of adsorption decreased in the order humid surfaces, however, higher RH has an adverse effect on
Fe2O3 > γ-Al2O3 > TiO2. CO2 uptake.
Likewise, CO2 reacts with ZnO and CuO surface hydroxyl
groups in dry conditions to form bicarbonate.168 Additionally, 7. CONCLUSIONS AND OUTLOOK
the reaction between CO2 and surface defect sites and lattice Water is a ubiquitous component in all potential CO2 sources,
oxygen in dry conditions was associated with the formation of including air and industrial large point sources, such as flue
surface monodentate and bidentate carbonate as well as gases, biogas, and natural gas. As far as adsorption is
carboxylate species. Bicarbonate formation was found to concerned, the wide variety of CO2 separation materials
decrease with increasing humidity. Beyond 20% and 70% RH exhibit as many different behaviors in the presence of water.
for CuO and ZnO, respectively, CO2 adsorbed in the form of Depending on the material, water may promote or impede
AU https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

CO2 adsorption, enhance or decrease the rate of CO2 uptake, Even when it plays a beneficial role, the main drawback of
or even affect the structural integrity of the adsorbent. water in CO2 capture by adsorption remains the energy penalty
Therefore, understanding how water affects the performance associated with water vaporization during adsorbent regener-
of different adsorbents is of paramount importance in terms of ation. Moreover, there is often an optimum amount of
material development and process design. moisture for maximum benefit. However, because adjusting the
In the presence of physical adsorbents such as ACs, zeolites, humidity of the gas feed does not seem to be economically
and MOFs, CO2 capture at low to moderate pressures is often viable, the solution lies in the design of either the adsorbent
adversely affected by water, mostly because of competitive and/or the process. Innovative materials should be designed to
adsorption of water on preferred sites such as functional take full advantage of the beneficial effects of water without
groups in carbons, extra-framework cations in zeolites, and excessive adsorption. It would also be advantageous if as little
coordinatively unsaturated sites in MOFs, or simply by pore water as possible desorbs during regeneration.
blocking. However, there are several exceptions to this Almost all investigations dealing with the effect of water on
statement. At relatively high pressure in the presence of CO2 adsorption were limited to three-component gas mixtures,
carbon, water plays a positive role through enhanced CO2 CO2/H2O/N2, often with humidity levels below the water
solubility and formation of CO2 hydrate. Some MOFs are content of typical large point source emissions. Future
unaffected by water because they are highly hydrophobic or directions may include the following:
have distinct sites for water and CO2 adsorption or they • To undertake the next stages of material and process
possess very specific enzyme-like pockets for CO2 adsorption, development, it is essential to focus on the effect of
with little interaction with water. As a general rule, MOFs with water vapor under industrially relevant concentrations
high CO2 capacity and low Henry constant for water and for significantly longer periods of time to gain new
adsorption are potential candidates for CO2 capture from insights into the long-term behavior of adsorbents.
humid gas streams. The lack of hydrolytic stability of some
MOFs is an additional complicating factor that can be • Other species that often occur in industrial flue gases,
overcome through appropriate material design. Another such as SOx, NOx, and O2 should be included in future
potential strategy to circumvent the competitive effect of studies to understand their impact and mutual
water is to achieve kinetic separation, where the adsorption interactions with water on CO2 uptake and the stability
stage is terminated while the water adsorption front is still of the material. Such insights are needed for process
located at the entrance of the adsorption bed. simulation, optimization, and scale-up.
Because of their tolerance to water, amine-containing • Another water-related issue that needs to be addressed is
adsorbents are among the most promising materials for DAC the effect of liquid water that may occur accidentally in
and postcombustion CO2 capture. In the presence of moisture, cold spots in the presence of water-saturated feeds, or
such materials often exhibit higher CO2 uptake because of during steam regeneration. Condensed water may have a
more favorable surface chemistry, shifting from carbamate number of negative effects. Liquid water may not only
(CO2/N = 0.5) to bicarbonate (CO2/N = 1) formation and/or block the adsorption of CO2 for most adsorbents but
enhanced diffusion of CO2 within the bulk of the material, may also induce significant changes in the number and
particularly for impregnated amine-containing polymers with spatial distribution of CO2 adsorption sites, particularly
high amine content. Higher CO2 diffusion may stem from for amine-functionalized materials. As all impregnated
reduced cross-linking due to the formation of bicarbonate and tethered amines and polyamines are water-soluble,
versus carbamate, polymer swelling, unfolding of the polymer they can be easily lixiviated out of the support and either
chain, or removal of hydrogen bonding in the presence of segregate on the material outer surface or be completely
water. More importantly, water is necessary to maintain the washed out, depending on whether liquid water is
long term stability of amine-containing adsorbents, precluding evaporated or removed as such. Similarly, grafted
the formation of stable urea linkages at the expense of CO2 aminosilanes leach readily in the presence of liquid
adsorption sites. Nonetheless, unexpected water condensation, water through hydrolysis of their surface siloxane
for example, during steam regeneration, may bring about bridges, leading to either significant removal of amine
polymer leaching or trigger −Si-O-Si− bond cleavage in groups or segregation of grafted amines into islands that
amine-grafted silicas. form as the water is evaporated. Segregation of amine
As for reactive adsorbents such as metal oxides, hydroxides, groups has a negative effect at least on the rate of
and carbonates, occurrence of water during CO2 capture is adsorption because of increased diffusion resistance.
either required or strongly recommended. Water often Therefore, innovative solutions to overcome the adverse
increases the carbonation efficiency and rate through different effects of liquid water on CO2 adsorption should be
processes such as the formation of reactive intermediates or the developed. This may include material hydrophobization
promotion of the material porosity, facilitating mass transfer through grafting of alkyl chains or coating with
deeper into the adsorbent. appropriate polymers.
As a general rule, water enhances CO2 permeation and, to • Another area of intensive research activity that was not
some extent, selectivity of polymer-based membranes through addressed in this review is the combined CO2 capture
a variety of mechanisms that may involve polymer swelling and catalytic conversion into a variety of value-added
with decreased viscosity and larger free volume, increased CO2 products, including synthetic natural gas,629,630 syn-
solubility, facilitated CO2−carrier interactions, or the creation gas,629,631 and other chemicals,632,633 over dual function
of preferential CO2 transport pathways. In the presence of materials (DFM). In single capture-conversion oper-
inorganic membranes, such as zeolites, water often has an ations, selection of appropriate adsorbents is dictated
adverse effect on CO2 permeability, particularly at low primarily by the conditions of the catalytic reaction,
temperatures, because of unfavorable competitive adsorption. particularly the temperature window.629,633 Moreover,
AV https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

the CO2 conversion may be affected not only by water group at the University of Ottawa as a postdoctoral fellow to develop
but also by other feed gas impurities such as O2, SOx, cost-effective silica-based adsorbents for CO2 capture. Later, he joined
and NOx. High temperature reactions at ca. 200−400 Enerkem as an industrial Postdoctoral Fellow to conduct research on
°C, such as CO2 methanation and syngas production, catalytic conversion of syngas to valuable chemicals. Currently, he is a
require metal oxide adsorbents, typically CaO, combined surface scientist at Ingevity, working on the characterization of a
with a catalyst, often a transition metal such as Ni or variety of adsorbents using experimental and theoretical methods.
Ru.634 Although nickel is an excellent and less costly Abdelhamid Sayari is Professor of Chemistry at the University of
methanation catalyst, it is readily deactivated through Ottawa since 2001. In 1983−1985, he was a Postdoctoral Fellow at
oxidation in the presence of typical flue gas containing the University of Pittsburgh, working on Fischer−Tropsch catalysis.
water and oxygen.634,635 CO2 conversion at lower Then he moved to the National Research Council of Canada
temperature, below ca. 120 °C, requires low temperature (Ottawa) as Research Associate to work on the genesis of metallic
adsorbents such as MOFs, polymers, and amine- nanoparticles in zeolites before joining Université Laval (Québec) as
containing materials.631,633 Under such conditions, Professor of Chemical Engineering. He was a Tier-1 Canada Research
water may be present not only in the gas phase, but Chair (2002−2015) in Catalysis by mesostructured materials and
also in liquid phase, if the reaction is carried out at high founding Director of the University of Ottawa’s Center for Catalysis
enough pressure.636 For example, in the popular reaction Research and Innovation. Abdel is currently a leading scientist in
of CO2 with epoxides, water may adversely affect the postcombustion CO2 capture. He published more than 280 papers on
yield of cyclic carbonate through the formation of porous materials, catalysis, and adsorption, in addition to eight
diols.636,637 Overall, the area of combined CO2 capture patents, and edited four books on nanoporous materials.
and conversion is still wide open, with many research
opportunities to explore the advantages and disadvan-
tages of single versus two-step operations and investigate ACKNOWLEDGMENTS
the role of water on the state of the catalyst under A.S. thanks the Natural Sciences and Engineering Research
working condition and the reaction outcome. Council of Canada (NSERC) for a Discovery Grant and a
Collaborative R&D Grant (CRDPJ 543657-19).
AUTHOR INFORMATION
Corresponding Author REFERENCES
Abdelhamid Sayari − Centre for Catalysis Research and (1) Houghton, J. Global Warming: The Complete Briefing, 5th ed.;
Innovation, Department of Chemistry and Biomolecular Cambridge University Press, 2015.
Sciences, University of Ottawa, Ottawa, Ontario K1N 6N5, (2) Global Energy Review 2020: The Impacts of the COVID-19 Crisis
Canada; orcid.org/0000-0002-5424-6466; on Global Energy Demand and CO2 Emissions; International Energy
Email: abdel.sayari@uottawa.ca Agency, 2020; https://www.iea.org/reports/global-energy-review-
2020/global-energy-and-CO2-emissions-in-2020 (accessed 2020-06-
Authors 08).
Joel M. Kolle − Centre for Catalysis Research and Innovation, (3) Le Quéré, C.; Jackson, R. B.; Jones, M. W.; Smith, A. J. P.;
Department of Chemistry and Biomolecular Sciences, Abernethy, S.; Andrew, R. M.; De-Gol, A. J.; Willis, D. R.; Shan, Y.;
Canadell, J. G.; Friedlingstein, P.; Creutzig, F.; Peters, G. P.
University of Ottawa, Ottawa, Ontario K1N 6N5, Canada
Temporary Reduction in Daily Global CO2 Emissions during the
Mohammadreza Fayaz − Centre for Catalysis Research and COVID-19 Forced Confinement. Nat. Clim. Change 2020, 10, 647−
Innovation, Department of Chemistry and Biomolecular 653.
Sciences, University of Ottawa, Ottawa, Ontario K1N 6N5, (4) Tans, P. Carbon Dioxide. NASA Global Climate Change; NASA
Canada Jet Propulsion Laboratory, 2020; https://climate.nasa.gov/vital-signs/
Complete contact information is available at: carbon-dioxide/ (accessed 2020-06-10).
(5) 2015 Paris Agreement (COP21); United Nations Framework
https://pubs.acs.org/10.1021/acs.chemrev.0c00762 Convention on Climate Change (UNFCCC)https://unfccc.int/files/
essential_background/convention/application/pdf/english_paris_
Notes agreement.pdf, 2015 (accessed 2020-06-10).
The authors declare no competing financial interest. (6) Masson-Delmotte, V.; Zhai, P.; Pörtner, H. O.; Roberts, D. J.;
Skea, P. R.; Shukla Pirani, A.; Moufouma-Okia, W.; Péan, C.; Pidcock,
Biographies R.; Connors, S.; Matthews, J. B. R.; Chen, Y.; Zhou, X.; Gomis, M. I.;
Joel M. Kolle obtained his B.S. (2007) and M.S. (2012) degrees in Lonnoy, E.; Maycock, T.; Tignor, M.; Waterfield, T. Global Warming
Chemistry from the University of Buea, Cameroon. He obtained a of 1.5 °C. An IPCC Special Report: Summary for Policymakers;
Intergovernmental Panel on Climate Change (IPCC): Geneva, 2018.
second M.S. degree in Advanced Materials and Processes (2015) from
(7) Lecomte, F.; Broutin, P.; Lebas, E. CO2 Capture: Technologies to
the University of Erlangen−Nürnberg, Germany, and a Ph.D. degree Reduce Greenhouse Gas Emissions; Editions Technips, 2010.
in Chemistry (2020) from the University of Ottawa under the (8) Papadopoulos, A. I., Seferlis, P., Eds.; Process Systems and
supervision of Prof. Abdelhamid Sayari. His doctoral work focused on Materials for CO2 Capture.Wiley and Sons, 2017.
the synthesis of porous organosilicas for CO2 adsorption and fixation (9) D’Alessandro, D. M.; Smit, B.; Long, J. R. Carbon Dioxide
to valued chemicals. He is currently a Research Associate in Prof. Capture: Prospects for New Materials. Angew. Chem., Int. Ed. 2010,
Sayari’s group, where he is looking at enhancing the oxidation/ 49, 6058−6082.
leaching stability and low-cost synthesis of solid amine CO2 sorbents. (10) Markewitz, P.; Kuckshinrichs, W.; Leitner, W.; Linssen, J.;
Zapp, P.; Bongartz, R.; Schreiber, A.; Müller, T. E. Worldwide
Mohammadreza Fayaz completed his Ph.D. in Environmental Innovations in the Development of Carbon Capture Technologies
Engineering at the University of Alberta, where he studied the and the Utilization of CO2. Energy Environ. Sci. 2012, 5, 7281−7305.
methods used to minimize irreversible adsorption of VOCs on (11) Kenarsari, S. D.; Yang, D.; Jiang, G.; Zhang, S.; Wang, J.;
activated carbon. Following his Ph.D., he joined Prof. Sayari’s research Russell, A. G.; Wei, Q.; Fan, M. Review of Recent Advances in

AW https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Carbon Dioxide Separation and Capture. RSC Adv. 2013, 3, 22739− Status of MHI CO2 Capture Plant Technology, 500 TPD CCS
22773. Demonstration of Test Results and Reliable Technologies Applied to
(12) Sreenivasulu, B.; Gayatri, D. V.; Sreedhar, I.; Raghavan, K. V. A Coal Fired Flue Gas. Energy Procedia 2014, 63, 6120−6128.
Journey into the Process and Engineering Aspects of Carbon Capture (30) Sjostrom, S.; Krutka, H. Evaluation of Solid Sorbents as a
Technologies. Renewable Sustainable Energy Rev. 2015, 41, 1324− Retrofit Technology for CO2 Capture. Fuel 2010, 89, 1298−1306.
1350. (31) Lu, A. H.; Hao, G. P. Porous Materials for Carbon Dioxide
(13) Yuan, Z.; Eden, M. R.; Gani, R. Toward the Development and Capture. Annu. Rep. Prog. Chem., Sect. A: Inorg. Chem. 2013, 109,
Deployment of Large-Scale Carbon Dioxide Capture and Conversion 484−503.
Processes. Ind. Eng. Chem. Res. 2016, 55, 3383−3419. (32) Nie, L.; Mu, Y.; Jin, J.; Chen, J.; Mi, J. Recent Developments
(14) Gibson, J. A. A.; Mangano, E.; Shiko, E.; Greenaway, A. G.; and Consideration Issues in Solid Adsorbents for CO2 Capture from
Gromov, A. V.; Lozinska, M. M.; Friedrich, D.; Campbell, E. E. B.; Flue Gas. Chin. J. Chem. Eng. 2018, 26, 2303−2317.
Wright, P. A.; Brandani, S. Adsorption Materials and Processes for (33) Patel, H. A.; Byun, J.; Yavuz, C. T. Carbon Dioxide Capture
Carbon Capture from Gas-Fired Power Plants: AMPGas. Ind. Eng. Adsorbents: Chemistry and Methods. ChemSusChem 2017, 10, 1303−
Chem. Res. 2016, 55, 3840−3851. 1317.
(15) Oschatz, M.; Antonietti, M. A Search for Selectivity to Enable (34) Khraisheh, M.; Mukherjee, S.; Kumar, A.; Al Momani, F.;
CO2 Capture with Porous Adsorbents. Energy Environ. Sci. 2018, 11, Walker, G.; Zaworotko, M. J. An Overview on Trace CO2 Removal
57−70. by Advanced Physisorbent Materials. J. Environ. Manage. 2020, 255,
(16) Bui, M.; Adjiman, C. S.; Bardow, A.; Anthony, E. J.; Boston, A.; 109874.
Brown, S.; Fennell, P. S.; Fuss, S.; Galindo, A.; Hackett, L. A.; Hallett, (35) Sumida, K.; Rogow, D. L.; Mason, J. A.; McDonald, T. M.;
J. P.; Herzog, H. J.; Jackson, G.; Kemper, J.; Krevor, S.; Maitland, G. Bloch, E. D.; Herm, Z. R.; Bae, T.-H.; Long, J. R. Carbon Dioxide
C.; Matuszewski, M.; Metcalfe, I. S.; Petit, C.; Puxty, G.; Reimer, J.; Capture in Metal-Organic Frameworks. Chem. Rev. 2012, 112, 724−
Reiner, D. M.; Rubin, E. S.; Scott, S. A.; Shah, N.; Smit, B.; Trusler, J. 781.
P. M.; Webley, P.; Wilcox, J.; Mac Dowell, N. Carbon Capture and (36) Bae, Y. S.; Snurr, R. Q. Development and Evaluation of Porous
Storage (CCS): The Way Forward. Energy Environ. Sci. 2018, 11, Materials for Carbon Dioxide Separation and Capture. Angew. Chem.,
1062−1176. Int. Ed. 2011, 50, 11586−11596.
(17) Kolster, C.; Masnadi, M. S.; Krevor, S.; Mac Dowell, N.; (37) Zhang, Z.; Zhao, Y.; Gong, Q.; Li, Z.; Li, J. MOFs for CO2
Brandt, A. R. CO2 Enhanced Oil Recovery: A Catalyst for Gigatonne- Capture and Separation from Flue Gas Mixtures: The Effect of
Scale Carbon Capture and Storage Deployment? Energy Environ. Sci. Multifunctional Sites on Their Adsorption Capacity and Selectivity.
2017, 10, 2594−2608. Chem. Commun. 2013, 49, 653−661.
(18) Kelemen, P.; Benson, S. M.; Pilorgé, H.; Psarras, P.; Wilcox, J. (38) Shalini, S.; Nandi, S.; Justin, A.; Maity, R.; Vaidhyanathan, R.
An Overview of the Status and Challenges of CO2 Storage in Potential of Ultramicroporous Metal-Organic Frameworks in CO2
Minerals and Geological Formations. Front. Clim. 2019, 1, 9. Clean-Up. Chem. Commun. 2018, 54, 13472−13490.
(19) Abanades, J. C.; Rubin, E. S.; Mazzotti, M.; Herzog, H. J. On (39) Andirova, D.; Cogswell, C. F.; Lei, Y.; Choi, S. Effect of the
the Climate Change Mitigation Potential of CO2 Conversion to Structural Constituents of Metal Organic Frameworks on Carbon
Fuels. Energy Environ. Sci. 2017, 10, 2491−2499. Dioxide Capture. Microporous Mesoporous Mater. 2016, 219, 276−
(20) Otto, A.; Grube, T.; Schiebahn, S.; Stolten, D. Closing the 305.
Loop: Captured CO2 as a Feedstock in the Chemical Industry. Energy (40) Ding, M.; Flaig, R. W.; Jiang, H. L.; Yaghi, O. M. Carbon
Environ. Sci. 2015, 8, 3283−3297. Capture and Conversion Using Metal-Organic Frameworks and
(21) Hepburn, C.; Adlen, E.; Beddington, J.; Carter, E. A.; Fuss, S.; MOF-Based Materials. Chem. Soc. Rev. 2019, 48, 2783−2828.
Mac Dowell, N.; Minx, J. C.; Smith, P.; Williams, C. K. The (41) Danaci, D.; Bui, M.; Mac Dowell, N.; Petit, C. Exploring the
Technological and Economic Prospects for CO2 Utilization and Limits of Adsorption-Based CO2 Capture Using MOFs with PVSA -
Removal. Nature 2019, 575, 87−97. from Molecular Design to Process Economics. Mol. Syst. Des. Eng.
(22) Feron, P. H. M. Absorption-Based Post-Combustion Capture of 2020, 5, 212−231.
Carbon Dioxide; Feron, P. H. M., Ed.; Elsevier, 2016. (42) Choi, H. S.; Suh, M. P. Highly Selective CO2 Capture in
(23) Oexmann, J.; Kather, A.; Linnenberg, S.; Liebenthal, U. Post- Flexible 3D Coordination Polymer Networks. Angew. Chem., Int. Ed.
Combustion CO2 Capture: Chemical Absorption Processes in Coal- 2009, 48, 6865−6869.
Fired Steam Power Plants. Greenhouse Gases: Sci. Technol. 2012, 2, (43) Piscopo, C. G.; Loebbecke, S. Strategies to Enhance Carbon
80−98. Dioxide Capture in Metal-Organic Frameworks. ChemPlusChem 2020,
(24) Nielsen, C. J.; Herrmann, H.; Weller, C. Atmospheric 85, 538−547.
Chemistry and Environmental Impact of the Use of Amines in (44) Patel, H. A.; Je, S. H.; Park, J.; Chen, D. P.; Jung, Y.; Yavuz, C.
Carbon Capture and Storage (CCS). Chem. Soc. Rev. 2012, 41, 6684− T.; Coskun, A. Unprecedented High-Temperature CO2 Selectivity in
6704. N2-Phobic Nanoporous Covalent Organic Polymers. Nat. Commun.
(25) Poste, A. E.; Grung, M.; Wright, R. F. Amines and Amine- 2013, 4, 1357.
Related Compounds in Surface Waters: A Review of Sources, (45) Xiang, Z.; Zhou, X.; Zhou, C.; Zhong, S.; He, X.; Qin, C.; Cao,
Concentrations and Aquatic Toxicity. Sci. Total Environ. 2014, 481, D. Covalent-Organic Polymers for Carbon Dioxide Capture. J. Mater.
274−279. Chem. 2012, 22, 22663−22669.
(26) Reynolds, A. J.; Verheyen, T. V.; Adeloju, S. B.; Meuleman, E.; (46) Wang, H.; Jiang, D.; Huang, D.; Zeng, G.; Xu, P.; Lai, C.; Chen,
Feron, P. Towards Commercial Scale Postcombustion Capture of M.; Cheng, M.; Zhang, C.; Wang, Z. Covalent Triazine Frameworks
CO2 with Monoethanolamine Solvent: Key Considerations for for Carbon Dioxide Capture. J. Mater. Chem. A 2019, 7, 22848−
Solvent Management and Environmental Impacts. Environ. Sci. 22870.
Technol. 2012, 46, 3643−3654. (47) Liu, Q.; Pham, T.; Porosoff, M. D.; Lobo, R. F. ZK-5: A CO2-
(27) Zhao, B.; Sun, Y.; Yuan, Y.; Gao, J.; Wang, S.; Zhuo, Y.; Chen, Selective Zeolite with High Working Capacity at Ambient Temper-
C. Study on Corrosion in CO2 Chemical Absorption Process Using ature and Pressure. ChemSusChem 2012, 5, 2237−2242.
Amine Solution. Energy Procedia 2011, 4, 93−100. (48) Hudson, M. R.; Queen, W. L.; Mason, J. A.; Fickel, D. W.;
(28) Gunasekaran, P.; Veawab, A.; Aroonwilas, A. Corrosivity of Lobo, R. F.; Brown, C. M. Unconventional, Highly Selective CO2
Amine-Based Absorbents for CO2 Capture. Energy Procedia 2017, Adsorption in Zeolite SSZ-13. J. Am. Chem. Soc. 2012, 134, 1970−
114, 2047−2054. 1973.
(29) Hirata, T.; Nagayasu, H.; Yonekawa, T.; Inui, M.; Kamijo, T.; (49) Aschenbrenner, O.; McGuire, P.; Alsamaq, S.; Wang, J.;
Kubota, Y.; Tsujiuchi, T.; Shimada, D.; Wall, T.; Thomas, J. Current Supasitmongkol, S.; Al-Duri, B.; Styring, P.; Wood, J. Adsorption of

AX https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Carbon Dioxide on Hydrotalcite-Like Compounds of Different (72) Mohammad, S. A.; Gasem, K. A. M. Modeling the Competitive
Compositions. Chem. Eng. Res. Des. 2011, 89, 1711−1721. Adsorption of CO2 and Water at High Pressures on Wet Coals.
(50) Bhatta, L. K. G.; Subramanyam, S.; Chengala, M. D.; Bhatta, U. Energy Fuels 2012, 26, 557−568.
M.; Venkatesh, K. Enhancement in CO2 Adsorption on Hydrotalcite- (73) Li, G.; Xiao, P.; Webley, P. A.; Zhang, J.; Singh, R. Competition
Based Material by Novel Carbon Support Combined with K2CO3 of CO2/H2O in Adsorption Based CO2 Capture. Energy Procedia
Impregnation. Ind. Eng. Chem. Res. 2015, 54, 10876−10884. 2009, 1, 1123−1130.
(51) Elkhalifah, A. E. I.; Maitra, S.; Bustam, M. A.; Murugesan, T. (74) Lee, S. C.; Chae, H. J.; Lee, S. J.; Choi, B. Y.; Yi, C. K.; Lee, J.
Effects of Exchanged Ammonium Cations on Structure Character- B.; Ryu, C. K.; Kim, J. C. Development of Regenerable MgO-Based
istics and CO2 Adsorption Capacities of Bentonite Clay. Appl. Clay Sorbent Promoted with K2CO3 for CO2 Capture at Low
Sci. 2013, 83−84, 391−398. Temperatures. Environ. Sci. Technol. 2008, 42, 2736−2741.
(52) Shen, W.; Fan, W. Nitrogen-Containing Porous Carbons: (75) Xu, X.; Song, C.; Miller, B. G.; Scaroni, A. W. Influence of
Synthesis and Application. J. Mater. Chem. A 2013, 1, 999−1013. Moisture on CO2 Separation from Gas Mixture by a Nanoporous
(53) Li, P. Z.; Zhao, Y. Nitrogen-Rich Porous Adsorbents for CO2 Adsorbent Based on Polyethylenimine-Modified Molecular Sieve
Capture and Storage. Chem. - Asian J. 2013, 8, 1680−1691. MCM-41. Ind. Eng. Chem. Res. 2005, 44, 8113−8119.
(54) Creamer, A. E.; Gao, B. Carbon-Based Adsorbents for (76) Harlick, P. J. E.; Sayari, A. Applications of Pore-Expanded
Postcombustion CO2 Capture: A Critical Review. Environ. Sci. Mesoporous Silica. 5. Triamine Grafted Material with Exceptional
Technol. 2016, 50, 7276−7289. CO2 Dynamic and Equilibrium Adsorption Performance. Ind. Eng.
(55) Sethia, G.; Sayari, A. Comprehensive Study of Ultra- Chem. Res. 2007, 46, 446−458.
Microporous Nitrogen-Doped Activated Carbon for CO2 Capture. (77) Sayari, A.; Belmabkhout, Y. Stabilization of Amine-Containing
Carbon 2015, 93, 68−80. CO2 Adsorbents: Dramatic Effect of Water Vapor. J. Am. Chem. Soc.
(56) Sethia, G.; Sayari, A. Nitrogen-Doped Carbons: Remarkably 2010, 132, 6312−6314.
Stable Materials for CO2 Capture. Energy Fuels 2014, 28, 2727−2731. (78) Sayari, A.; Belmabkhout, Y. Stabilization of Amine-Containing
(57) Belmabkhout, Y.; Serna-Guerrero, R.; Sayari, A. Adsorption of CO2 Adsorbents and Related Systems and Methods. U.S. Patent
CO2 from Dry Gases on MCM-41 Silica at Ambient Temperature US10112174, 2018.
and High Pressure. 1: Pure CO2 Adsorption. Chem. Eng. Sci. 2009, (79) Jahandar Lashaki, M.; Khiavi, S.; Sayari, A. Stability of Amine-
64, 3721−3728. Functionalized CO2 Adsorbents: A Multifaceted Puzzle. Chem. Soc.
(58) Belmabkhout, Y.; Sayari, A. Adsorption of CO2 from Dry Gases Rev. 2019, 48, 3320−3405.
on MCM-41 Silica at Ambient Temperature and High Pressure. 2: (80) Yang, R. T. Adsorbents: Fundamentals and Applications; Wiley &
Adsorption of CO2/N2, CO2/CH4 and CO2/H2 Binary Mixtures. Sons, 2003.
Chem. Eng. Sci. 2009, 64, 3729−3735. (81) Lin, R.; Ladshaw, A.; Nan, Y.; Liu, J.; Yiacoumi, S.; Tsouris, C.;
(59) Radfarnia, H. R.; Sayari, A. A Highly Efficient CaO-Based CO2 DePaoli, D. W.; Tavlarides, L. L. Isotherms for Water Adsorption on
Sorbent Prepared by a Citrate-Assisted Sol-Gel Technique. Chem. Molecular Sieve 3A: Influence of Cation Composition. Ind. Eng.
Eng. J. 2015, 262, 913−920. Chem. Res. 2015, 54, 10442−10448.
(60) Kierzkowska, A. M.; Pacciani, R.; Müller, C. R. CaO-Based (82) Castillo, J. M.; Silvestre-Albero, J.; Rodriguez-Reinoso, F.;
CO2 Sorbents: From Fundamentals to the Development of New, Vlugt, T. J. H.; Calero, S. Water Adsorption in Hydrophilic Zeolites:
Highly Effective Materials. ChemSusChem 2013, 6, 1130−1148. Experiment and Simulation. Phys. Chem. Chem. Phys. 2013, 15,
(61) Sayari, A.; Belmabkhout, Y.; Serna-Guerrero, R. Flue Gas 17374−17382.
Treatment via CO2 Adsorption. Chem. Eng. J. 2011, 171, 760−774. (83) Rege, S. U.; Yang, R. T. A Novel FTIR Method for Studying
(62) Choi, S.; Drese, J. H.; Jones, C. W. Adsorbent Materials for Mixed Gas Adsorption at Low Concentrations: H2O and CO2 on
Carbon Dioxide Capture from Large Anthropogenic Point Sources. NaX Zeolite and γ-Alumina. Chem. Eng. Sci. 2001, 56, 3781−3796.
ChemSusChem 2009, 2, 796−854. (84) Brandani, F.; Ruthven, D. M. The Effect of Water on the
(63) Bollini, P.; Didas, S. A.; Jones, C. W. Amine-Oxide Hybrid Adsorption of CO2 and C3H8 on Type X Zeolites. Ind. Eng. Chem.
Materials for Acid Gas Separations. J. Mater. Chem. 2011, 21, 15100− Res. 2004, 43, 8339−8344.
15120. (85) Sun, Y.; Wang, Y.; Zhang, Y.; Zhou, Y.; Zhou, L. CO2 Sorption
(64) Zhao, X.; Cui, Q.; Wang, B.; Yan, X.; Singh, S.; Zhang, F.; Gao, in Activated Carbon in the Presence of Water. Chem. Phys. Lett. 2007,
X.; Li, Y. Recent Progress of Amine Modified Sorbents for Capturing 437, 14−16.
CO2 from Flue Gas. Chin. J. Chem. Eng. 2018, 26, 2292−2302. (86) Tan, K.; Nijem, N.; Gao, Y.; Zuluaga, S.; Li, J.; Thonhauser, T.;
(65) Gelles, T.; Lawson, S.; Rownaghi, A. A.; Rezaei, F. Recent Chabal, Y. J. Water Interactions in Metal Organic Frameworks.
Advances in Development of Amine Functionalized Adsorbents for CrystEngComm 2015, 17, 247−260.
CO2 Capture. Adsorption 2020, 26, 5−50. (87) Guo, P.; Dutta, D.; Wong-Foy, A. G.; Gidley, D. W.; Matzger,
(66) Samanta, A.; Zhao, A.; Shimizu, G. K. H.; Sarkar, P.; Gupta, R. A. J. Water Sensitivity in Zn4O-Based MOFs Is Structure and History
Post-Combustion CO2 Capture Using Solid Sorbents: A Review. Ind. Dependent. J. Am. Chem. Soc. 2015, 137, 2651−2657.
Eng. Chem. Res. 2012, 51, 1438−1463. (88) Canivet, J.; Fateeva, A.; Guo, Y.; Coasne, B.; Farrusseng, D.
(67) Chen, C.; Zhang, S.; Row, K. H.; Ahn, W. S. Amine-Silica Water Adsorption in MOFs: Fundamentals and Applications. Chem.
Composites for CO2 Capture: A Short Review. J. Energy Chem. 2017, Soc. Rev. 2014, 43, 5594−5617.
26, 868−880. (89) Kaye, S. S.; Dailly, A.; Yaghi, O. M.; Long, J. R. Impact of
(68) Hu, X. E.; Liu, L.; Luo, X.; Xiao, G.; Shiko, E.; Zhang, R.; Fan, Preparation and Handling on the Hydrogen Storage Properties of
X.; Zhou, Y.; Liu, Y.; Zeng, Z.; Li, C. A Review of N-Functionalized Zn4O(1,4-Benzenedicarboxylate)3 (MOF-5). J. Am. Chem. Soc. 2007,
Solid Adsorbents for Post-Combustion CO2 Capture. Appl. Energy 129, 14176−14177.
2020, 260, 114244. (90) Cavka, J. H.; Jakobsen, S.; Olsbye, U.; Guillou, N.; Lamberti,
(69) Zevenhoven, R.; Kilpinen, P. Flue Gases and Fuel Gases. In C.; Bordiga, S.; Lillerud, K. P. A New Zirconium Inorganic Building
Control of Pollutants in Flue Gases and Fuel Gases; Helsinki University Brick Forming Metal Organic Frameworks with Exceptional Stability.
of Technology: Helsinki, 2001; Chapter 2. J. Am. Chem. Soc. 2008, 130, 13850−13851.
(70) Biogas Composition. In The Biogas; Naskeo Environnement, (91) Hong, D. Y.; Hwang, Y. K.; Serre, C.; Férey, G.; Chang, J. S.
2009; https://www.biogas-renewable-energy.info/biogas_ Porous Chromium Terephthalate MIL-101 with Coordinatively
composition.html (accessed 2020-05-10). Unsaturated Sites: Surface Functionalization, Encapsulation, Sorption
(71) Li, G.; Xiao, P.; Webley, P. Binary Adsorption Equilibrium of and Catalysis. Adv. Funct. Mater. 2009, 19, 1537−1552.
Carbon Dioxide and Water Vapor on Activated Alumina. Langmuir (92) Saccoccia, B.; Bohnsack, A. M.; Waggoner, N. W.; Cho, K. H.;
2009, 25, 10666−10675. Lee, J. S.; Hong, D. Y.; Lynch, V. M.; Chang, J. S.; Humphrey, S. M.

AY https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Separation of P-Divinylbenzene by Selective Room-Temperature (110) Sujan, A. R.; Kumar, D. R.; Sakwa-Novak, M.; Ping, E. W.;
Adsorption Inside Mg-CUK-1 Prepared by Aqueous Microwave Hu, B.; Park, S. J.; Jones, C. W. Poly(Glycidyl Amine)-Loaded SBA-
Synthesis. Angew. Chem., Int. Ed. 2015, 54, 5394−5398. 15 Sorbents for CO2 Capture from Dilute and Ultradilute Gas
(93) Samokhvalov, A. Aluminum Metal-Organic Frameworks for Mixtures. ACS Appl. Polym. Mater. 2019, 1, 3137−3147.
Sorption in Solution: A Review. Coord. Chem. Rev. 2018, 374, 236− (111) Lee, J. J.; Sievers, C.; Jones, C. W. Silica-Supported Hindered
253. Aminopolymers for CO2 Capture. Ind. Eng. Chem. Res. 2019, 58,
(94) Wang, S.; Serre, C. Toward Green Production of Water-Stable 22551−22560.
Metal-Organic Frameworks Based on High-Valence Metals with Low (112) Ma, X.; Wang, X.; Song, C. Molecular Basket” Sorbents for
Toxicities. ACS Sustainable Chem. Eng. 2019, 7, 11911−11927. Separation of CO2 and H2S from Various Gas Streams. J. Am. Chem.
(95) Yuan, S.; Feng, L.; Wang, K.; Pang, J.; Bosch, M.; Lollar, C.; Soc. 2009, 131, 5777−5783.
Sun, Y.; Qin, J.; Yang, X.; Zhang, P.; Wang, Q.; Zou, L.; Zhang, Y.; (113) Xu, X.; Song, C.; Miller, B. G.; Scaroni, A. W. Adsorption
Zhang, L.; Fang, Y.; Li, J.; Zhou, H. C. Stable Metal-Organic Separation of Carbon Dioxide from Flue Gas of Natural Gas-Fired
Frameworks: Design, Synthesis, and Applications. Adv. Mater. 2018, Boiler by a Novel Nanoporous “Molecular Basket” Adsorbent. Fuel
30, 1704303. Process. Technol. 2005, 86, 1457−1472.
(96) Burtch, N. C.; Jasuja, H.; Walton, K. S. Water Stability and (114) Klinthong, W.; Huang, C. H.; Tan, C. S. One-Pot Synthesis
Adsorption in Metal-Organic Frameworks. Chem. Rev. 2014, 114, and Pelletizing of Polyethylenimine-Containing Mesoporous Silica
10575−10612. Powders for CO2 Capture. Ind. Eng. Chem. Res. 2016, 55, 6481−6491.
(97) Howarth, A. J.; Liu, Y.; Li, P.; Li, Z.; Wang, T. C.; Hupp, J. T.; (115) Goeppert, A.; Czaun, M.; May, R. B.; Prakash, G. K. S.; Olah,
Farha, O. K. Chemical, Thermal and Mechanical Stabilities of Metal- G. A.; Narayanan, S. R. Carbon Dioxide Capture from the Air Using a
Organic Frameworks. Nat. Rev. Mater. 2016, 1, 15018. Polyamine Based Regenerable Solid Adsorbent. J. Am. Chem. Soc.
(98) Yaghi, O. M.; Kalmutzki, M. J.; Diercks, C. S. Water Sorption 2011, 133, 20164−20167.
Applications of MOFs. In Introduction to Reticular Chemistry: Metal- (116) Min, K.; Choi, W.; Kim, C.; Choi, M. Oxidation-Stable
Organic Frameworks and Covalent Organic Frameworks; Wiley & Sons, Amine-Containing Adsorbents for Carbon Dioxide Capture. Nat.
2019; Chapter 17, pp 395−427. Commun. 2018, 9, 726.
(99) Yazaydın, A. Ö .; Benin, A. I.; Faheem, S. A.; Jakubczak, P.; (117) Heydari-Gorji, A.; Yang, Y.; Sayari, A. Effect of the Pore
Low, J. J.; Willis, R. R.; Snurr, R. Q. Enhanced CO2 Adsorption in Length on CO2 Adsorption over Amine-Modified Mesoporous
Metal-Organic Frameworks via Occupation of Open-Metal Sites by Silicas. Energy Fuels 2011, 25, 4206−4210.
Coordinated Water Molecules. Chem. Mater. 2009, 21, 1425−1430. (118) Zhang, Z.; Ma, X.; Wang, D.; Song, C.; Wang, Y.
(100) Liu, J.; Wang, Y.; Benin, A. I.; Jakubczak, P.; Willis, R. R.; Development of Silica-Gel-Supported Polyethylenimine Sorbents for
LeVan, M. D. CO2/H2O Adsorption Equilibrium and Rates on CO2 Capture from Flue Gas. AIChE J. 2012, 58, 2495−2502.
Metal-Organic Frameworks: HKUST-1 and Ni/DOBDC. Langmuir (119) Dao, D. S.; Yamada, H.; Yogo, K. Large-Pore Mesostructured
2010, 26, 14301−14307. Silica Impregnated with Blended Amines for CO2 Capture. Ind. Eng.
(101) Grajciar, L.; Wiersum, A. D.; Llewellyn, P. L.; Chang, J. S.; Chem. Res. 2013, 52, 13810−13817.
Nachtigall, P. Understanding CO2 Adsorption in CuBTC MOF: (120) Gadipelli, S.; Patel, H. A.; Guo, Z. An Ultrahigh Pore Volume
Comparing Combined DFT-Ab Initio Calculations with Micro- Drives Up the Amine Stability and Cyclic CO2 Capacity of a Solid-
calorimetry Experiments. J. Phys. Chem. C 2011, 115, 17925−17933. Amine@Carbon Sorbent. Adv. Mater. 2015, 27, 4903−4909.
(102) Serre, C.; Millange, F.; Thouvenot, C.; Noguès, M.; Marsolier, (121) Gibson, J. A. A.; Gromov, A. V.; Brandani, S.; Campbell, E. E.
G.; Louër, D.; Férey, G. Very Large Breathing Effect in the First B. The Effect of Pore Structure on the CO2 Adsorption Efficiency of
Nanoporous Chromium(III)-Based Solids: MIL-53 or Cr III (OH)· Polyamine Impregnated Porous Carbons. Microporous Mesoporous
{O2C-C6H4-CO2 }·{HO2C-C6H4-CO2H} X·H2Oy. J. Am. Chem. Mater. 2015, 208, 129−139.
Soc. 2002, 124, 13519−13526. (122) Wang, D.; Ma, X.; Sentorun-Shalaby, C.; Song, C. Develop-
(103) Llewellyn, P. L.; Bourrelly, S.; Serre, C.; Filinchuk, Y.; Férey, ment of Carbon-Based “Molecular Basket” Sorbent for CO2 Capture.
G. How Hydration Drastically Improves Adsorption Selectivity for Ind. Eng. Chem. Res. 2012, 51, 3048−3057.
CO2 over CH4 in the Flexible Chromium Terephthalate MIL-53. (123) Su, F.; Lu, C.; Kuo, S. C.; Zeng, W. Adsorption of CO2 on
Angew. Chem., Int. Ed. 2006, 45, 7751−7754. Amine-Functionalized Y-Type Zeolites. Energy Fuels 2010, 24, 1441−
(104) Song, J. H.; Kim, D. W.; Kang, D. W.; Lee, W. R.; Hong, C. S. 1448.
Humidity-Triggered Single-Crystal-to-Single-Crystal Structural Trans- (124) Chen, C.; Kim, S. S.; Cho, W. S.; Ahn, W. S.
formations in a Zn(II) Coordination Polymer Displaying Unusual Polyethylenimine-Incorporated Zeolite 13X with Mesoporosity for
Activation Energy Change in Proton Conductivity. Chem. Commun. Post-Combustion CO2 Capture. Appl. Surf. Sci. 2015, 332, 167−171.
2019, 55, 9713−9716. (125) Kim, Y. K.; Mo, Y. H.; Lee, J.; You, H. S.; Yi, C. K.; Park, Y.
(105) Shin, S.; Jeong, S.; Kim, D.; Lah, M. S. Reversible Single- C.; Park, S. E. Tetraethylenepentamine Embedded Zeolite A for
Crystal-to-Single-Crystal Transformations of Metal-Organic Frame- Carbon Dioxide Adsorption. J. Nanosci. Nanotechnol. 2013, 13, 2703−
works That Accompany Two-Dimensional Framework Reorganiza- 2707.
tions. Cryst. Growth Des. 2017, 17, 2228−2237. (126) Xian, S.; Xu, F.; Ma, C.; Wu, Y.; Xia, Q.; Wang, H.; Li, Z.
(106) Heydari-Gorji, A.; Sayari, A. Thermal, Oxidative, and CO2- Vapor-Enhanced CO2 Adsorption Mechanism of Composite PEI@
Induced Degradation of Supported Polyethylenimine Adsorbents. Ind. ZIF-8 Modified by Polyethyleneimine for CO2/N2 Separation. Chem.
Eng. Chem. Res. 2012, 51, 6887−6894. Eng. J. 2015, 280, 363−369.
(107) Chen, C.; Son, W. J.; You, K. S.; Ahn, J. W.; Ahn, W. S. (127) Martínez, F.; Sanz, R.; Orcajo, G.; Briones, D.; Yángüez, V.
Carbon Dioxide Capture Using Amine-Impregnated HMS Having Amino-Impregnated MOF Materials for CO2 Capture at Post-
Textural Mesoporosity. Chem. Eng. J. 2010, 161, 46−52. Combustion Conditions. Chem. Eng. Sci. 2016, 142, 55−61.
(108) Sanz, R.; Calleja, G.; Arencibia, A.; Sanz-Pérez, E. S. CO2 (128) Lin, Y.; Lin, H.; Wang, H.; Suo, Y.; Li, B.; Kong, C.; Chen, L.
Capture with Pore-Expanded MCM-41 Silica Modified with Amino Enhanced Selective CO2 Adsorption on Polyamine/MIL-101(Cr)
Groups by Double Functionalization. Microporous Mesoporous Mater. Composites. J. Mater. Chem. A 2014, 2, 14658−14665.
2015, 209, 165−171. (129) Belmabkhout, Y.; Sayari, A. Effect of Pore Expansion and
(109) Choi, W.; Min, K.; Kim, C.; Ko, Y. S.; Jeon, J. W.; Seo, H.; Amine Functionalization of Mesoporous Silica on CO2 Adsorption
Park, Y. K.; Choi, M. Epoxide-Functionalization of Polyethyleneimine over a Wide Range of Conditions. Adsorption 2009, 15, 318−328.
for Synthesis of Stable Carbon Dioxide Adsorbent in Temperature (130) Serna-Guerrero, R.; Belmabkhout, Y.; Sayari, A. Triamine-
Swing Adsorption. Nat. Commun. 2016, 7, 12640. Grafted Pore-Expanded Mesoporous Silica for CO2 Capture: Effect of

AZ https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Moisture and Adsorbent Regeneration Strategies. Adsorption 2010, Capability of Diamine-Grafted Metal-Organic Framework Adsorbents
16, 567−575. Through Amine Functionalization. ChemSusChem 2017, 10, 541−
(131) Serna-Guerrero, R.; Da’na, E.; Sayari, A. New Insights into the 550.
Interactions of CO2 with Amine-Functionalized Silica. Ind. Eng. (150) Lee, W. R.; Kim, J. E.; Lee, S. J.; Kang, M.; Kang, D. W.; Lee,
Chem. Res. 2008, 47, 9406−9412. H. Y.; Hiremath, V.; Seo, J. G.; Jin, H.; Moon, D.; Cho, M.; Jung, Y.;
(132) Huang, H. Y.; Yang, R. T.; Chinn, D.; Munson, C. L. Amine- Hong, C. S. Diamine-Functionalization of a Metal-Organic Frame-
Grafted MCM-48 and Silica Xerogel as Superior Sorbents for Acidic work Adsorbent for Superb Carbon Dioxide Adsorption and
Gas Removal from Natural Gas. Ind. Eng. Chem. Res. 2003, 42, 2427− Desorption Properties. ChemSusChem 2018, 11, 1694−1707.
2433. (151) McDonald, T. M.; Mason, J. A.; Kong, X.; Bloch, E. D.; Gygi,
(133) Belmabkhout, Y.; Serna-Guerrero, R.; Sayari, A. Amine- D.; Dani, A.; Crocellà, V.; Giordanino, F.; Odoh, S. O.; Drisdell, W.
Bearing Mesoporous Silica for CO2 Removal from Dry and Humid S.; Vlaisavljevich, B.; Dzubak, A. L.; Poloni, R.; Schnell, S. K.; Planas,
Air. Chem. Eng. Sci. 2010, 65, 3695−3698. N.; Lee, K.; Pascal, T.; Wan, L. F.; Prendergast, D.; Neaton, J. B.;
(134) Foo, G. S.; Lee, J. J.; Chen, C. H.; Hayes, S. E.; Sievers, C.; Smit, B.; Kortright, J. B.; Gagliardi, L.; Bordiga, S.; Reimer, J. A.;
Jones, C. W. Elucidation of Surface Species through in Situ FTIR Long, J. R. Cooperative Insertion of CO2 in Diamine-Appended
Spectroscopy of Carbon Dioxide Adsorption on Amine-Grafted SBA- Metal-Organic Frameworks. Nature 2015, 519, 303−308.
15. ChemSusChem 2017, 10, 266−276. (152) Siegelman, R. L.; McDonald, T. M.; Gonzalez, M. I.; Martell,
(135) Harlick, P. J. E.; Sayari, A. Applications of Pore-Expanded J. D.; Milner, P. J.; Mason, J. A.; Berger, A. H.; Bhown, A. S.; Long, J.
Mesoporous Silicas. 3. Triamine Silane Grafting for Enhanced CO2 R. Controlling Cooperative CO2 Adsorption in Diamine-Appended
Adsorption. Ind. Eng. Chem. Res. 2006, 45, 3248−3255. Mg2 (Dobpdc) Metal-Organic Frameworks. J. Am. Chem. Soc. 2017,
(136) Heydari-Gorji, A.; Belmabkhout, Y.; Sayari, A. Degradation of 139, 10526−10538.
Amine-Supported CO2 Adsorbents in the Presence of Oxygen- (153) Milner, P. J.; Siegelman, R. L.; Forse, A. C.; Gonzalez, M. I.;
Containing Gases. Microporous Mesoporous Mater. 2011, 145, 146− Runčevski, T.; Martell, J. D.; Reimer, J. A.; Long, J. R. A
149. Diaminopropane-Appended Metal-Organic Framework Enabling
(137) Jahandar Lashaki, M.; Sayari, A. CO2 Capture Using Efficient CO2 Capture from Coal Flue Gas via a Mixed Adsorption
Triamine-Grafted SBA-15: The Impact of the Support Pore Structure. Mechanism. J. Am. Chem. Soc. 2017, 139, 13541−13553.
Chem. Eng. J. 2018, 334, 1260−1269. (154) Couck, S.; Denayer, J. F. M.; Baron, G. V.; Remy, T.; Gascon,
(138) Jahandar Lashaki, M.; Ziaei-Azad, H.; Sayari, A. Insights into J.; Kapteijn, F. An Amine-Functionalized MIL-53 Metal-Organic
the Hydrothermal Stability of Triamine-Functionalized SBA-15 Silica Framework with Large Separation Power for CO2 and CH4. J. Am.
for CO2 Adsorption. ChemSusChem 2017, 10, 4037−4045. Chem. Soc. 2009, 131, 6326−6327.
(139) Serna-Guerrero, R.; Belmabkhout, Y.; Sayari, A. Influence of (155) Flaig, R. W.; Osborn Popp, T. M.; Fracaroli, A. M.; Kapustin,
Regeneration Conditions on the Cyclic Performance of Amine- E. A.; Kalmutzki, M. J.; Altamimi, R. M.; Fathieh, F.; Reimer, J. A.;
Grafted Mesoporous Silica for CO2 Capture: An Experimental and Yaghi, O. M. The Chemistry of CO2 Capture in an Amine-
Statistical Study. Chem. Eng. Sci. 2010, 65, 4166−4172. Functionalized Metal-Organic Framework under Dry and Humid
(140) Brunelli, N. A.; Didas, S. A.; Venkatasubbaiah, K.; Jones, C. W. Conditions. J. Am. Chem. Soc. 2017, 139, 12125−12128.
Tuning Cooperativity by Controlling the Linker Length of Silica- (156) Ghalei, B.; Sakurai, K.; Kinoshita, Y.; Wakimoto, K.; Isfahani,
Supported Amines in Catalysis and CO2 Capture. J. Am. Chem. Soc. A. P.; Song, Q.; Doitomi, K.; Furukawa, S.; Hirao, H.; Kusuda, H.;
2012, 134, 13950−13953. Kitagawa, S.; Sivaniah, E. Enhanced Selectivity in Mixed Matrix
(141) Didas, S. A.; Kulkarni, A. R.; Sholl, D. S.; Jones, C. W. Role of Membranes for CO2 Capture through Efficient Dispersion of Amine-
Amine Structure on Carbon Dioxide Adsorption from Ultradilute Gas Functionalized MOF Nanoparticles. Nat. Energy 2017, 2, 17086.
Streams Such as Ambient Air. ChemSusChem 2012, 5, 2058−2064. (157) Hicks, J. C.; Drese, J. H.; Fauth, D. J.; Gray, M. L.; Qi, G.;
(142) Yoo, C. J.; Lee, L. C.; Jones, C. W. Probing Intramolecular Jones, C. W. Designing Adsorbents for CO2 Capture from Flue Gas-
versus Intermolecular CO2 Adsorption on Amine-Grafted SBA-15. Hyperbranched Aminosilicas Capable of Capturing CO2 Reversibly. J.
Langmuir 2015, 31, 13350−13360. Am. Chem. Soc. 2008, 130, 2902−2903.
(143) Sayari, A.; Hamoudi, S. Periodic Mesoporous Silica-Based (158) Chaikittisilp, W.; Didas, S. A.; Kim, H. J.; Jones, C. W. Vapor-
Organic-Inorganic Nanocomposite Materials. Chem. Mater. 2001, 13, Phase Transport as a Novel Route to Hyperbranched Polyamine-
3151−3168. Oxide Hybrid Materials. Chem. Mater. 2013, 25, 613−622.
(144) Su, X.; Bromberg, L.; Martis, V.; Simeon, F.; Huq, A.; Hatton, (159) Kolle, J. M.; Sayari, A. Covalently Immobilized Polyethyle-
T. A. Postsynthetic Functionalization of Mg-MOF-74 with nimine for CO2 Adsorption. Ind. Eng. Chem. Res. 2020, 59, 6944−
Tetraethylenepentamine: Structural Characterization and Enhanced 6950.
CO2 Adsorption. ACS Appl. Mater. Interfaces 2017, 9, 11299−11306. (160) Zhou, Z.; Balijepalli, S. K.; Nguyen-Sorenson, A. H. T.;
(145) Choi, S.; Watanabe, T.; Bae, T. H.; Sholl, D. S.; Jones, C. W. Anderson, C. M.; Park, J. L.; Stowers, K. J. Steam-Stable Covalently
Modification of the Mg/DOBDC MOF with Amines to Enhance Bonded Polyethylenimine Modified Multiwall Carbon Nanotubes for
CO2 Adsorption from Ultradilute Gases. J. Phys. Chem. Lett. 2012, 3, Carbon Dioxide Capture. Energy Fuels 2018, 32, 11701−11709.
1136−1141. (161) Mane, S.; Gao, Z. Y.; Li, Y. X.; Liu, X. Q.; Sun, L. B. Rational
(146) Cao, Y.; Song, F.; Zhao, Y.; Zhong, Q. Capture of Carbon Fabrication of Polyethylenimine-Linked Microbeads for Selective
Dioxide from Flue Gas on TEPA-Grafted Metal-Organic Framework CO2 Capture. Ind. Eng. Chem. Res. 2018, 57, 250−258.
Mg2(Dobdc). J. Environ. Sci. 2013, 25, 2081−2087. (162) Franchi, R. S.; Harlick, P. J. E.; Sayari, A. Applications of Pore-
(147) Liao, P. Q.; Chen, X. W.; Liu, S. Y.; Li, X. Y.; Xu, Y. T.; Tang, Expanded Mesoporous Silica. 2. Development of a High-Capacity,
M.; Rui, Z.; Ji, H.; Zhang, J. P.; Chen, X. M. Putting an Ultrahigh Water-Tolerant Adsorbent for CO2. Ind. Eng. Chem. Res. 2005, 44,
Concentration of Amine Groups into a Metal-Organic Framework for 8007−8013.
CO2 Capture at Low Pressures. Chem. Sci. 2016, 7, 6528−6533. (163) Lee, J. J.; Chen, C. H.; Shimon, D.; Hayes, S. E.; Sievers, C.;
(148) Lee, W. R.; Hwang, S. Y.; Ryu, D. W.; Lim, K. S.; Han, S. S.; Jones, C. W. Effect of Humidity on the CO2 Adsorption of Tertiary
Moon, D.; Choi, J.; Hong, C. S. Diamine-Functionalized Metal- Amine Grafted SBA-15. J. Phys. Chem. C 2017, 121, 23480−23487.
Organic Framework: Exceptionally High CO2 Capacities from (164) Ko, Y. G.; Shin, S. S.; Choi, U. S. Primary, Secondary, and
Ambient Air and Flue Gas, Ultrafast CO2 Uptake Rate, and Tertiary Amines for CO2 Capture: Designing for Mesoporous CO2
Adsorption Mechanism. Energy Environ. Sci. 2014, 7, 744−751. Adsorbents. J. Colloid Interface Sci. 2011, 361, 594−602.
(149) Jo, H.; Lee, W. R.; Kim, N. W.; Jung, H.; Lim, K. S.; Kim, J. E.; (165) Sayari, A.; Heydari-Gorji, A.; Yang, Y. CO2-Induced
Kang, D. W.; Lee, H.; Hiremath, V.; Seo, J. G.; Jin, H.; Moon, D.; Degradation of Amine-Containing Adsorbents: Reaction Products
Han, S. S.; Hong, C. S. Fine-Tuning of the Carbon Dioxide Capture and Pathways. J. Am. Chem. Soc. 2012, 134, 13834−13842.

BA https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

(166) Azmi, N.; Yusup, S.; Sabil, K. M. Effect of Water onto Porous (187) Wang, S.; Li, X.; Wu, H.; Tian, Z.; Xin, Q.; He, G.; Peng, D.;
CaO for CO2 Adsorption: Experimental and Extended Isotherm Chen, S.; Yin, Y.; Jiang, Z.; Guiver, M. D. Advances in High
Model. J. Cleaner Prod. 2017, 168, 973−982. Permeability Polymer-Based Membrane Materials for CO2 Separa-
(167) Ding, Y. D.; Song, G.; Liao, Q.; Zhu, X.; Chen, R. Bench Scale tions. Energy Environ. Sci. 2016, 9, 1863−1890.
Study of CO2 Adsorption Performance of MgO in the Presence of (188) Tong, Z.; Vakharia, V. K.; Gasda, M.; Ho, W. S. W. Water
Water Vapor. Energy 2016, 112, 101−110. Vapor and CO2 Transport through Amine-Containing Facilitated
(168) Gankanda, A.; Cwiertny, D. M.; Grassian, V. H. Role of Transport Membranes. React. Funct. Polym. 2015, 86, 111−116.
Atmospheric CO2 and H2O Adsorption on ZnO and CuO (189) Yang, Q.; Lin, Q.; Sammarchi, S.; Li, J.; Li, S.; Wang, D. Water
Nanoparticle Aging: Formation of New Surface Phases and the Vapor Effects on CO2 Separation of Amine-Containing Facilitated
Impact on Nanoparticle Dissolution. J. Phys. Chem. C 2016, 120, Transport Membranes (AFTMs) Module: Mathematical Modeling
19195−19203. Using Tanks-in-Series Approach. Greenhouse Gases: Sci. Technol. 2021,
(169) Lee, S. C.; Choi, B. Y.; Lee, T. J.; Ryu, C. K.; Ahn, Y. S.; Kim, 11, 52−68.
J. C. CO2 Absorption and Regeneration of Alkali Metal-Based Solid (190) Wong, K. C.; Goh, P. S.; Ismail, A. F. Thin Film
Sorbents. Catal. Today 2006, 111, 385−390. Nanocomposite: The next Generation Selective Membrane for CO2
(170) Nanayakkara, C. E.; Larish, W. A.; Grassian, V. H. Titanium Removal. J. Mater. Chem. A 2016, 4, 15726−15748.
Dioxide Nanoparticle Surface Reactivity with Atmospheric Gases, (191) Yue, M. B.; Sun, L. B.; Cao, Y.; Wang, Z. J.; Wang, Y.; Yu, Q.;
CO2, SO2, and NO2: Roles of Surface Hydroxyl Groups and Zhu, J. H. Promoting the CO2 Adsorption in the Amine-Containing
Adsorbed Water in the Formation and Stability of Adsorbed Products. SBA-15 by Hydroxyl Group. Microporous Mesoporous Mater. 2008,
J. Phys. Chem. C 2014, 118, 23011−23021. 114, 74−81.
(171) Wang, J.; Wang, S.; Xin, Q.; Li, Y. Perspectives on Water- (192) Knowles, G. P.; Graham, J. V.; Delaney, S. W.; Chaffee, A. L.
Facilitated CO2 Capture Materials. J. Mater. Chem. A 2017, 5, 6794− Aminopropyl-Functionalized Mesoporous Silicas as CO2 Adsorbents.
6816. Fuel Process. Technol. 2005, 86, 1435−1448.
(172) González-Zamora, E.; Ibarra, I. A. CO2 Capture Under (193) Knowles, G. P.; Delaney, S. W.; Chaffee, A. L. Amine-
Humid Conditions in Metal-Organic Frameworks. Mater. Chem. Functionalised Mesoporous Silicas as CO2 Adsorbents. Stud. Surf. Sci.
Front. 2017, 1, 1471−1484. Catal. 2005, 156, 887−896.
(173) Li, P.; Chen, J.; Zhang, J.; Wang, X. Water Stability and (194) Knowles, G. P.; Delaney, S. W.; Chaffee, A. L.
Competition Effects toward CO2 Adsorption on Metal Organic Diethylenetriamine[Propyl(Silyl)]-Functionalized (DT) Mesoporous
Frameworks. Sep. Purif. Rev. 2015, 44, 19−27. Silicas as CO2 Adsorbents. Ind. Eng. Chem. Res. 2006, 45, 2626−2633.
(174) Han, Y.; Ho, W. S. W. Recent Advances in Polymeric (195) Chanut, N.; Bourrelly, S.; Kuchta, B.; Serre, C.; Chang, J. S.;
Membranes for CO2 Capture. Chin. J. Chem. Eng. 2018, 26, 2238− Wright, P. A.; Llewellyn, P. L. Screening the Effect of Water Vapour
2254. on Gas Adsorption Performance: Application to CO2 Capture from
(175) Zhao, Y.; Winston Ho, W. S. Steric Hindrance Effect on Flue Gas in Metal-Organic Frameworks. ChemSusChem 2017, 10,
Amine Demonstrated in Solid Polymer Membranes for CO2 1543−1553.
Transport. J. Membr. Sci. 2012, 415−416, 132−138. (196) Peralta, R. A.; Alcántar-Vázquez, B.; Sánchez-Serratos, M.;
(176) Venna, S. R.; Carreon, M. A. Metal Organic Framework González-Zamora, E.; Ibarra, I. A. Carbon Dioxide Capture in the
Membranes for Carbon Dioxide Separation. Chem. Eng. Sci. 2015, Presence of Water Vapour in InOF-1. Inorg. Chem. Front. 2015, 2,
124, 3−19. 898−903.
(177) Pera-Titus, M. Porous Inorganic Membranes for CO2 (197) Monazam, E. R.; Breault, R. W.; Fauth, D. J.; Shadle, L. J.;
Capture: Present and Prospects. Chem. Rev. 2014, 114, 1413−1492. Bayham, S. Insights into the Adsorption of Carbon Dioxide in the
(178) Shah, M.; McCarthy, M. C.; Sachdeva, S.; Lee, A. K.; Jeong, Presence of Water Vapor Utilizing a Low Molecular Weight
H. K. Current Status of Metal-Organic Framework Membranes for Polyethylenimine-Impregnated CARiACT Silica Sorbent. Ind. Eng.
Gas Separations: Promises and Challenges. Ind. Eng. Chem. Res. 2012, Chem. Res. 2017, 56, 9054−9064.
51, 2179−2199. (198) Sánchez-González, E.; Á lvarez, J. R.; Peralta, R. A.; Campos-
(179) Yeo, Z. Y.; Chew, T. L.; Zhu, P. W.; Mohamed, A. R.; Chai, S.- Reales-Pineda, A.; Tejeda-Cruz, A.; Lima, E.; Balmaseda, J.; González-
P. Synthesis and Performance of Microporous Inorganic Membranes Zamora, E.; Ibarra, I. A. Water Adsorption Properties of NOTT-401
for CO2 Separation: A Review. J. Porous Mater. 2013, 20, 1457−1475. and CO2 Capture under Humid Conditions. ACS Omega 2016, 1,
(180) Zito, P. F.; Brunetti, A.; Barbieri, G. Selective Mass Transport 305−310.
of CO2 Containing Mixtures through Zeolite Membranes. J. Membr. (199) Ebner, A. D.; Gray, M. L.; Chisholm, N. G.; Black, Q. T.;
Sci. Res. 2020, 6, 333−343. Mumford, D. D.; Nicholson, M. A.; Ritter, J. A. Suitability of a Solid
(181) Lei, L.; Bai, L.; Lindbråthen, A.; Pan, F.; Zhang, X.; He, X. Amine Sorbent for CO2 Capture by Pressure Swing Adsorption. Ind.
Carbon Membranes for CO2 Removal: Status and Perspectives from Eng. Chem. Res. 2011, 50, 5634−5641.
Materials to Processes. Chem. Eng. J. 2020, 401, 126084. (200) Liu, S. H.; Wu, C. H.; Lee, H. K.; Liu, S.-B. Highly Stable
(182) Janakiram, S.; Ansaloni, L.; Jin, S.-A.; Yu, X.; Dai, Z.; Spontak, Amine-Modified Mesoporous Silica Materials for Efficient CO2
R. J.; Deng, L. Humidity-Responsive Molecular Gate-Opening Capture. Top. Catal. 2010, 53, 210−217.
Mechanism for Gas Separation in Ultraselective Nanocellulose/IL (201) Chen, C.; Yang, S. T.; Ahn, W. S.; Ryoo, R. Amine-
Hybrid Membranes. Green Chem. 2020, 22, 3546−3557. Impregnated Silica Monolith with a Hierarchical Pore Structure:
(183) Klemm, A.; Lee, Y.-Y.; Mao, H.; Gurkan, B. Facilitated Enhancement of CO2 Capture Capacity. Chem. Commun. 2009,
Transport Membranes With Ionic Liquids for CO2 Separations. Front. 3627−3629.
Chem. 2020, 8, 637. (202) Subagyono, D. J. N.; Marshall, M.; Knowles, G. P.; Chaffee, A.
(184) Fu, Y.; Jiang, Y.-B.; Dunphy, D.; Xiong, H.; Coker, E.; Chou, L. CO2 Adsorption by Amine Modified Siliceous Mesostructured
S. S.; Zhang, H.; Vanegas, J. M.; Croissant, J. G.; Cecchi, J. L.; Rempe, Cellular Foam (MCF) in Humidified Gas. Microporous Mesoporous
S. B.; Brinker, C. J. Ultra-Thin Enzymatic Liquid Membrane for CO2 Mater. 2014, 186, 84−93.
Separation and Capture. Nat. Commun. 2018, 9, 990. (203) Smal, I. M.; Yu, Q.; Veneman, R.; Fränzel-Luiten, B.; Brilman,
(185) Saeed, M.; Deng, L. CO2 Facilitated Transport Membrane D. W. F. TG-FTIR Measurement of CO2-H2O Co-Adsorption for
Promoted by Mimic Enzyme. J. Membr. Sci. 2015, 494, 196−204. CO2 Air Capture Sorbent Screening. Energy Procedia 2014, 63,
(186) Liu, Y.; Wang, Z.; Shi, M.; Li, N.; Zhao, S.; Wang, J. Carbonic 6834−6841.
Anhydrase Inspired Poly(N-Vinylimidazole)/Zeolite Zn-β Hybrid (204) Pirngruber, G. D.; Hamon, L.; Bourrelly, S.; Llewellyn, P. L.;
Membranes for CO2 Capture. Chem. Commun. 2018, 54, 7239−7242. Lenoir, E.; Guillerm, V.; Serre, C.; Devic, T. A Method for Screening

BB https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

the Potential of MOFs as CO2 Adsorbents in Pressure Swing (222) Xu, D.; Zhang, J.; Li, G.; Xiao, P.; Webley, P.; Zhai, Y. Effect
Adsorption Processes. ChemSusChem 2012, 5, 762−776. of Water Vapor from Power Station Flue Gas on CO2 Capture by
(205) Mason, J. A.; McDonald, T. M.; Bae, T. H.; Bachman, J. E.; Vacuum Swing Adsorption with Activated Carbon. J. Fuel Chem.
Sumida, K.; Dutton, J. J.; Kaye, S. S.; Long, J. R. Application of a Technol. 2011, 39, 169−174.
High-Throughput Analyzer in Evaluating Solid Adsorbents for Post- (223) Younas, M.; Leong, L. K.; Mohamed, A. R.; Sethupathi, S.
Combustion Carbon Capture via Multicomponent Adsorption of CO2 Adsorption by Modified Palm Shell Activated Carbon (PSAC)
CO2, N2, and H2O. J. Am. Chem. Soc. 2015, 137, 4787−4803. Via Chemical and Physical Activation and Metal Impregnation. Chem.
(206) Alvarado-Alvarado, D.; González-Estefan, J. H.; Flores, J. G.; Eng. Commun. 2016, 203, 1455−1463.
Á lvarez, J. R.; Aguilar-Pliego, J.; Islas-Jácome, A.; Chastanet, G.; (224) Zerze, H.; Tipirneni, A.; McHugh, A. J. Reusable Poly-
González-Zamora, E.; Lara-García, H. A.; Alcántar-Vázquez, B.; (Allylamine)-Based Solid Materials for Carbon Dioxide Capture
Gonidec, M.; Ibarra, I. A. Water Adsorption Properties of Fe(Pz)- under Continuous Flow of Ambient Air. Sep. Sci. Technol. 2017, 52,
[Pt(CN)4] and the Capture of CO2 and CO. Organometallics 2020, 2513−2522.
39, 949−955. (225) Zhao, P.; Zhang, G.; Sun, Y.; Xu, Y. CO2 Adsorption Behavior
(207) Belmabkhout, Y.; Serna-Guerrero, R.; Sayari, A. Adsorption of and Kinetics on Amine-Functionalized Composites Silica with
CO2-Containing Gas Mixtures over Amine-Bearing Pore-Expanded Trimodal Nanoporous Structure. Energy Fuels 2017, 31, 12508−
MCM-41 Silica: Application for Gas Purification. Ind. Eng. Chem. Res. 12520.
2010, 49, 359−365. (226) Brilman, D. W. F.; Veneman, R. Capturing Atmospheric CO2
(208) Durán, I.; Rubiera, F.; Pevida, C. Separation of CO2 in a Solid Using Supported Amine Sorbents. Energy Procedia 2013, 37, 6070−
Waste Management Incineration Facility Using Activated Carbon 6078.
Derived from Pine Sawdust. Energies 2017, 10, 827. (227) Veneman, R.; Frigka, N.; Zhao, W.; Li, Z.; Kersten, S.;
(209) Plaza, M. G.; González, A. S.; Pevida, C.; Rubiera, F. Influence Brilman, W. Adsorption of H2O and CO2 on Supported Amine
of Water Vapor on CO2 Adsorption Using a Biomass-Based Carbon. Sorbents. Int. J. Greenhouse Gas Control 2015, 41, 268−275.
Ind. Eng. Chem. Res. 2014, 53, 15488−15499. (228) Alesi, W. R.; Gray, M.; Kitchin, J. R. CO2 Adsorption on
(210) Chiang, Y. C.; Chen, Y. J.; Wu, C. Y. Effect of Relative Supported Molecular Amidine Systems on Activated Carbon.
Humidity on Adsorption Breakthrough of CO2 on Activated Carbon ChemSusChem 2010, 3, 948−956.
Fibers. Materials 2017, 10, 1296. (229) Drese, J. H.; Choi, S.; Lively, R. P.; Koros, W. J.; Fauth, D. J.;
(211) Liu, Y.; Ye, Q.; Shen, M.; Shi, J.; Chen, J.; Pan, H.; Shi, Y. Gray, M. L.; Jones, C. W. Synthesis-Structure-Property Relationships
Carbon Dioxide Capture by Functionalized Solid Amine Sorbents for Hyperbranched Aminosilica CO2 Adsorbents. Adv. Funct. Mater.
with Simulated Flue Gas Conditions. Environ. Sci. Technol. 2011, 45, 2009, 19, 3821−3832.
5710−5716. (230) Bansal, R. C.; Goyal, M. Activated Carbon Adsorption; CRC
(212) Wang, J.; Huang, H.; Wang, M.; Yao, L.; Qiao, W.; Long, D.; Press: Boca Raton, FL, 2005.
Ling, L. Direct Capture of Low-Concentration CO2 on Mesoporous (231) Activated Carbon: Synthesis, Properties and Uses; Chemistry
Carbon-Supported Solid Amine Adsorbents at Ambient Temperature. Research and Applications Series; Hsu, M., Davies, E., Eds.; Nova
Ind. Eng. Chem. Res. 2015, 54, 5319−5327. Science Publishers, 2017.
(213) Wang, J.; Wang, M.; Li, W.; Qiao, W.; Long, D.; Ling, L. (232) Shafeeyan, M. S.; Daud, W. M. A. W.; Houshmand, A.;
Application of Polyethylenimine-Impregnated Solid Adsorbents for Shamiri, A. A Review on Surface Modification of Activated Carbon for
Direct Capture of Low-Concentration CO2. AIChE J. 2015, 61, 972− Carbon Dioxide Adsorption. J. Anal. Appl. Pyrolysis 2010, 89, 143−
980. 151.
(214) Hiyoshi, N.; Yogo, K.; Yashima, T. Adsorption Characteristics (233) Chen, E.; Rolin, K.; Stillman, Z. Activated Carbon-Based
of Carbon Dioxide on Organically Functionalized SBA-15. Micro- Carbon Dioxide Adsorption Process. Scholarly Commons; University
porous Mesoporous Mater. 2005, 84, 357−365. of Pennsylvania, 2017; https://repository.upenn.edu/cbe_sdr/89/.
(215) Jadhav, P. D.; Chatti, R. V.; Biniwale, R. B.; Labhsetwar, N. K.; (234) Balsamo, M.; Silvestre-Albero, A.; Silvestre-Albero, J.; Erto, A.;
Devotta, S.; Rayalu, S. S. Monoethanol Amine Modified Zeolite 13X Rodríguez-Reinoso, F.; Lancia, A. Assessment of CO2 Adsorption
for CO2 Adsorption at Different Temperatures. Energy Fuels 2007, Capacity on Activated Carbons by a Combination of Batch and
21, 3555−3559. Dynamic Tests. Langmuir 2014, 30, 5840−5848.
(216) Soubeyrand-Lenoir, E.; Vagner, C.; Yoon, J. W.; Bazin, P.; (235) Boyjoo, Y.; Cheng, Y.; Zhong, H.; Tian, H.; Pan, J.; Pareek, V.
Ragon, F.; Hwang, Y. K.; Serre, C.; Chang, J. S.; Llewellyn, P. L. How K.; Jiang, S. P.; Lamonier, J. F.; Jaroniec, M.; Liu, J. From Waste Coca
Water Fosters a Remarkable 5-Fold Increase in Low-Pressure CO2 Cola® to Activated Carbons with Impressive Capabilities for CO2
Uptake Within Mesoporous MIL-100(Fe). J. Am. Chem. Soc. 2012, Adsorption and Supercapacitors. Carbon 2017, 116, 490−499.
134, 10174−10181. (236) Singh, G.; Lakhi, K. S.; Sathish, C. I.; Ramadass, K.; Yang, J.
(217) Sayari, A.; Liu, Q.; Mishra, P. Enhanced Adsorption Efficiency H.; Vinu, A. Oxygen-Functionalized Mesoporous Activated Carbons
Through Materials Design for Direct Air Capture Over Supported Derived from Casein and Their Superior CO2 Adsorption Capacity at
Polyethylenimine. ChemSusChem 2016, 9, 2796−2803. Both Low- And High-Pressure Regimes. ACS Appl. Nano Mater. 2019,
(218) Serna-Guerrero, R.; Belmabkhout, Y.; Sayari, A. Further 2, 1604−1613.
Investigations of CO2 Capture Using Triamine-Grafted Pore- (237) Zhao, Y.; Yu, B.; Yang, Z.; Zhang, H.; Hao, L.; Gao, X.; Liu, Z.
Expanded Mesoporous Silica. Chem. Eng. J. 2010, 158, 513−519. A Protic Ionic Liquid Catalyzes CO2 Conversion at Atmospheric
(219) Gray, M. L.; Hoffman, J. S.; Hreha, D. C.; Fauth, D. J.; Pressure and Room Temperature: Synthesis of Quinazoline-2,4-
Hedges, S. W.; Champagne, K. J.; Pennline, H. W. Parametric Study (1H,3H)-Diones. Angew. Chem., Int. Ed. 2014, 53, 5922−5925.
of Solid Amine Sorbents for the Capture of Carbon Dioxide. Energy (238) Shafeeyan, M. S.; Daud, W. M. A. W.; Houshmand, A.; Arami-
Fuels 2009, 23, 4840−4844. Niya, A. Ammonia Modification of Activated Carbon to Enhance
(220) Sehaqui, H.; Gálvez, M. E.; Becatinni, V.; Cheng Ng, Y.; Carbon Dioxide Adsorption: Effect of Pre-Oxidation. Appl. Surf. Sci.
Steinfeld, A.; Zimmermann, T.; Tingaut, P. Fast and Reversible Direct 2011, 257, 3936−3942.
CO2 Capture From Air onto All-Polymer Nanofibrillated Cellulose- (239) Ngoy, J. M.; Wagner, N.; Riboldi, L.; Bolland, O. A CO2
Polyethylenimine Foams. Environ. Sci. Technol. 2015, 49, 3167−3174. Capture Technology Using Multi-Walled Carbon Nanotubes with
(221) Gebald, C.; Wurzbacher, J. A.; Borgschulte, A.; Zimmermann, Polyaspartamide Surfactant. Energy Procedia 2014, 63, 2230−2248.
T.; Steinfeld, A. Single-Component and Binary CO2 and H2O (240) Wahby, A.; Silvestre-Albero, J.; Sepúlveda-Escribano, A.;
Adsorption of Amine-Functionalized Cellulose. Environ. Sci. Technol. Rodríguez-Reinoso, F. CO2 Adsorption on Carbon Molecular Sieves.
2014, 48, 2497−2504. Microporous Mesoporous Mater. 2012, 164, 280−287.

BC https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

(241) Ghosh, A.; Subrahmanyam, K. S.; Krishna, K. S.; Datta, S.; (263) Č ejka, J.; Morris, R. E.; Nachtigall, P. Zeolites in Catalysis:
Govindaraj, A.; Pati, S. K.; Rao, C. N. R. Uptake of H2 and CO2 by Properties and Applications; Royal Society of Chemistry, 2017.
Graphene. J. Phys. Chem. C 2008, 112, 15704−15707. (264) Aguado, S.; Bergeret, G.; Daniel, C.; Farrusseng, D. Absolute
(242) Wickramaratne, N. P.; Jaroniec, M. Importance of Small Molecular Sieve Separation of Ethylene/Ethane Mixtures with Silver
Micropores in CO2 Capture by Phenolic Resin-Based Activated Zeolite A. J. Am. Chem. Soc. 2012, 134, 14635−14637.
Carbon Spheres. J. Mater. Chem. A 2013, 1, 112−116. (265) Lin, C. C. H.; Sawada, J. A.; Wu, L.; Haastrup, T.; Kuznicki, S.
(243) Nandi, M.; Okada, K.; Dutta, A.; Bhaumik, A.; Maruyama, J.; M. Anion-Controlled Pore Size of Titanium Silicate Molecular Sieves.
Derks, D.; Uyama, H. Unprecedented CO2 Uptake over Highly J. Am. Chem. Soc. 2009, 131, 609−614.
Porous N-Doped Activated Carbon Monoliths Prepared by Physical (266) Song, Z.; Huang, Y.; Xu, W. L.; Wang, L.; Bao, Y.; Li, S.; Yu,
Activation. Chem. Commun. 2012, 48, 10283−10285. M. Continuously Adjustable, Molecular-Sieving “Gate” on 5A Zeolite
(244) Chen, J.; Yang, J.; Hu, G.; Hu, X.; Li, Z.; Shen, S.; Radosz, M.; for Distinguishing Small Organic Molecules by Size. Sci. Rep. 2015, 5,
Fan, M. Enhanced CO2 Capture Capacity of Nitrogen-Doped 13981.
Biomass-Derived Porous Carbons. ACS Sustainable Chem. Eng. (267) Siriwardane, R. V.; Shen, M. S.; Fisher, E. P.; Losch, J.
2016, 4, 1439−1445. Adsorption of CO2 on Zeolites at Moderate Temperatures. Energy
(245) Keller, J. U.; Staudt, R. Adsorption Isotherms. In Gas Fuels 2005, 19, 1153−1159.
Adsorption Equilibria: Experimental Methods and Adsorption Isotherms; (268) Siriwardane, R. V.; Shen, M. S.; Fisher, E. P.; Poston, J. A.
Springer, 2005; Chapter 7, pp 359−413. Adsorption of CO2 on Molecular Sieves and Activated Carbon.
(246) Cossarutto, L.; Zimny, T.; Kaczmarczyk, J.; Siemieniewska, T.; Energy Fuels 2001, 15, 279−284.
Bimer, J.; Weber, J. V. Transport and Sorption of Water Vapour in (269) Cavenati, S.; Grande, C. A.; Rodrigues, A. E. Adsorption
Activated Carbons. Carbon 2001, 39, 2339−2346. Equilibrium of Methane, Carbon Dioxide, and Nitrogen on Zeolite
(247) Do, D. D.; Do, H. D. A Model for Water Adsorption in 13X at High Pressures. J. Chem. Eng. Data 2004, 49, 1095−1101.
Activated Carbon. Carbon 2000, 38, 767−773. (270) Harlick, P. J. E.; Tezel, F. H. An Experimental Adsorbent
(248) Dubinin, M. M. Water Vapor Adsorption and the Micro- Screening Study for CO2 Removal From N2. Microporous Mesoporous
porous Structures of Carbonaceous Adsorbents. Carbon 1980, 18, Mater. 2004, 76, 71−79.
355−364. (271) Maurin, G.; Llewellyn, P. L.; Bell, R. G. Adsorption
(249) Barton, S. S.; Evans, M. J. B.; MacDonald, J. A. F. The Mechanism of Carbon Dioxide in Faujasites: Grand Canonical
Adsorption of Water Vapor by Porous Carbon. Carbon 1991, 29, Monte Carlo Simulations and Microcalorimetry Measurements. J.
1099−1105. Phys. Chem. B 2005, 109, 16084−16091.
(250) Müller, E. A.; Rull, L. F.; Vega, L. F.; Gubbins, K. E. (272) Kumar, A.; Madden, D. G.; Lusi, M.; Chen, K. J.; Daniels, E.
Adsorption of Water on Activated Carbons: A Molecular Simulation A.; Curtin, T.; Perry, J. J.; Zaworotko, M. J. Direct Air Capture of
Study. J. Phys. Chem. 1996, 100, 1189−1196. CO2 by Physisorbent Materials. Angew. Chem., Int. Ed. 2015, 54,
(251) Koresh, J.; Soffer, A. Study of Molecular Sieve Carbons. Part 14372−14377.
1. Pore Structure, Gradual Pore Opening and Mechanism of (273) Stuckert, N. R.; Yang, R. T. CO2 Capture from the
Molecular Sieving. J. Chem. Soc., Faraday Trans. 1 1980, 76, 2457− Atmosphere and Simultaneous Concentration Using Zeolites and
2471. Amine-Grafted SBA-15. Environ. Sci. Technol. 2011, 45, 10257−
(252) Koresh, J. E.; Kim, T. H.; Koros, W. J. Study of 10264.
Ultramicroporous Carbons by High-Pressure Sorption. J. Chem. Soc., (274) Li, G.; Xiao, P.; Webley, P.; Zhang, J.; Singh, R.; Marshall, M.
Faraday Trans. 1 1989, 85, 1537−1544. Capture of CO2 from High Humidity Flue Gas by Vacuum Swing
(253) Liu, L.; Jin, S.; Park, Y.; Park, Y. C.; Lee, C. H. Sorption Adsorption with Zeolite 13X. Adsorption 2008, 14, 415−422.
Equilibria and Kinetics of CO2, N2, and H2O on KOH-Treated (275) Chen, H.; Wang, W.; Ding, J.; Wei, X.; Lu, J. CO2 Adsorption
Activated Carbon. Ind. Eng. Chem. Res. 2018, 57, 17218−17225. Capacity of FAU Zeolites in Presence of H2O: A Monte Carlo
(254) Plaza, M. G.; Durán, I.; Querejeta, N.; Rubiera, F.; Pevida, C. Simulation Study. Energy Procedia 2017, 105, 4370−4376.
Experimental and Simulation Study of Adsorption in Postcombustion (276) Mazaj, M.; Bjelica, M.; Ž agar, E.; Logar, N. Z.; Kovačič, S.
Conditions Using a Microporous Biochar. 2. H2O, CO2, and N2 Zeolite Nanocrystals Embedded in Microcellular Carbon Foam as a
Adsorption. Ind. Eng. Chem. Res. 2016, 55, 6854−6865. High-Performance CO2 Capture Adsorbent with Energy-Saving
(255) Plaza, M. G.; González, A. S.; Rubiera, F.; Pevida, C. Regeneration Properties. ChemSusChem 2020, 13, 2089−2097.
Evaluation of Microporous Biochars Produced by Single-Step (277) Song, M. K.; Datta, S. J.; Yoon, K. B. DFT Calculated
Oxidation for Postcombustion CO2 Capture under Humid Structures and Electronic Properties of ETS-10, AM-6, and SGU-29:
Conditions. Energy Procedia 2014, 63, 693−702. Structure Stabilization through Periodic Distortions. J. Phys. Chem. C
(256) Plaza, M. G.; González, A. S.; Rubiera, F.; Pevida, C. Water 2016, 120, 20206−20215.
Vapour Adsorption by a Coffee-Based Microporous Carbon: Effect on (278) Datta, S. J.; Khumnoon, C.; Lee, Z. H.; Moon, W. K.; Docao,
CO2 Capture. J. Chem. Technol. Biotechnol. 2015, 90, 1592−1600. S.; Nguyen, T. H.; Hwang, I. C.; Moon, D.; Oleynikov, P.; Terasaki,
(257) Wang, Y.; Zhou, Y.; Liu, C.; Zhou, L. Comparative Studies of O.; Yoon, K. B. CO2 Capture from Humid Flue Gases and Humid
CO2 and CH4 Sorption on Activated Carbon in Presence of Water. Atmosphere Using a Microporous Coppersilicate. Science 2015, 350,
Colloids Surf., A 2008, 322, 14−18. 302−306.
(258) Jalilov, A. S.; Li, Y.; Kittrell, C.; Tour, J. M. Increased CO2 (279) Kim, C.; Cho, H. S.; Chang, S.; Cho, S. J.; Choi, M. An
Selectivity of Asphalt-Derived Porous Carbon through Introduction of Ethylenediamine-Grafted Y Zeolite: A Highly Regenerable Carbon
Water into Pore Space. Nat. Energy 2017, 2, 932−938. Dioxide Adsorbent: Via Temperature Swing Adsorption without Urea
(259) Psarras, P.; He, J.; Wilcox, J. Effect of Water on the CO2 Formation. Energy Environ. Sci. 2016, 9, 1803−1811.
Adsorption Capacity of Amine-Functionalized Carbon Sorbents. Ind. (280) Jeong, W.; Kim, J. Understanding the Mechanisms of CO2
Eng. Chem. Res. 2017, 56, 6317−6325. Adsorption Enhancement in Pure Silica Zeolites under Humid
(260) Psarras, P.; He, J.; Wilcox, J. Molecular Simulations of Conditions. J. Phys. Chem. C 2016, 120, 23500−23510.
Nitrogen-Doped Hierarchical Carbon Adsorbents for Post-Combus- (281) Hiyoshi, N.; Yogo, K.; Yashima, T. Adsorption of Carbon
tion CO2 Capture. Phys. Chem. Chem. Phys. 2016, 18, 28747−28758. Dioxide on Amine Modified SBA-15 in the Presence of Water Vapor.
(261) You, Y. Y.; Liu, X. J. Modeling of CO2 Adsorption and Chem. Lett. 2004, 33, 510−511.
Recovery from Wet Flue Gas by Using Activated Carbon. Chem. Eng. (282) Ren, J.; Dyosiba, X.; Musyoka, N. M.; Langmi, H. W.; Mathe,
J. 2019, 369, 672−685. M.; Liao, S. Review on the Current Practices and Efforts Towards
(262) Jacobs, P. A.; Flanigen, E. M.; Jansen, J. C.; van Bekkum, H. Pilot-Scale Production of Metal-Organic Frameworks (MOFs). Coord.
Introduction to Zeolite Science and Practice, 2nd ed.; Elsevier, 2001. Chem. Rev. 2017, 352, 187−219.

BD https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

(283) Tan, Y. X.; Wang, F.; Zhang, J. Design and Synthesis of (302) Farrusseng, D. Metal-Organic Frameworks: Applications from
Multifunctional Metal-Organic Zeolites. Chem. Soc. Rev. 2018, 47, Catalysis to Gas Storage; Farrusseng, D., Ed.; Wiley and Sons, 2011.
2130−2144. (303) Kang, Z.; Fan, L.; Sun, D. Recent Advances and Challenges of
(284) Kirchon, A.; Feng, L.; Drake, H. F.; Joseph, E. A.; Zhou, H. C. Metal-Organic Framework Membranes For Gas Separation. J. Mater.
From Fundamentals to Applications: A Toolbox for Robust and Chem. A 2017, 5, 10073−10091.
Multifunctional MOF Materials. Chem. Soc. Rev. 2018, 47, 8611− (304) Baumann, A. E.; Burns, D. A.; Liu, B.; Thoi, V. S. Metal-
8638. Organic Framework Functionalization and Design Strategies for
(285) Sumida, K.; Liang, K.; Reboul, J.; Ibarra, I. A.; Furukawa, S.; Advanced Electrochemical Energy Storage Devices. Commun. Chem.
Falcaro, P. Sol-Gel Processing of Metal-Organic Frameworks. Chem. 2019, 2, 86.
Mater. 2017, 29, 2626−2645. (305) Pascanu, V.; González Miera, G.; Inge, A. K.; Martín-Matute,
(286) Masoomi, M. Y.; Morsali, A.; Dhakshinamoorthy, A.; Garcia, B. Metal-Organic Frameworks as Catalysts for Organic Synthesis: A
H. Mixed-Metal MOFs: Unique Opportunities in Metal-Organic Critical Perspective. J. Am. Chem. Soc. 2019, 141, 7223−7234.
Framework (MOF) Functionality and Design. Angew. Chem., Int. Ed. (306) Yang, D.; Gates, B. C. Catalysis by Metal Organic
2019, 58, 15188−15205. Frameworks: Perspective and Suggestions for Future Research. ACS
(287) Kalmutzki, M. J.; Hanikel, N.; Yaghi, O. M. Secondary Catal. 2019, 9, 1779−1798.
Building Units as the Turning Point in the Development of the (307) Zhu, H.; Liu, D. The Synthetic Strategies of Metal-Organic
Reticular Chemistry of MOFs. Sci. Adv. 2018, 4, No. eaat9180. Framework Membranes, Films and 2D MOFs and Their Applications
(288) Farha, O. K.; Eryazici, I.; Jeong, N. C.; Hauser, B. G.; Wilmer, in Devices. J. Mater. Chem. A 2019, 7, 21004−21035.
C. E.; Sarjeant, A. A.; Snurr, R. Q.; Nguyen, S. T.; Yazaydin, A. Ö .; (308) Kalmutzki, M. J.; Diercks, C. S.; Yaghi, O. M. Metal-Organic
Hupp, J. T. Metal-Organic Framework Materials with Ultrahigh Frameworks for Water Harvesting from Air. Adv. Mater. 2018, 30,
Surface Areas: Is the Sky the Limit? J. Am. Chem. Soc. 2012, 134, 1704304.
15016−15021. (309) Meng, J.; Liu, X.; Niu, C.; Pang, Q.; Li, J.; Liu, F.; Liu, Z.; Mai,
(289) Hönicke, I. M.; Senkovska, I.; Bon, V.; Baburin, I. A.; Bönisch, L. Advances in Metal-Organic Framework Coatings: Versatile
N.; Raschke, S.; Evans, J. D.; Kaskel, S. Balancing Mechanical Stability Synthesis and Broad Applications. Chem. Soc. Rev. 2020, 49, 3142−
and Ultrahigh Porosity in Crystalline Framework Materials. Angew. 3186.
Chem., Int. Ed. 2018, 57, 13780−13783. (310) Wang, C.; Liu, X.; Keser Demir, N.; Chen, J. P.; Li, K.
(290) Getman, R. B.; Bae, Y. S.; Wilmer, C. E.; Snurr, R. Q. Review Applications of Water Stable Metal-Organic Frameworks. Chem. Soc.
and Analysis of Molecular Simulations of Methane, Hydrogen, and Rev. 2016, 45, 5107−5134.
Acetylene Storage in Metal-Organic Frameworks. Chem. Rev. 2012, (311) Dietzel, P. D. C.; Johnsen, R. E.; Blom, R.; Fjellvåg, H.
112, 703−723. Structural Changes and Coordinatively Unsaturated Metal Atoms on
(291) Langmi, H. W.; Ren, J.; North, B.; Mathe, M.; Bessarabov, D. Dehydration of Honeycomb Analogous Microporous Metal-Organic
Hydrogen Storage in Metal-Organic Frameworks. Electrochim. Acta Frameworks. Chem. - Eur. J. 2008, 14, 2389−2397.
2014, 128, 368−392. (312) Gelfand, B. S.; Shimizu, G. K. H. Parameterizing and Grading
(292) Ahmed, A.; Seth, S.; Purewal, J.; Wong-Foy, A. G.; Veenstra, Hydrolytic Stability in Metal-Organic Frameworks. Dalt. Trans. 2016,
M.; Matzger, A. J.; Siegel, D. J. Exceptional Hydrogen Storage 45, 3668−3678.
Achieved by Screening Nearly Half a Million Metal-Organic (313) Cychosz, K. A.; Matzger, A. J. Water Stability of Microporous
Frameworks. Nat. Commun. 2019, 10, 1568. Coordination Polymers and the Adsorption of Pharmaceuticals from
(293) Trickett, C. A.; Helal, A.; Al-Maythalony, B. A.; Yamani, Z. H.; Water. Langmuir 2010, 26, 17198−17202.
Cordova, K. E.; Yaghi, O. M. The Chemistry of Metal-Organic (314) Küsgens, P.; Rose, M.; Senkovska, I.; Fröde, H.; Henschel, A.;
Frameworks for CO2 Capture, Regeneration and Conversion. Nat. Siegle, S.; Kaskel, S. Characterization of Metal-Organic Frameworks
Rev. Mater. 2017, 2, 17045. by Water Adsorption. Microporous Mesoporous Mater. 2009, 120,
(294) Wang, J.; Huang, L.; Yang, R.; Zhang, Z.; Wu, J.; Gao, Y.; 325−330.
Wang, Q.; O’Hare, D.; Zhong, Z. Recent Advances in Solid Sorbents (315) Maity, R.; Singh, H. D.; Yadav, A. K.; Chakraborty, D.;
for CO2 Capture and New Development Trends. Energy Environ. Sci. Vaidhyanathan, R. Water-Stable Adenine-Based MOFs with Polar
2014, 7, 3478−3518. Pores for Selective CO2 Capture. Chem. - Asian J. 2019, 14, 3736−
(295) Sabouni, R.; Kazemian, H.; Rohani, S. Carbon Dioxide 3741.
Capturing Technologies: A Review Focusing on Metal Organic (316) Low, J. J.; Benin, A. I.; Jakubczak, P.; Abrahamian, J. F.;
Framework Materials (MOFs). Environ. Sci. Pollut. Res. 2014, 21, Faheem, S. A.; Willis, R. R. Virtual High Throughput Screening
5427−5449. Confirmed Experimentally: Porous Coordination Polymer Hydration.
(296) Yu, J.; Xie, L. H.; Li, J. R.; Ma, Y.; Seminario, J. M.; Balbuena, J. Am. Chem. Soc. 2009, 131, 15834−15842.
P. B. CO2 Capture and Separations Using MOFs: Computational and (317) Decoste, J. B.; Peterson, G. W.; Jasuja, H.; Glover, T. G.;
Experimental Studies. Chem. Rev. 2017, 117, 9674−9754. Huang, Y. G.; Walton, K. S. Stability and Degradation Mechanisms of
(297) Xiang, S.; He, Y.; Zhang, Z.; Wu, H.; Zhou, W.; Krishna, R.; Metal-Organic Frameworks Containing the Zr6O4(OH)4 Secondary
Chen, B. Microporous Metal-Organic Framework with Potential for Building Unit. J. Mater. Chem. A 2013, 1, 5642−5650.
Carbon Dioxide Capture at Ambient Conditions. Nat. Commun. (318) Greathouse, J. A.; Allendorf, M. D. The Interaction of Water
2012, 3, 954−959. with MOF-5 Simulated by Molecular Dynamics. J. Am. Chem. Soc.
(298) Liu, J.; Thallapally, P. K.; McGrail, B. P.; Brown, D. R.; Liu, J. 2006, 128, 10678−10679.
Progress in Adsorption-Based CO2 Capture by Metal-Organic (319) Tan, K.; Nijem, N.; Canepa, P.; Gong, Q.; Li, J.; Thonhauser,
Frameworks. Chem. Soc. Rev. 2012, 41, 2308−2322. T.; Chabal, Y. J. Stability and Hydrolyzation of Metal Organic
(299) Khan, N. A.; Hasan, Z.; Jhung, S. H. Adsorptive Removal of Frameworks with Paddle-Wheel SBUs upon Hydration. Chem. Mater.
Hazardous Materials Using Metal-Organic Frameworks (MOFs): A 2012, 24, 3153−3167.
Review. J. Hazard. Mater. 2013, 244−245, 444−456. (320) Serre, C.; Millange, F.; Surblé, S.; Férey, G. A Route to the
(300) Dias, E. M.; Petit, C. Towards the Use of Metal-Organic Synthesis of Trivalent Transition-Metal Porous Carboxylates with
Frameworks for Water Reuse: A Review of the Recent Advances in Trimeric Secondary Building Units. Angew. Chem., Int. Ed. 2004, 43,
the Field of Organic Pollutants Removal and Degradation and the 6285−6289.
Next Steps in the Field. J. Mater. Chem. A 2015, 3, 22484−22506. (321) Férey, C.; Mellot-Draznieks, C.; Serre, C.; Millange, F.;
(301) Ramaswamy, P.; Wong, N. E.; Shimizu, G. K. H. MOFs as Dutour, J.; Surblé, S.; Margiolaki, I. Chemistry: A Chromium
Proton Conductors - Challenges and Opportunities. Chem. Soc. Rev. Terephthalate-Based Solid with Unusually Large Pore Volumes and
2014, 43, 5913−5932. Surface Area. Science 2005, 309, 2040−2042.

BE https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

(322) Kim, M.; Cohen, S. M. Discovery, Development, and Isoreticular, Highly Stable, Porous Zirconium Oxide Based Metal-
Functionalization of Zr(IV)-Based Metal-Organic Frameworks. Organic Frameworks. Angew. Chem., Int. Ed. 2012, 51, 9267−9271.
CrystEngComm 2012, 14, 4096−4104. (341) Feng, D.; Chung, W. C.; Wei, Z.; Gu, Z. Y.; Jiang, H. L.; Chen,
(323) Bai, Y.; Dou, Y.; Xie, L. H.; Rutledge, W.; Li, J. R.; Zhou, H. Y. P.; Darensbourg, D. J.; Zhou, H. C. Construction of Ultrastable
C. Zr-Based Metal-Organic Frameworks: Design, Synthesis, Structure, Porphyrin Zr Metal-Organic Frameworks through Linker Elimination.
and Applications. Chem. Soc. Rev. 2016, 45, 2327−2367. J. Am. Chem. Soc. 2013, 135, 17105−17110.
(324) Bosch, M.; Zhang, M.; Zhou, H.-C. Increasing the Stability of (342) Jiang, H. L.; Feng, D.; Wang, K.; Gu, Z. Y.; Wei, Z.; Chen, Y.
Metal-Organic Frameworks. Adv. Chem. 2014, 2014, 182327. P.; Zhou, H. C. An Exceptionally Stable, Porphyrinic Zr Metal-
(325) Furukawa, H.; Gándara, F.; Zhang, Y. B.; Jiang, J.; Queen, W. Organic Framework Exhibiting PH-Dependent Fluorescence. J. Am.
L.; Hudson, M. R.; Yaghi, O. M. Water Adsorption in Porous Metal- Chem. Soc. 2013, 135, 13934−13938.
Organic Frameworks and Related Materials. J. Am. Chem. Soc. 2014, (343) Feng, D.; Gu, Z. Y.; Li, J. R.; Jiang, H. L.; Wei, Z.; Zhou, H. C.
136, 4369−4381. Zirconium-Metalloporphyrin PCN-222: Mesoporous Metal-Organic
(326) Park, K. S.; Ni, Z.; Cote, A. P.; Choi, J. Y.; Huang, R.; Uribe- Frameworks with Ultrahigh Stability as Biomimetic Catalysts. Angew.
Romo, F. J.; Chae, H. K.; O’Keeffe, M.; Yaghi, O. M. Exceptional Chem., Int. Ed. 2012, 51, 10307−10310.
Chemical and Thermal Stability of Zeolitic Imidazolate Frameworks. (344) Bon, V.; Senkovska, I.; Baburin, I. A.; Kaskel, S. Zr- and Hf-
Proc. Natl. Acad. Sci. U. S. A. 2006, 103, 10186−10191. Based Metal-Organic Frameworks: Tracking down the Polymor-
(327) Choi, H. J.; Dincǎ, M.; Dailly, A.; Long, J. R. Hydrogen phism. Cryst. Growth Des. 2013, 13, 1231−1237.
Storage in Water-Stable Metal-Organic Frameworks Incorporating (345) Taylor, J. M.; Vaidhyanathan, R.; Iremonger, S. S.; Shimizu, G.
1,3- and 1,4-Benzenedipyrazolate. Energy Environ. Sci. 2010, 3, 117− K. H. Enhancing Water Stability of Metal-Organic Frameworks via
123. Phosphonate Monoester Linkers. J. Am. Chem. Soc. 2012, 134,
(328) Colombo, V.; Galli, S.; Choi, H. J.; Han, G. D.; Maspero, A.; 14338−14340.
Palmisano, G.; Masciocchi, N.; Long, J. R. High Thermal and (346) Gelfand, B. S.; Lin, J.-B.; Shimizu, G. K. H. Design of a
Chemical Stability in Pyrazolate-Bridged Metal-Organic Frameworks Humidity-Stable Metal-Organic Framework Using a Phosphonate
with Exposed Metal Sites. Chem. Sci. 2011, 2, 1311−1319. Monoester Ligand. Inorg. Chem. 2015, 54, 1185−1187.
(329) Kang, I. J.; Khan, N. A.; Haque, E.; Jhung, S. H. Chemical and (347) Chen, X.; Peng, Y.; Han, X.; Liu, Y.; Lin, X.; Cui, Y. Sixteen
Thermal Stability of Isotypic Metal-Organic Frameworks: Effect of Isostructural Phosphonate Metal-Organic Frameworks with Con-
Metal Ions. Chem. - Eur. J. 2011, 17, 6437−6442. trolled Lewis Acidity and Chemical Stability for Asymmetric Catalysis.
(330) Jiao, Y.; Morelock, C. R.; Burtch, N. C.; Mounfield, W. P.; Nat. Commun. 2017, 8, 2171.
Hungerford, J. T.; Walton, K. S. Tuning the Kinetic Water Stability (348) Emerson, A. J.; Hawes, C. S.; Marshall, M.; Knowles, G. P.;
and Adsorption Interactions of Mg-MOF-74 by Partial Substitution Chaffee, A. L.; Batten, S. R.; Turner, D. R. High-Connectivity
with Co or Ni. Ind. Eng. Chem. Res. 2015, 54, 12408−12414. Approach to a Hydrolytically Stable Metal-Organic Framework for
(331) Howe, J. D.; Morelock, C. R.; Jiao, Y.; Chapman, K. W.; CO2 Capture from Flue Gas. Chem. Mater. 2018, 30, 6614−6618.
Walton, K. S.; Sholl, D. S. Understanding Structure, Metal (349) Jasuja, H.; Walton, K. S. Effect of Catenation and Basicity of
Distribution, and Water Adsorption in Mixed-Metal MOF-74. J. Pillared Ligands on the Water Stability of MOFs. Dalt. Trans. 2013,
Phys. Chem. C 2017, 121, 627−635. 42, 15421−15426.
(332) Zuluaga, S.; Fuentes-Fernandez, E. M. A.; Tan, K.; Xu, F.; Li, (350) Gupta, V.; Mandal, S. K. A Robust and Water-Stable Two-
J.; Chabal, Y. J.; Thonhauser, T. Understanding and Controlling Fold Interpenetrated Metal-Organic Framework Containing Both
Water Stability of MOF-74. J. Mater. Chem. A 2016, 4, 5176−5183. Rigid Tetrapodal Carboxylate and Rigid Bifunctional Nitrogen
(333) Jasuja, H.; Walton, K. S. Experimental Study of CO2, CH4, Linkers Exhibiting Selective CO2 Capture. Dalt. Trans. 2019, 48,
and Water Vapor Adsorption on a Dimethyl-Functionalized UiO-66 415−425.
Framework. J. Phys. Chem. C 2013, 117, 7062−7068. (351) Pal, A.; Chand, S.; Das, M. C. A Water-Stable Twofold
(334) Bellarosa, L.; Gutiérrez-Sevillano, J. J.; Calero, S.; López, N. Interpenetrating Microporous MOF for Selective CO2 Adsorption
How Ligands Improve the Hydrothermal Stability and Affect the and Separation. Inorg. Chem. 2017, 56, 13991−13997.
Adsorption in the IRMOF Family. Phys. Chem. Chem. Phys. 2013, 15, (352) Zheng, B.; Luo, X.; Wang, Z.; Zhang, S.; Yun, R.; Huang, L.;
17696−17704. Zeng, W.; Liu, W. An Unprecedented Water Stable Acylamide-
(335) Wu, T.; Shen, L.; Luebbers, M.; Hu, C.; Chen, Q.; Ni, Z.; Functionalized Metal-Organic Framework for Highly Efficient CH4/
Masel, R. I. Enhancing the Stability of Metal-Organic Frameworks in CO2 Gas Storage/Separation and Acid-Base Cooperative Catalytic
Humid Air by Incorporating Water Repellent Functional Groups. Activity. Inorg. Chem. Front. 2018, 5, 2355−2363.
Chem. Commun. 2010, 46, 6120−6122. (353) Schoenecker, P. M.; Carson, C. G.; Jasuja, H.; Flemming, C. J.
(336) Cadiau, A.; Belmabkhout, Y.; Adil, K.; Bhatt, P. M.; Pillai, R. J.; Walton, K. S. Effect of Water Adsorption on Retention of Structure
S.; Shkurenko, A.; Martineau-Corcos, C.; Maurin, G.; Eddaoudi, M. and Surface Area of Metal-Organic Frameworks. Ind. Eng. Chem. Res.
Molecular Sorption: Hydrolytically Stable Fluorinated Metal-Organic 2012, 51, 6513−6519.
Frameworks for Energy-Efficient Dehydration. Science 2017, 356, (354) Jasuja, H.; Huang, Y. G.; Walton, K. S. Adjusting the Stability
731−735. of Metal-Organic Frameworks under Humid Conditions by Ligand
(337) Zhang, Z.; Nguyen, H. T. H.; Miller, S. A.; Ploskonka, A. M.; Functionalization. Langmuir 2012, 28, 16874−16880.
Decoste, J. B.; Cohen, S. M. Polymer-Metal-Organic Frameworks (355) Jasuja, H.; Burtch, N. C.; Huang, Y. G.; Cai, Y.; Walton, K. S.
(PolyMOFs) as Water Tolerant Materials for Selective Carbon Kinetic Water Stability of an Isostructural Family of Zinc-Based
Dioxide Separations. J. Am. Chem. Soc. 2016, 138, 920−925. Pillared Metal-Organic Frameworks. Langmuir 2013, 29, 633−642.
(338) Kitao, T.; Zhang, Y.; Kitagawa, S.; Wang, B.; Uemura, T. (356) Tang, Y.; Kourtellaris, A.; Tasiopoulos, A. J.; Teat, S. J.;
Hybridization of MOFs and Polymers. Chem. Soc. Rev. 2017, 46, Dubbeldam, D.; Rothenberg, G.; Tanase, S. Selective CO2
3108−3133. Adsorption in Water-Stable Alkaline-Earth Based Metal-Organic
(339) Decoste, J. B.; Peterson, G. W.; Schindler, B. J.; Killops, K. L.; Frameworks. Inorg. Chem. Front. 2018, 5, 541−549.
Browe, M. A.; Mahle, J. J. The Effect of Water Adsorption on the (357) Mondloch, J. E.; Katz, M. J.; Planas, N.; Semrouni, D.;
Structure of the Carboxylate Containing Metal-Organic Frameworks Gagliardi, L.; Hupp, J. T.; Farha, O. K. Are Zr6-Based MOFs Water
Cu-BTC, Mg-MOF-74, and UiO-66. J. Mater. Chem. A 2013, 1, Stable? Linker Hydrolysis vs. Capillary-Force-Driven Channel
11922−11932. Collapse. Chem. Commun. 2014, 50, 8944−8946.
(340) Guillerm, V.; Ragon, F.; Dan-Hardi, M.; Devic, T.; (358) Brinker, C. J.; Scherer, G. W. Drying. In Sol−Gel Science, The
Vishnuvarthan, M.; Campo, B.; Vimont, A.; Clet, G.; Yang, Q.; Physics and Chemistry of Sol−Gel Processing; Academic Press: San
Maurin, G.; Férey, G.; Vittadini, A.; Gross, S.; Serre, C. A Series of Diego, 1990; Chapter 8, pp 452−513.

BF https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

(359) Millange, F.; Guillou, N.; Walton, R. I.; Grenèche, J. M.; (377) Ho, N. L.; Porcheron, F.; Pellenq, R. J. M. Experimental and
Margiolaki, I.; Férey, G. Effect of the Nature of the Metal on the Molecular Simulation Investigation of Enhanced CO2 Solubility in
Breathing Steps in MOFs with Dynamic Frameworks. Chem. Hybrid Adsorbents. Langmuir 2010, 26, 13287−13296.
Commun. 2008, 4732−4734. (378) Ho, L. N.; Perez Pellitero, J.; Porcheron, F.; Pellenq, R. J. M.
(360) Kökçam-Demir, Ü .; Goldman, A.; Esrafili, L.; Gharib, M.; Enhanced CO2 Solubility in Hybrid MCM-41: Molecular Simulations
Morsali, A.; Weingart, O.; Janiak, C. Coordinatively Unsaturated and Experiments. Langmuir 2011, 27, 8187−8197.
Metal Sites (Open Metal Sites) in Metal-Organic Frameworks: (379) González-Martínez, G. A.; Jurado-Vázquez, T.; Solís-Ibarra,
Design and Applications. Chem. Soc. Rev. 2020, 49, 2751−2798. D.; Vargas, B.; Sánchez-González, E.; Martínez, A.; Vargas, R.;
(361) Gul-E-Noor, F.; Jee, B.; Pöppl, A.; Hartmann, M.; Himsl, D.; González-Zamora, E.; Ibarra, I. A. Confinement of H2O and EtOH to
Bertmer, M. Effects of Varying Water Adsorption on a Cu3(BTC)2 Enhance CO2 Capture in MIL-53(Al)-TDC. Dalt. Trans. 2018, 47,
Metal-Organic Framework (MOF) as Studied by 1H and 13C Solid- 9459−9465.
State NMR Spectroscopy. Phys. Chem. Chem. Phys. 2011, 13, 7783− (380) Sánchez-González, E.; Mileo, P. G. M.; Sagastuy-Breña, M.;
7788. Á lvarez, J. R.; Reynolds, J. E.; Villarreal, A.; Gutiérrez-Alejandre, A.;
(362) Yu, J.; Ma, Y.; Balbuena, P. B. Evaluation of the Impact of Ramírez, J.; Balmaseda, J.; González-Zamora, E.; Maurin, G.;
H2O, O2, and SO2 on Postcombustion CO2 Capture in Metal- Humphrey, S. M.; Ibarra, I. A. Highly Reversible Sorption of H2S
Organic Frameworks. Langmuir 2012, 28, 8064−8071. and CO2 by an Environmentally Friendly Mg-Based MOF. J. Mater.
(363) Kizzie, A. C.; Wong-Foy, A. G.; Matzger, A. J. Effect of Chem. A 2018, 6, 16900−16909.
Humidity on the Performance of Microporous Coordination (381) Sagastuy-Breña, M.; Mileo, P. G. M.; Sánchez-González, E.;
Polymers as Adsorbents for CO2 Capture. Langmuir 2011, 27, Reynolds, J. E.; Jurado-Vázquez, T.; Balmaseda, J.; González-Zamora,
6368−6373. E.; Devautour-Vinot, S.; Humphrey, S. M.; Maurin, G.; Ibarra, I. A.
(364) Liu, J.; Benin, A. I.; Furtado, A. M. B.; Jakubczak, P.; Willis, R. Humidity-Induced CO2 Capture Enhancement in Mg-CUK-1. Dalt.
R.; Levan, M. D. Stability Effects on CO2 Adsorption for the DOBDC Trans. 2018, 47, 15827−15834.
Series of Metal-Organic Frameworks. Langmuir 2011, 27, 11451− (382) Sánchez-Serratos, M.; Bayliss, P. A.; Peralta, R. A.; González-
11456. Zamora, E.; Lima, E.; Ibarra, I. A. CO2 Capture in the Presence of
(365) Liu, J.; Tian, J.; Thallapally, P. K.; McGrail, B. P. Selective Water Vapour in MIL-53(Al). New J. Chem. 2016, 40, 68−72.
CO2 Capture from Flue Gas Using Metal-Organic Frameworks - A (383) Zárate, A.; Peralta, R. A.; Bayliss, P. A.; Howie, R.; Sánchez-
Fixed Bed Study. J. Phys. Chem. C 2012, 116, 9575−9581. Serratos, M.; Carmona-Monroy, P.; Solis-Ibarra, D.; González-
(366) Yu, J.; Balbuena, P. B. Water Effects on Postcombustion CO2 Zamora, E.; Ibarra, I. A. CO2 Capture under Humid Conditions in
Capture in Mg-MOF-74. J. Phys. Chem. C 2013, 117, 3383−3388.
NH2-MIL-53(Al): The Influence of the Amine Functional Group.
(367) Lin, L. C.; Lee, K.; Gagliardi, L.; Neaton, J. B.; Smit, B. Force-
RSC Adv. 2016, 6, 9978−9983.
Field Development from Electronic Structure Calculations with
(384) Liao, P. Q.; Chen, H.; Zhou, D. D.; Liu, S. Y.; He, C. T.; Rui,
Periodic Boundary Conditions: Applications to Gaseous Adsorption
Z.; Ji, H.; Zhang, J. P.; Chen, X. M. Monodentate Hydroxide as a
and Transport in Metal-Organic Frameworks. J. Chem. Theory
Super Strong yet Reversible Active Site for CO2 Capture from High-
Comput. 2014, 10, 1477−1488.
Humidity Flue Gas. Energy Environ. Sci. 2015, 8, 1011−1016.
(368) Xian, S.; Peng, J.; Zhang, Z.; Xia, Q.; Wang, H.; Li, Z. Highly
(385) Wright, A. M.; Wu, Z.; Zhang, G.; Mancuso, J. L.; Comito, R.
Enhanced and Weakened Adsorption Properties of Two MOFs by
J.; Day, R. W.; Hendon, C. H.; Miller, J. T.; Dincá, M. A Structural
Water Vapor for Separation of CO2/CH4 and CO2/N2 Binary
Mixtures. Chem. Eng. J. 2015, 270, 385−392. Mimic of Carbonic Anhydrase in a Metal-Organic Framework. Chem.
(369) Coelho, J. A.; Ribeiro, A. M.; Ferreira, A. F. P.; Lucena, S. M. 2018, 4, 2894−2901.
P.; Rodrigues, A. E.; de Azevedo, D. C. S. Stability of an Al-Fumarate (386) Bien, C. E.; Chen, K. K.; Chien, S. C.; Reiner, B. R.; Lin, L. C.;
MOF and Its Potential for CO2 Capture from Wet Stream. Ind. Eng. Wade, C. R.; Ho, W. S. W. Bioinspired Metal-Organic Framework for
Chem. Res. 2016, 55, 2134−2143. Trace CO2 Capture. J. Am. Chem. Soc. 2018, 140, 12662−12666.
(370) Coelho, J. A.; Lima, A. E. O.; Rodrigues, A. E.; de Azevedo, D. (387) Bien, C. E.; Liu, Q.; Wade, C. R. Assessing the Role of Metal
C. S.; Lucena, S. M. P. Computer Simulation of Adsorption and Identity on CO2 Adsorption in MOFs Containing M-OH Functional
Sitting of CO2, N2, CH4 and Water on a New Al(OH)-Fumarate Groups. Chem. Mater. 2020, 32, 489−497.
MOF. Adsorption 2017, 23, 423−431. (388) Jayaramulu, K.; Geyer, F.; Schneemann, A.; Kment, Š .;
(371) Zhang, Z.; Li, Z.; Li, J. Computational Study of Adsorption Otyepka, M.; Zboril, R.; Vollmer, D.; Fischer, R. A. Hydrophobic
and Separation of CO2, CH4, and N2 by an Rht-Type Metal - Metal-Organic Frameworks. Adv. Mater. 2019, 31, 1970230.
Organic Framework. Langmuir 2012, 28, 12122−12133. (389) Xie, L. H.; Xu, M. M.; Liu, X. M.; Zhao, M. J.; Li, J. R.
(372) Masala, A.; Vitillo, J. G.; Mondino, G.; Grande, C. A.; Blom, Hydrophobic Metal-Organic Frameworks: Assessment, Construction,
R.; Manzoli, M.; Marshall, M.; Bordiga, S. CO2 Capture in Dry and and Diverse Applications. Adv. Sci. 2020, 7, 1901758.
Wet Conditions in UTSA-16 Metal-Organic Framework. ACS Appl. (390) Nguyen, N. T. T.; Furukawa, H.; Gándara, F.; Nguyen, H. T.;
Mater. Interfaces 2017, 9, 455−463. Cordova, K. E.; Yaghi, O. M. Selective Capture of Carbon Dioxide
(373) Á lvarez, J. R.; Peralta, R. A.; Balmaseda, J.; González-Zamora, under Humid Conditions by Hydrophobic Chabazite-Type Zeolitic
E.; Ibarra, I. A. Water Adsorption Properties of a Sc(III) Porous Imidazolate Frameworks. Angew. Chem., Int. Ed. 2014, 53, 10645−
Coordination Polymer for CO2 Capture Applications. Inorg. Chem. 10648.
Front. 2015, 2, 1080−1084. (391) Liu, B.; Smit, B. Molecular Simulation Studies of Separation of
(374) Lara-García, H. A.; Gonzalez, M. R.; González-Estefan, J. H.; CO2 /N2, CO2 /CH4, and CH4 /N2 by ZIFs. J. Phys. Chem. C 2010,
Sánchez-Camacho, P.; Lima, E.; Ibarra, I. A. Removal of CO2 from 114, 8515−8522.
CH4 and CO2 Capture in the Presence of H2O Vapour in NOTT- (392) Liu, Y.; Liu, J.; Lin, Y. S.; Chang, M. Effects of Water Vapor
401. Inorg. Chem. Front. 2015, 2, 442−447. and Trace Gas Impurities in Flue Gas on CO2/N2 Separation Using
(375) López-Cervantes, V. B.; Sánchez-González, E.; Jurado- ZIF-68. J. Phys. Chem. C 2014, 118, 6744−6751.
Vázquez, T.; Tejeda-Cruz, A.; González-Zamora, E.; Ibarra, I. A. (393) Bhatt, P. M.; Belmabkhout, Y.; Cadiau, A.; Adil, K.; Shekhah,
CO2 Adsorption under Humid Conditions: Self-Regulated Water O.; Shkurenko, A.; Barbour, L. J.; Eddaoudi, M. A Fine-Tuned
Content in CAU-10. Polyhedron 2018, 155, 163−169. Fluorinated MOF Addresses the Needs for Trace CO2 Removal and
(376) Chen, Y.; Qiao, Z.; Huang, J.; Wu, H.; Xiao, J.; Xia, Q.; Xi, H.; Air Capture Using Physisorption. J. Am. Chem. Soc. 2016, 138, 9301−
Hu, J.; Zhou, J.; Li, Z. Unusual Moisture-Enhanced CO2 Capture 9307.
within Microporous PCN-250 Frameworks. ACS Appl. Mater. (394) Belmabkhout, Y.; Bhatt, P. M.; Adil, K.; Pillai, R. S.; Cadiau,
Interfaces 2018, 10, 38638−38647. A.; Shkurenko, A.; Maurin, G.; Liu, G.; Koros, W. J.; Eddaoudi, M.

BG https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Natural Gas Upgrading Using a Fluorinated MOF with Tuned H2S (410) McDonald, T. M.; Lee, W. R.; Mason, J. A.; Wiers, B. M.;
and CO2 Adsorption Selectivity. Nat. Energy 2018, 3, 1059−1066. Hong, C. S.; Long, J. R. Capture of Carbon Dioxide from Air and Flue
(395) Moghadam, P. Z.; Ivy, J. F.; Arvapally, R. K.; Dos Santos, A. Gas in the Alkylamine-Appended Metal-Organic Framework Mmen-
M.; Pearson, J. C.; Zhang, L.; Tylianakis, E.; Ghosh, P.; Oswald, I. W. Mg2(Dobpdc). J. Am. Chem. Soc. 2012, 134, 7056−7065.
H.; Kaipa, U.; Wang, X.; Wilson, A. K.; Snurr, R. Q.; Omary, M. A. (411) Siegelman, R. L.; Milner, P. J.; Forse, A. C.; Lee, J.-H. H.;
Adsorption and Molecular Siting of CO2, Water, and Other Gases in Colwell, K. A.; Neaton, J. B.; Reimer, J. A.; Weston, S. C.; Long, J. R.
the Superhydrophobic, Flexible Pores of FMOF-1 from Experiment Water Enables Efficient CO2 Capture from Natural Gas Flue
and Smulation. Chem. Sci. 2017, 8, 3989−4000. Emissions in an Oxidation-Resistant Diamine-Appended Metal-
(396) Yu, J.; Balbuena, P. B. How Impurities Affect CO2 Capture in Organic Framework. J. Am. Chem. Soc. 2019, 141, 13171−13186.
Metal-Organic Frameworks Modified with Different Functional (412) Milner, P. J.; Martell, J. D.; Siegelman, R. L.; Gygi, D.;
Groups. ACS Sustainable Chem. Eng. 2015, 3, 117−124. Weston, S. C.; Long, J. R. Overcoming Double-Step CO2 Adsorption
(397) Bae, Y. S.; Liu, J.; Wilmer, C. E.; Sun, H.; Dickey, A. N.; Kim, and Minimizing Water Co-Adsorption in Bulky Diamine-Appended
M. B.; Benin, A. I.; Willis, R. R.; Barpaga, D.; LeVan, M. D.; Snurr, R. Variants of Mg2(Dobpdc). Chem. Sci. 2018, 9, 160−174.
Q. The Effect of Pyridine Modification of Ni-DOBDC on CO2 (413) McDonald, T. M.; D’Alessandro, D. M.; Krishna, R.; Long, J.
Capture under Humid Conditions. Chem. Commun. 2014, 50, 3296− R. Enhanced Carbon Dioxide Capture upon Incorporation of N,N′-
3298. Dimethylethylenediamine in the Metal-Organic Framework CuBTTri.
(398) Zhang, W.; Hu, Y.; Ge, J.; Jiang, H. L.; Yu, S. H. A Facile and Chem. Sci. 2011, 2, 2022−2028.
General Coating Approach to Moisture/Water-Resistant Metal- (414) Luz, I.; Soukri, M.; Lail, M. Synthesis of Fluidized CO2
Organic Frameworks with Intact Porosity. J. Am. Chem. Soc. 2014, Sorbents Based on Diamine Coordinated to Metal-Organic Frame-
136, 16978−16981. works by Direct Conversion of Metal Oxides Supported on
(399) Fernandez, C. A.; Nune, S. K.; Annapureddy, H. V.; Dang, L. Mesoporous Silica. Chem. - Eur. J. 2018, 24, 10612−10616.
X.; McGrail, B. P.; Zheng, F.; Polikarpov, E.; King, D. L.; Freeman, (415) Wu, Y.; Lv, Z.; Zhou, X.; Peng, J.; Tang, Y.; Li, Z. Tuning
C.; Brooks, K. P. Hydrophobic and Moisture-Stable Metal-Organic Secondary Building Unit of Cu-BTC to Simultaneously Enhance Its
Frameworks. Dalt. Trans. 2015, 44, 13490−13497. CO2 Selective Adsorption and Stability under Moisture. Chem. Eng. J.
(400) Kang, M.; Kim, J. E.; Kang, D. W.; Lee, H. Y.; Moon, D.; 2019, 355, 815−821.
Hong, C. S. A Diamine-Grafted Metal-Organic Framework with (416) Montoro, C.; García, E.; Calero, S.; Pérez-Fernández, M. A.;
Outstanding CO2 Capture Properties and a Facile Coating Approach López, A. L.; Barea, E.; Navarro, J. A. R. Functionalisation of MOF
for Imparting Exceptional Moisture Stability. J. Mater. Chem. A 2019, Open Metal Sites with Pendant Amines for CO2 Capture. J. Mater.
7, 8177−8183. Chem. 2012, 22, 10155−10158.
(401) Peng, L.; Yang, S.; Sun, D. T.; Asgari, M.; Queen, W. L. (417) Vlaisavljevich, B.; Odoh, S. O.; Schnell, S. K.; Dzubak, A. L.;
MOF/Polymer Composite Synthesized Using a Double Solvent Lee, K.; Planas, N.; Neaton, J. B.; Gagliardi, L.; Smit, B. CO2 Induced
Phase Transitions in Diamine-Appended Metal-Organic Frameworks.
Method Offers Enhanced Water and CO2 Adsorption Properties.
Chem. Sci. 2015, 6, 5177−5185.
Chem. Commun. 2018, 54, 10602−10605.
(418) Jeon, S.; Jung, H.; Jo, D. H.; Kim, S. H. Development of
(402) Ding, N.; Li, H.; Feng, X.; Wang, Q.; Wang, S.; Ma, L.; Zhou,
Crosslinked PEI Solid Adsorbents for CO2 Capture. Energy Procedia
J.; Wang, B. Partitioning MOF-5 into Confined and Hydrophobic
2017, 114, 2287−2293.
Compartments for Carbon Capture under Humid Conditions. J. Am.
(419) Kim, E. J.; Siegelman, R. L.; Jiang, H. Z. H.; Forse, A. C.; Lee,
Chem. Soc. 2016, 138, 10100−10103.
J.-H.; Martell, J. D.; Milner, P. J.; Falkowski, J. M.; Neaton, J. B.;
(403) Qian, X.; Sun, F.; Sun, J.; Wu, H.; Xiao, F.; Wu, X.; Zhu, G.
Reimer, J. A.; Weston, S. C.; Long, J. R. Cooperative Carbon Capture
Imparting Surface Hydrophobicity to Metal-Organic Frameworks
and Steam Regeneration with Tetraamine-Appended Metal-Organic
Using a Facile Solution-Immersion Process to Enhance Water Frameworks. Science 2020, 369, 392−396.
Stability for CO2 Capture. Nanoscale 2017, 9, 2003−2008. (420) Loiseau, T.; Serre, C.; Huguenard, C.; Fink, G.; Taulelle, F.;
(404) Nandi, S.; Haldar, S.; Chakraborty, D.; Vaidhyanathan, R. Henry, M.; Bataille, T.; Férey, G. A Rationale for the Large Breathing
Strategically Designed Azolyl-Carboxylate MOFs for Potential Humid of the Porous Aluminum Terephthalate (MIL-53) Upon Hydration.
CO2 Capture. J. Mater. Chem. A 2017, 5, 535−543. Chem. - Eur. J. 2004, 10, 1373−1382.
(405) Shi, Z.; Tao, Y.; Wu, J.; Zhang, C.; He, H.; Long, L.; Lee, Y.; (421) Bourrelly, S.; Llewellyn, P. L.; Serre, C.; Millange, F.; Loiseau,
Li, T.; Zhang, Y. B. Robust Metal-Triazolate Frameworks for CO2 T.; Férey, G. Different Adsorption Behaviors of Methane and Carbon
Capture from Flue Gas. J. Am. Chem. Soc. 2020, 142, 2750−2754. Dioxide in the Isotypic Nanoporous Metal Terephthalates MIL-53
(406) Boyd, P. G.; Chidambaram, A.; García-Díez, E.; Ireland, C. P.; and MIL-47. J. Am. Chem. Soc. 2005, 127, 13519−13521.
Daff, T. D.; Bounds, R.; Gładysiak, A.; Schouwink, P.; Moosavi, S. M.; (422) Madden, D. G.; Scott, H. S.; Kumar, A.; Chen, K. J.; Sanii, R.;
Maroto-Valer, M. M.; Reimer, J. A.; Navarro, J. A. R.; Woo, T. K.; Bajpai, A.; Lusi, M.; Curtin, T.; Perry, J. J.; Zaworotko, M. J. Flue-Gas
Garcia, S.; Stylianou, K. C.; Smit, B. Data-Driven Design of Metal- and Direct-Air Capture of CO2 by Porous Metal-Organic Materials.
Organic Frameworks for Wet Flue Gas CO2 Capture. Nature 2019, Philos. Trans. R. Soc., A 2017, 375, 20160025.
576, 253−256. (423) Mukherjee, S.; Sikdar, N.; O’Nolan, D.; Franz, D. M.; Gascón,
(407) Fracaroli, A. M.; Furukawa, H.; Suzuki, M.; Dodd, M.; V.; Kumar, A.; Kumar, N.; Scott, H. S.; Madden, D. G.; Kruger, P. E.;
Okajima, S.; Gándara, F.; Reimer, J. A.; Yaghi, O. M. Metal-Organic Space, B.; Zaworotko, M. J. Trace CO2 Capture by an Ultra-
Frameworks with Precisely Designed Interior for Carbon Dioxide microporous Physisorbent with Low Water Affinity. Sci. Adv. 2019, 5,
Capture in the Presence of Water. J. Am. Chem. Soc. 2014, 136, 8863− No. eaax9171.
8866. (424) Kumar, A.; Hua, C.; Madden, D. G.; O’Nolan, D.; Chen, K. J.;
(408) Lee, W. R.; Jo, H.; Yang, L. M.; Lee, H.; Ryu, D. W.; Lim, K. Keane, L. A. J.; Perry, J. J.; Zaworotko, M. J. Hybrid Ultramicroporous
S.; Song, J. H.; Min, D. Y.; Han, S. S.; Seo, J. G.; Park, Y. K.; Moon, Materials (HUMs) with Enhanced Stability and Trace Carbon
D.; Hong, C. S. Exceptional CO2 Working Capacity in a Capture Performance. Chem. Commun. 2017, 53, 5946−5949.
Heterodiamine-Grafted Metal-Organic Framework. Chem. Sci. 2015, (425) Nugent, P.; Belmabkhout, Y.; Burd, S. D.; Cairns, A. J.;
6, 3697−3705. Luebke, R.; Forrest, K.; Pham, T.; Ma, S.; Space, B.; Wojtas, L.;
(409) Yeon, J. S.; Lee, W. R.; Kim, N. W.; Jo, H.; Lee, H.; Song, J. Eddaoudi, M.; Zaworotko, M. J. Porous Materials with Optimal
H.; Lim, K. S.; Kang, D. W.; Seo, J. G.; Moon, D.; Wiers, B.; Hong, C. Adsorption Thermodynamics and Kinetics for CO2 Separation.
S. Homodiamine-Functionalized Metal-Organic Frameworks with a Nature 2013, 495, 80−84.
MOF-74-Type Extended Structure for Superior Selectivity of CO2 (426) Shekhah, O.; Belmabkhout, Y.; Chen, Z.; Guillerm, V.; Cairns,
over N2. J. Mater. Chem. A 2015, 3, 19177−19185. A.; Adil, K.; Eddaoudi, M. Made-to-Order Metal-Organic Frameworks

BH https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

for Trace Carbon Dioxide Removal and Air Capture. Nat. Commun. Process Using PEI Impregnated Mesoporous Precipitated Silica.
2014, 5, 4228. Greenhouse Gases: Sci. Technol. 2015, 5, 91−101.
(427) Loiseau, T.; Mellot-Draznieks, C.; Muguerra, H.; Férey, G.; (446) Lara, Y.; Romeo, L. M. Amine-Impregnated Alumina Solid
Haouas, M.; Taulelle, F. Hydrothermal Synthesis and Crystal Sorbents for CO2 Capture. Lessons Learned. Energy Procedia 2017,
Structure of a New Three-Dimensional Aluminum-Organic Frame- 114, 2372−2379.
work MIL-69 with 2,6-Naphthalenedicarboxylate (Ndc), Al(OH)- (447) Gholidoust, A.; Atkinson, J. D.; Hashisho, Z. Enhancing CO2
(Ndc)·H2O. C. R. Chim. 2005, 8, 765−772. Adsorption via Amine-Impregnated Activated Carbon from Oil Sands
(428) Huang, H.; Zhang, W.; Liu, D.; Zhong, C. Understanding the Coke. Energy Fuels 2017, 31, 1756−1763.
Effect of Trace Amount of Water on CO2 Capture in Natural Gas (448) Gibson, J. A. A.; Gromov, A. V.; Brandani, S.; Campbell, E. E.
Upgrading in Metal - Organic Frameworks: A Molecular Simulation B. Comparison of Amine-Impregnated Mesoporous Carbon with
Study. Ind. Eng. Chem. Res. 2012, 51, 10031−10038. Microporous Activated Carbon and 13X Zeolite for Biogas
(429) Bahamon, D.; Vega, L. F. Systematic Evaluation of Materials Purification. J. Porous Mater. 2017, 24, 1473−1479.
for Post-Combustion CO2 Capture in a Temperature Swing (449) Karka, S.; Kodukula, S.; Nandury, S. V.; Pal, U.
Adsorption Process. Chem. Eng. J. 2016, 284, 438−447. Polyethylenimine-Modified Zeolite 13X for CO2 Capture: Adsorp-
(430) Liang, Z.; Marshall, M.; Chaffee, A. L. CO2 Adsorption-Based tion and Kinetic Studies. ACS Omega 2019, 4, 16441−16449.
Separation by Metal Organic Framework (Cu-BTC) versus Zeolite (450) Gómez-Pozuelo, G.; Sanz-Pérez, E. S.; Arencibia, A.; Pizarro,
(13X). Energy Fuels 2009, 23, 2785−2789. P.; Sanz, R.; Serrano, D. P. CO2 Adsorption on Amine-Functionalized
(431) Kohl, A. L.; Nielsen, R. Gas Purification, 5th ed.; Gulf Clays. Microporous Mesoporous Mater. 2019, 282, 38−47.
Professional Publishing: Houston, 1997. (451) Bali, S.; Leisen, J.; Foo, G. S.; Sievers, C.; Jones, C. W.
(432) Dutcher, B.; Fan, M.; Russell, A. G. Amine-Based CO2 Aminosilanes Grafted to Basic Alumina as CO2 Adsorbents-Role of
Capture Technology Development from the Beginning of 2013-A Grafting Conditions on CO2 Adsorption Properties. ChemSusChem
Review. ACS Appl. Mater. Interfaces 2015, 7, 2137−2148. 2014, 7, 3145−3156.
(433) Mumford, K. A.; Wu, Y.; Smith, K. H.; Stevens, G. W. Review (452) Su, F.; Lu, C.; Chen, H. S. Adsorption, Desorption, and
of Solvent Based Carbon-Dioxide Capture Technologies. Front. Chem. Thermodynamic Studies of CO2 with High-Amine-Loaded Multi-
Sci. Eng. 2015, 9, 125−141. walled Carbon Nanotubes. Langmuir 2011, 27, 8090−8098.
(434) Yu, C. H.; Huang, C. H.; Tan, C. S. A Review of CO2 Capture (453) Ali Akbar Razavi, S.; Morsali, A. Linker Functionalized Metal-
by Absorption and Adsorption. Aerosol Air Qual. Res. 2012, 12, 745− Organic Frameworks. Coord. Chem. Rev. 2019, 399, 213023.
769. (454) Otsuji, K.; Hirao, M.; Satoh, S. A Regenerable Carbon Dioxide
(435) Kamijo, T.; Sorimachi, Y.; Shimada, D.; Miyamoto, O.; Endo, Removal and Oxygen Recovery System for the Japanese Experiment
T.; Nagayasu, H.; Mangiaracina, A. Result of the 60 Tpd CO2 Module. Acta Astronaut. 1987, 15, 45−54.
Capture Pilot Plant in European Coal Power Plant with KS-1 Solvent. (455) Xu, X.; Song, C.; Andresen, J. M.; Miller, B. G.; Scaroni, A. W.
Energy Procedia 2013, 37, 813−816. Novel Polyethylenimine-Modified Mesoporous Molecular Sieve of
(436) Iijima, M.; Nagayasu, T.; Kamijyo, T.; Nakatan, S. MHI’s MCM-41 Type as High-Capacity Adsorbent for CO2 Capture. Energy
Energy Efficient Flue Gas CO2 Capture Technology and Large Scale Fuels 2002, 16, 1463−1469.
CCS Demonstration Test at Coal-Fired Power Plants in USA. (456) Zhao, Y.; Zhou, J.; Fan, L.; Chen, L.; Li, L.; Xu, Z. P.; Qian, G.
Mitsubishi Heavy Ind. Technol. Rev. 2011, 48, 26−32. Indoor CO2 Control through Mesoporous Amine-Functionalized
(437) Zhao, W.; Veneman, R.; Chen, D.; Li, Z.; Cai, N.; Brilmana, Silica Monoliths. Ind. Eng. Chem. Res. 2019, 58, 19465−19474.
D. W. F. Post-Combustion CO2 Capture Demonstration Using (457) Min, K.; Choi, W.; Choi, M. Macroporous Silica with Thick
Supported Amine Sorbents: Design and Evaluation of 200 KWth Framework for Steam-Stable and High-Performance Poly-
Pilot. Energy Procedia 2014, 63, 2374−2383. (Ethyleneimine)/Silica CO2 Adsorbent. ChemSusChem 2017, 10,
(438) Yu, Q.; Delgado, J. D. L. P.; Veneman, R.; Brilman, D. W. F. 2518−2526.
Stability of a Benzyl Amine Based CO2 Capture Adsorbent in View of (458) Zhang, H.; Goeppert, A.; Olah, G. A.; Prakash, G. K. S.
Regeneration Strategies. Ind. Eng. Chem. Res. 2017, 56, 3259−3269. Remarkable Effect of Moisture on the CO2 Adsorption of Nano-Silica
(439) Sardo, M.; Afonso, R.; Juźków, J.; Pacheco, M.; Bordonhos, Supported Linear and Branched Polyethylenimine. J. CO2 Util. 2017,
M.; Pinto, M. L.; Gomes, J. R. B.; Mafra, L. Unravelling Moisture- 19, 91−99.
Induced CO2 Chemisorption Mechanisms in Amine-Modified (459) Hashida, T.; Tashiro, K.; Aoshima, S.; Inaki, Y. Structural
Sorbents at the Molecular Scale. J. Mater. Chem. A 2021, 9, 5542− Investigation on Water-Induced Phase Transitions of Poly(Ethylene
5555. Imine). 1. Time-Resolved Infrared Spectral Measurements in the
(440) Szego, A. E.; Jaworski, A.; Hedin, N. Chemisorption of CO2 Hydration Process. Macromolecules 2002, 35, 4330−4336.
on Diaminated Silica as Bicarbonates and Different Types of (460) He, H.; Zhuang, L.; Chen, S.; Liu, H. Solid Amine Adsorbent
Carbamate Ammonium Ion Pairs. Mater. Adv. 2021, 2, 448−454. Prepared by Molecular Imprinting and Its Carbon Dioxide
(441) Ben Said, R.; Kolle, J. M.; Essalah, K.; Tangour, B.; Sayari, A. Adsorption Properties. Chem. - Asian J. 2016, 11, 3055−3061.
A Unified Approach to CO2-Amine Reaction Mechanisms. ACS (461) Liu, F.; Kuang, Y.; Wang, S.; Chen, S.; Fu, W. Preparation and
Omega 2020, 5, 26125−26133. Characterization of Molecularly Imprinted Solid Amine Adsorbent for
(442) Li, K.; Kress, J. D.; Mebane, D. S. The Mechanism of CO2 CO2 Adsorption. New J. Chem. 2018, 42, 10016−10023.
Adsorption under Dry and Humid Conditions in Mesoporous Silica- (462) Koutsianos, A.; Barron, A. R.; Andreoli, E. CO2 Capture
Supported Amine Sorbents. J. Phys. Chem. C 2016, 120, 23683− Partner Molecules in Highly Loaded PEI Sorbents. J. Phys. Chem. C
23691. 2017, 121, 21772−21781.
(443) Yu, J.; Zhai, Y.; Chuang, S. S. C. Water Enhancement in CO2 (463) Hamdy, L. B.; Wakeham, R. J.; Taddei, M.; Barron, A. R.;
Capture by Amines: An Insight into CO2-H2O Interactions on Andreoli, E. Epoxy Cross-Linked Polyamine CO2 Sorbents Enhanced
Amine Films and Sorbents. Ind. Eng. Chem. Res. 2018, 57, 4052− via Hydrophobic Functionalization. Chem. Mater. 2019, 31, 4673−
4062. 4684.
(444) Heydari-Gorji, A.; Belmabkhout, Y.; Sayari, A. Polyethyleni- (464) Qi, G.; Fu, L.; Giannelis, E. P. Sponges with Covalently
mine-Impregnated Mesoporous Silica: Effect of Amine Loading and Tethered Amines for High-Efficiency Carbon Capture. Nat. Commun.
Surface Alkyl Chains on CO2 Adsorption. Langmuir 2011, 27, 2014, 5, 5796.
12411−12416. (465) He, H.; Tang, H.; Chen, X.; Hou, X.; Zhou, X.; Chen, H.; Wu,
(445) Quang, D. V.; Dindi, A.; Rayer, A. V.; Hadri, N. El; S.; Wang, S. Structure Design of Low-Temperature Regenerative
Abdulkadir, A.; Abu-Zahra, M. R. M. Effect of Moisture on the Heat Hyperbranched Polyamine Adsorbent for CO2 Capture. Langmuir
Capacity and the Regeneration Heat Required for CO2 Capture 2018, 34, 14169−14179.

BI https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

(466) Didas, S. A.; Choi, S.; Chaikittisilp, W.; Jones, C. W. Amine- (486) Climeworks; Climeworks, 2020; https://www.climeworks.
Oxide Hybrid Materials for CO2 Capture from Ambient Air. Acc. com/ (accessed 2021-03-13).
Chem. Res. 2015, 48, 2680−2687. (487) Mebane, D. S.; Kress, J. D.; Storlie, C. B.; Fauth, D. J.; Gray,
(467) Sanz-Pérez, E. S.; Murdock, C. R.; Didas, S. A.; Jones, C. W. M. L.; Li, K. Transport, Zwitterions, and the Role of Water for CO2
Direct Capture of CO2 from Ambient Air. Chem. Rev. 2016, 116, Adsorption in Mesoporous Silica-Supported Amine Sorbents. J. Phys.
11840−11876. Chem. C 2013, 117, 26617−26627.
(468) Shi, X.; Xiao, H.; Azarabadi, H.; Song, J.; Wu, X.; Chen, X.; (488) Bos, M. J.; Pietersen, S.; Brilman, D. W. F. Production of High
Lackner, K. S. Sorbents for Direct Capture of CO2 from Ambient Air. Purity CO2 from Air Using Solid Amine Sorbents. Chem. Eng. Sci. X
Angew. Chem., Int. Ed. 2020, 59, 6984−7006. 2019, 2, 100020.
(469) Yang, M.; Ma, C.; Xu, M.; Wang, S.; Xu, L. Recent Advances (489) Kim, H. J.; Chaikittisilp, W.; Jang, K. S.; Didas, S. A.; Johnson,
in CO2 Adsorption from Air: A Review. Curr. Pollut. Reports 2019, 5, J. R.; Koros, W. J.; Nair, S.; Jones, C. W. Aziridine-Functionalized
272−293. Mesoporous Silica Membranes on Polymeric Hollow Fibers: Syn-
(470) Wijesiri, R. P.; Knowles, G. P.; Yeasmin, H.; Hoadley, A. F. A.; thesis and Single-Component CO2 and N2 Permeation Properties.
Chaffee, A. L. CO2 Capture from Air Using Pelletized Polyethyle- Ind. Eng. Chem. Res. 2015, 54, 4407−4413.
nimine Impregnated MCF Silica. Ind. Eng. Chem. Res. 2019, 58, (490) Buijs, W.; De Flart, S. Direct Air Capture of CO2 with an
3293−3303. Amine Resin: A Molecular Modeling Study of the CO2 Capturing
(471) Kwon, H. T.; Sakwa-Novak, M. A.; Pang, S. H.; Sujan, A. R.; Process. Ind. Eng. Chem. Res. 2017, 56, 12297−12304.
Ping, E. W.; Jones, C. W. Aminopolymer-Impregnated Hierarchical (491) Wang, T.; Lackner, K. S.; Wright, A. Moisture Swing Sorbent
Silica Structures: Unexpected Equivalent CO2 Uptake under for Carbon Dioxide Capture from Ambient Air. Environ. Sci. Technol.
Simulated Air Capture and Flue Gas Capture Conditions. Chem. 2011, 45, 6670−6675.
Mater. 2019, 31, 5229−5237. (492) Shi, X.; Xiao, H.; Lackner, K. S.; Chen, X. Capture CO2 from
(472) Sujan, A. R.; Pang, S. H.; Zhu, G.; Jones, C. W.; Lively, R. P. Ambient Air Using Nanoconfined Ion Hydration. Angew. Chem. 2016,
Direct CO2 Capture from Air Using Poly(Ethylenimine)-Loaded 128, 4094−4097.
Polymer/Silica Fiber Sorbents. ACS Sustainable Chem. Eng. 2019, 7, (493) Shi, X.; Xiao, H.; Liao, X.; Armstrong, M.; Chen, X.; Lackner,
5264−5273. K. S. Humidity Effect on Ion Behaviors of Moisture-Driven CO2
(473) Wang, J.; Chen, H.; Zhou, H.; Liu, X.; Qiao, W.; Long, D.; Sorbents. J. Chem. Phys. 2018, 149, 164708.
Ling, L. Carbon Dioxide Capture Using Polyethylenimine-Loaded (494) He, H.; Zhong, M.; Konkolewicz, D.; Yacatto, K.; Rappold, T.;
Mesoporous Carbons. J. Environ. Sci. 2013, 25, 124−132. Sugar, G.; David, N. E.; Matyjaszewski, K. Carbon Black Function-
(474) Alesi, W. R.; Kitchin, J. R. Evaluation of a Primary Amine- alized with Hyperbranched Polymers: Synthesis, Characterization,
Functionalized Ion-Exchange Resin for CO2 Capture. Ind. Eng. Chem. and Application in Reversible CO2 Capture. J. Mater. Chem. A 2013,
1, 6810−6821.
Res. 2012, 51, 6907−6915.
(495) Song, J.; Liu, J.; Zhao, W.; Chen, Y.; Xiao, H.; Shi, X.; Liu, Y.;
(475) Leal, O.; Bolívar, C.; Ovalles, C.; García, J. J.; Espidel, Y.
Chen, X. Quaternized Chitosan/PVA Aerogels for Reversible CO2
Reversible Adsorption of Carbon Dioxide on Amine Surface-Bonded
Capture from Ambient Air. Ind. Eng. Chem. Res. 2018, 57, 4941−
Silica Gel. Inorg. Chim. Acta 1995, 240, 183−189.
4948.
(476) Tsuda, T.; Fujiwara, T. Polyethyleneimine and Macrocyclic
(496) Madden, D.; Curtin, T. Carbon Dioxide Capture with Amino-
Polyamine Silica Gels Acting as Carbon Dioxide Absorbents. J. Chem.
Functionalised Zeolite-β: A Temperature Programmed Desorption
Soc., Chem. Commun. 1992, 22, 1659−1661.
Study under Dry and Humid Conditions. Microporous Mesoporous
(477) Didas, S. A.; Sakwa-Novak, M. A.; Foo, G. S.; Sievers, C.;
Mater. 2016, 228, 310−317.
Jones, C. W. Effect of Amine Surface Coverage on the Co-Adsorption (497) Nguyen, T. H.; Kim, S.; Yoon, M.; Bae, T. H. Hierarchical
of CO2 and Water: Spectral Deconvolution of Adsorbed Species. J. Zeolites with Amine-Functionalized Mesoporous Domains for
Phys. Chem. Lett. 2014, 5, 4194−4200. Carbon Dioxide Capture. ChemSusChem 2016, 9, 455−461.
(478) Cho, M.; Park, J.; Yavuz, C. T.; Jung, Y. A Catalytic Role of (498) Wang, Y.; Du, T.; Qiu, Z.; Song, Y.; Che, S.; Fang, X. CO2
Surface Silanol Groups in CO2 Capture on the Amine-Anchored Adsorption on Polyethylenimine-Modified ZSM-5 Zeolite Synthe-
Silica Support. Phys. Chem. Chem. Phys. 2018, 20, 12149−12156. sized from Rice Husk Ash. Mater. Chem. Phys. 2018, 207, 105−113.
(479) Yu, J.; Chuang, S. S. C. The Role of Water in CO2 Capture by (499) Bezerra, D. P.; Silva, F. W. M. D.; Moura, P. A. S. D.; Sousa, A.
Amine. Ind. Eng. Chem. Res. 2017, 56, 6337−6347. G. S.; Vieira, R. S.; Rodriguez-Castellon, E.; Azevedo, D. C. S. CO2
(480) Zheng, F.; Tran, D. N.; Busche, B. J.; Fryxell, G. E.; Addleman, Adsorption in Amine-Grafted Zeolite 13X. Appl. Surf. Sci. 2014, 314,
R. S.; Zemanian, T. S.; Aardahl, C. L. Ethylenediamine-Modified SBA- 314−321.
15 as Regenerable CO2 Sorbent. Ind. Eng. Chem. Res. 2005, 44, (500) Wang, Y.; Du, T.; Song, Y.; Che, S.; Fang, X.; Zhou, L. Amine-
3099−3105. Functionalized Mesoporous ZSM-5 Zeolite Adsorbents for Carbon
(481) Chen, C. H.; Shimon, D.; Lee, J. J.; Mentink-Vigier, F.; Hung, Dioxide Capture. Solid State Sci. 2017, 73, 27−35.
I.; Sievers, C.; Jones, C. W.; Hayes, S. E. The “Missing” Bicarbonate in (501) Chatti, R.; Bansiwal, A. K.; Thote, J. A.; Kumar, V.; Jadhav, P.;
CO2 Chemisorption Reactions on Solid Amine Sorbents. J. Am. Lokhande, S. K.; Biniwale, R. B.; Labhsetwar, N. K.; Rayalu, S. S.
Chem. Soc. 2018, 140, 8648−8651. Amine Loaded Zeolites for Carbon Dioxide Capture: Amine Loading
(482) Sayari, A.; Belmabkhout, Y.; Da’na, E. CO2 Deactivation of and Adsorption Studies. Microporous Mesoporous Mater. 2009, 121,
Supported Amines: Does the Nature of Amine Matter? Langmuir 84−89.
2012, 28, 4241−4247. (502) Hammache, S.; Hoffman, J. S.; Gray, M. L.; Fauth, D. J.;
(483) Didas, S. A.; Zhu, R.; Brunelli, N. A.; Sholl, D. S.; Jones, C. W. Howard, B. H.; Pennline, H. W. Comprehensive Study of the Impact
Thermal, Oxidative and CO2 Induced Degradation of Primary of Steam on Polyethyleneimine on Silica for CO2 Capture. Energy
Amines Used for CO2 Capture: Effect of Alkyl Linker on Stability. J. Fuels 2013, 27, 6899−6905.
Phys. Chem. C 2014, 118, 12302−12311. (503) Langeroudi, E. G.; Kleitz, F.; Iliuta, M. C.; Larachi, F. Grafted
(484) Gebald, C.; Wurzbacher, J. A.; Tingaut, P.; Zimmermann, T.; Amine/CO2 Interactions in (Gas-)Liquid - Solid Adsorption/
Steinfeld, A. Amine-Based Nanofibrillated Cellulose as Adsorbent for Absorption Equilibria. J. Phys. Chem. C 2009, 113, 21866−21876.
CO2 Capture from Air. Environ. Sci. Technol. 2011, 45, 9101−9108. (504) Asenath Smith, E.; Chen, W. How to Prevent the Loss of
(485) Wurzbacher, J. A.; Gebald, C.; Piatkowski, N.; Steinfeld, A. Surface Functionality Derived from Aminosilanes. Langmuir 2008, 24,
Concurrent Separation of CO2 and H2O from Air by a Temperature- 12405−12409.
Vacuum Swing Adsorption/Desorption Cycle. Environ. Sci. Technol. (505) Etienne, M.; Goubert-Renaudin, S.; Rousselin, Y.; Marichal,
2012, 46, 9191−9198. C.; Denat, F.; Lebeau, B.; Walcarius, A. Multiarm Cyclam-Grafted

BJ https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

Mesoporous Silica: A Strategy to Improve the Chemical Stability of (524) Kulkarni, A. R.; Sholl, D. S. Analysis of Equilibrium-Based
Silica Materials Functionalized with Amine Ligands. Langmuir 2009, TSA Processes for Direct Capture of CO2 from Air. Ind. Eng. Chem.
25, 3137−3145. Res. 2012, 51, 8631−8645.
(506) Li, W.; Bollini, P.; Didas, S. A.; Choi, S.; Drese, J. H.; Jones, C. (525) Sinha, A.; Darunte, L. A.; Jones, C. W.; Realff, M. J.; Kawajiri,
W. Structural Changes of Silica Mesocellular Foam Supported Amine- Y. Systems Design and Economic Analysis of Direct Air Capture of
Functionalized CO2 Adsorbents Upon Exposure to Steam. ACS Appl. CO2 through Temperature Vacuum Swing Adsorption Using MIL-
Mater. Interfaces 2010, 2, 3363−3372. 101(Cr)-PEI-800 and Mmen-Mg2(Dobpdc) MOF Adsorbents. Ind.
(507) Li, W.; Choi, S.; Drese, J. H.; Hornbostel, M.; Krishnan, G.; Eng. Chem. Res. 2017, 56, 750−764.
Eisenberger, P. M.; Jones, C. W. Steam-Stripping for Regeneration of (526) Yamada, H.; Fujiki, J.; Chowdhury, F. A.; Yogo, K. Effect of
Supported Amine-Based CO2 Adsorbents. ChemSusChem 2010, 3, Isopropyl-Substituent Introduction into Tetraethylenepentamine-
899−903. Based Solid Sorbents for CO2 Capture. Fuel 2018, 214, 14−19.
(508) Fayaz, M.; Sayari, A. Long-Term Effect of Steam Exposure on (527) Yamada, H.; Chowdhury, F. A.; Fujiki, J.; Yogo, K.
CO2 Capture Performance of Amine-Grafted Silica. ACS Appl. Mater. Enhancement Mechanism of the CO2 Adsorption-Desorption
Interfaces 2017, 9, 43747−43754. Efficiency of Silica-Supported Tetraethylenepentamine by Chemical
(509) Guo, M.; Liang, S.; Liu, J.; Jin, J.; Mi, J. Epoxide- Modification of Amino Groups. ACS Sustainable Chem. Eng. 2019, 7,
Functionalization of Grafted Tetraethylenepentamine on the Frame- 9574−9581.
work of an Acrylate Copolymer as a CO2 Sorbent with Long Cycle (528) Fujiki, J.; Chowdhury, F. A.; Yamada, H.; Yogo, K. Highly
Stability. ACS Sustainable Chem. Eng. 2020, 8, 3853−3864. Efficient Post-Combustion CO2 Capture by Low-Temperature
(510) Begag, R.; Krutka, H.; Dong, W.; Mihalcik, D.; Rhine, W.; Steam-Aided Vacuum Swing Adsorption Using a Novel Polyamine-
Gould, G.; Baldic, J.; Nahass, P. Superhydrophobic Amine Function- Based Solid Sorbent. Chem. Eng. J. 2017, 307, 273−282.
alized Aerogels as Sorbents for CO2 Capture. Greenhouse Gases: Sci. (529) Choi, W.; Park, J.; Kim, C.; Choi, M. Structural Effects of
Technol. 2013, 3, 30−39. Amine Polymers on Stability and Energy Efficiency of Adsorbents in
(511) Rather, A. M.; Srikrishnarka, P.; Baidya, A.; Shome, A.; Post-Combustion CO2 capture. Chem. Eng. J. 2021, 408, 127289.
Pradeep, T.; Manna, U. Evaluating the Impact of Tailored Water (530) Tong, Z.; Ho, W. S. W. Facilitated Transport Membranes for
Wettability on Performance of CO2 Capture. ACS Appl. Energy Mater. CO2 Separation and Capture. Sep. Sci. Technol. 2017, 52, 156−167.
2020, 3, 10541−10549. (531) Li, Y.; Wang, S.; He, G.; Wu, H.; Pan, F.; Jiang, Z. Facilitated
(512) Baidya, A.; Yatheendran, A.; Ahuja, T.; Sudhakar, C.; Das, S. Transport of Small Molecules and Ions for Energy-Efficient
K.; Ras, R. H. A.; Pradeep, T. Waterborne Fluorine-Free Super- Membranes. Chem. Soc. Rev. 2015, 44, 103−118.
hydrophobic Surfaces Exhibiting Simultaneous CO2 and Humidity (532) Rea, R.; De Angelis, M. G.; Baschetti, M. G. Models for
Sorption. Adv. Mater. Interfaces 2019, 6, 1901013. Facilitated Transport Membranes: A Review. Membranes 2019, 9, 26.
(513) Abu-Zahra, M. R. M.; Schneiders, L. H. J.; Niederer, J. P. M.; (533) Zhao, S.; Feron, P. H. M.; Deng, L.; Favre, E.; Chabanon, E.;
Feron, P. H. M.; Versteeg, G. F. CO2 Capture from Power Plants: Yan, S.; Hou, J.; Chen, V.; Qi, H. Status and Progress of Membrane
Part I. A Parametric Study of the Technical Performance Based on Contactors in Post-Combustion Carbon Capture: A State-of-the-Art
Monoethanolamine. Int. J. Greenhouse Gas Control 2007, 1, 37−46. Review of New Developments. J. Membr. Sci. 2016, 511, 180−206.
(514) Yang, R. T. Gas Separation by Adsorption Processes, 1st ed.; (534) Norahim, N.; Yaisanga, P.; Faungnawakij, K.; Charinpanitkul,
Butterworth-Heinemann, 1986. T.; Klaysom, C. Recent Membrane Developments for CO2
(515) Ruthven, D. M. Principles of Adsorption and Adsorption Separation and Capture. Chem. Eng. Technol. 2018, 41, 211−223.
Processes; Wiley & Sons, 1984. (535) Baker, R. W.; Wijmans, J. G.; Huang, Y. Permeability,
(516) Pullumbi, P.; Brandani, F.; Brandani, S. Gas Separation by Permeance and Selectivity: A Preferred Way of Reporting
Adsorption: Technological Drivers and Opportunities for Improve- Pervaporation Performance Data. J. Membr. Sci. 2010, 348, 346−352.
ment. Curr. Opin. Chem. Eng. 2019, 24, 131−142. (536) Robeson, L. M. The Upper Bound Revisited. J. Membr. Sci.
(517) Zhang, W.; Liu, H.; Sun, Y.; Cakstins, J.; Sun, C.; Snape, C. E. 2008, 320, 390−400.
Parametric Study on the Regeneration Heat Requirement of an (537) Han, Y.; Ho, W. S. W. Polymeric Membranes for CO2
Amine-Based Solid Adsorbent Process for Post-Combustion Carbon Separation and Capture. J. Membr. Sci. 2021, 628, 119244.
Capture. Appl. Energy 2016, 168, 394−405. (538) Kueh, B.; Kapsi, M.; Veziri, C. M.; Athanasekou, C.; Pilatos,
(518) Zhang, W.; Sun, C.; Snape, C. E.; Irons, R.; Stebbing, S.; G.; Reddy, K. S. K.; Raj, A.; Karanikolos, G. N. Asphaltene-Derived
Alderson, T.; Fitzgerald, D.; Liu, H. Process Simulations of Post- Activated Carbon and Carbon Nanotube Membranes for CO2
Combustion CO2 Capture for Coal and Natural Gas-Fired Power Separation. Energy Fuels 2018, 32, 11718−11730.
Plants Using a Polyethyleneimine/Silica Adsorbent. Int. J. Greenhouse (539) Zhou, Y.; Yan, K.; Yang, Z.; Bauer, R. A.; Hong, N.; Verweij,
Gas Control 2017, 58, 276−289. H. Nanometer-Thick Supported Graphene Oxide Membrane for CO2
(519) Min, K.; Choi, W.; Kim, C.; Choi, M. Rational Design of the Capture. ACS Appl. Nano Mater. 2020, 3, 6654−6663.
Polymeric Amines in Solid Adsorbents for Postcombustion Carbon (540) Yoo, B. M.; Shin, J. E.; Lee, H. D.; Park, H. B. Graphene and
Dioxide Capture. ACS Appl. Mater. Interfaces 2018, 10, 23825−23833. Graphene Oxide Membranes for Gas Separation Applications. Curr.
(520) Li, K.; Jiang, J.; Yan, F.; Tian, S.; Chen, X. The Influence of Opin. Chem. Eng. 2017, 16, 39−47.
Polyethyleneimine Type and Molecular Weight on the CO2 Capture (541) Wang, B.; Dutta, P. K. Influence of Cross-Linking,
Performance of PEI-Nano Silica Adsorbents. Appl. Energy 2014, 136, Temperature, and Humidity on CO2/N2 Separation Performance
750−755. of PDMS Coated Zeolite Membranes Grown within a Porous
(521) Veneman, R.; Kamphuis, H.; Brilman, D. W. F. Post- Poly(Ether Sulfone) Polymer. Ind. Eng. Chem. Res. 2017, 56, 6065−
Combustion CO2 Capture Using Supported Amine Sorbents: A 6077.
Process Integration Study. Energy Procedia 2013, 37, 2100−2108. (542) Gu, X.; Dong, J.; Nenoff, T. M. Synthesis of Defect-Free FAU-
(522) Elfving, J.; Kauppinen, J.; Jegoroff, M.; Ruuskanen, V.; Type Zeolite Membranes and Separation for Dry and Moist CO2/
Järvinen, L.; Sainio, T. Experimental Comparison of Regeneration N2Mixtures. Ind. Eng. Chem. Res. 2005, 44, 937−944.
Methods for CO2 Concentration from Air Using Amine-Based (543) Liu, B.; Zhou, R.; Yogo, K.; Kita, H. Preparation of CHA
Adsorbent. Chem. Eng. J. 2021, 404, 126337. Zeolite (Chabazite) Crystals and Membranes without Organic
(523) Dijkstra, J. W.; Walspurger, S.; Elzinga, G. D.; Pieterse, J. A. Structural Directing Agents for CO2 Separation. J. Membr. Sci.
Z.; Boon, J.; Haije, W. G. Evaluation of Postcombustion CO2 Capture 2019, 573, 333−343.
by a Solid Sorbent with Process Modeling Using Experimental CO2 (544) Bernal, M. P.; Coronas, J.; Menéndez, M.; Santamaría, J.
and H2O Adsorption Characteristics. Ind. Eng. Chem. Res. 2018, 57, Separation of CO2/N2Mixtures Using MFI-Type Zeolite Mem-
1245−1261. branes. AIChE J. 2004, 50, 127−135.

BK https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

(545) Wijmans, J. G.; Baker, R. W. The Solution-Diffusion Model: A of Supported Ionic Liquid Membrane for CO2/N2. J. Membr. Sci.
Review. J. Membr. Sci. 1995, 107, 1−21. 2010, 350, 279−285.
(546) Liu, L.; Chakma, A.; Feng, X. Gas Permeation through Water- (565) Tomé, L. C.; Marrucho, I. M. Ionic Liquid-Based Materials: A
Swollen Hydrogel Membranes. J. Membr. Sci. 2008, 310, 66−75. Platform to Design Engineered CO2 Separation Membranes. Chem.
(547) Dai, Z.; Deng, J.; Aboukeila, H.; Yan, J.; Ansaloni, L.; Mineart, Soc. Rev. 2016, 45, 2785−2824.
K. P.; Giacinti Baschetti, M.; Spontak, R. J.; Deng, L. Highly CO2- (566) Li, Y.; Li, X.; Wu, H.; Xin, Q.; Wang, S.; Liu, Y.; Tian, Z.;
Permeable Membranes Derived from a Midblock-Sulfonated Multi- Zhou, T.; Jiang, Z.; Tian, H.; Cao, X.; Wang, B. Anionic Surfactant-
block Polymer after Submersion in Water. NPG Asia Mater. 2019, 11, Doped Pebax Membrane with Optimal Free Volume Characteristics
53. for Efficient CO2 Separation. J. Membr. Sci. 2015, 493, 460−469.
(548) Venturi, D.; Grupkovic, D.; Sisti, L.; Baschetti, M. G. Effect of (567) Lasseuguette, E.; Carta, M.; Brandani, S.; Ferrari, M. C. Effect
Humidity and Nanocellulose Content on Polyvinylamine-Nano- of Humidity and Flue Gas Impurities on CO2 Permeation of a
cellulose Hybrid Membranes for CO2 Capture. J. Membr. Sci. 2018, Polymer of Intrinsic Microporosity for Post-Combustion Capture. Int.
548, 263−274. J. Greenhouse Gas Control 2016, 50, 93−99.
(549) Zhao, Y.; Ho, W. S. W. CO2-Selective Membranes Containing (568) Cersosimo, M.; Brunetti, A.; Drioli, E.; Fiorino, F.; Dong, G.;
Sterically Hindered Amines for CO2/H2 Separation. Ind. Eng. Chem. Woo, K. T.; Lee, J.; Lee, Y. M.; Barbieri, G. Separation of CO2 from
Res. 2013, 52, 8774−8782. Humidified Ternary Gas Mixtures Using Thermally Rearranged
(550) Huang, J.; Zou, J.; Ho, W. S. W. Carbon Dioxide Capture Polymeric Membranes. J. Membr. Sci. 2015, 492, 257−262.
Using a CO2-Selective Facilitated Transport Membrane. Ind. Eng. (569) Qazi, S.; Gómez-Coma, L.; Albo, J.; Druon-Bocquet, S.;
Chem. Res. 2008, 47, 1261−1267. Irabien, A.; Sanchez-Marcano, J. CO2 Capture in a Hollow Fiber
(551) Belaissaoui, B.; Lasseuguette, E.; Janakiram, S.; Deng, L.; Membrane Contactor Coupled with Ionic Liquid: Influence of
Ferrari, M.-C. Analysis of CO2 Facilitation Transport Effect through a Membrane Wetting and Process Parameters. Sep. Purif. Technol. 2020,
Hybrid Poly(Allyl Amine) Membrane: Pathways for Further 233, 115986.
Improvement. Membranes 2020, 10, 367. (570) Yong, Z.; Mata, V.; Rodrigues, A. E. Adsorption of Carbon
(552) Kasahara, S.; Kamio, E.; Ishigami, T.; Matsuyama, H. Effect of Dioxide at High Temperature - A Review. Sep. Purif. Technol. 2002,
Water in Ionic Liquids on CO 2 Permeability in Amino Acid Ionic 26, 195−205.
Liquid-Based Facilitated Transport Membranes. J. Membr. Sci. 2012, (571) Lee, K. B.; Beaver, M. G.; Caram, H. S.; Sircar, S. Reversible
415−416, 168−175. Chemisorbents for Carbon Dioxide and Their Potential Applications.
(553) Suleman, M. S.; Lau, K. K.; Yeong, Y. F. Plasticization and Ind. Eng. Chem. Res. 2008, 47, 8048−8062.
Swelling in Polymeric Membranes in CO2 Removal from Natural (572) Hirano, S.; Shigemoto, N.; Yamada, S.; Hayashi, H. Cyclic
Gas. Chem. Eng. Technol. 2016, 39, 1604−1616. Fixed-Bed Operations over K2CO3-on-Carbon for the Recovery of
(554) Mahajan, R.; Koros, W. J. Factors Controlling Successful
Carbon Dioxide under Moist Conditions. Bull. Chem. Soc. Jpn. 1995,
Formation of Mixed-Matrix Gas Separation Materials. Ind. Eng. Chem.
68, 1030−1035.
Res. 2000, 39, 2692−2696.
(573) Lee, S. C.; Choi, B. Y.; Ryu, C. K.; Ahn, Y. S.; Lee, T. J.; Kim,
(555) Janakiram, S.; Ahmadi, M.; Dai, Z.; Ansaloni, L.; Deng, L.
J. C. The Effect of Water on the Activation and the CO2 Capture
Performance of Nanocomposite Membranes Containing 0D to 2D
Capacities of Alkali Metal-Based Sorbents. Korean J. Chem. Eng. 2006,
Nanofillers for CO2 Separation: A Review. Membranes 2018, 8, 24.
23, 374−379.
(556) Zhao, S.; Wang, Z.; Qiao, Z.; Wei, X.; Zhang, C.; Wang, J.;
(574) Sharonov, V. E.; Tyshchishchin, E. A.; Moroz, E. M.; Okunev,
Wang, S. Gas Separation Membrane with CO2-Facilitated Transport
Highway Constructed from Amino Carrier Containing Nanorods and A. G.; Aristov, Y. I. Sorption of CO2 from Humid Gases on Potassium
Macromolecules. J. Mater. Chem. A 2013, 1, 246−249. Carbonate Supported by Porous Matrix. Russ. J. Appl. Chem. 2001, 74,
(557) Liu, Y.; Peng, D.; He, G.; Wang, S.; Li, Y.; Wu, H.; Jiang, Z. 409−413.
Enhanced CO2 Permeability of Membranes by Incorporating (575) Feng, B.; An, H.; Tan, E. Screening of CO2 Adsorbing
Polyzwitterion@CNT Composite Particles into Polyimide Matrix. Materials for Zero Emission Power Generation Systems. Energy Fuels
ACS Appl. Mater. Interfaces 2014, 6, 13051−13060. 2007, 21, 426−434.
(558) Li, X.; Wang, M.; Wang, S.; Li, Y.; Jiang, Z.; Guo, R.; Wu, H.; (576) Gunathilake, C.; Jaroniec, M. Mesoporous Calcium Oxide-
Cao, X. Z.; Yang, J.; Wang, B. Constructing CO2 Transport Silica and Magnesium Oxide-Silica Composites for CO2 Capture at
Passageways in Matrimid® Membranes Using Nanohydrogels for Ambient and Elevated Temperatures. J. Mater. Chem. A 2016, 4,
Efficient Carbon Capture. J. Membr. Sci. 2015, 474, 156−166. 10914−10924.
(559) Liao, J.; Wang, Z.; Gao, C.; Wang, M.; Yan, K.; Xie, X.; Zhao, (577) Kumar, S.; Saxena, S. K. A Comparative Study of CO2
S.; Wang, J.; Wang, S. A High Performance PVAm-HT Membrane Sorption Properties for Different Oxides. Mater. Renew. Sustain. Energy
Containing High-Speed Facilitated Transport Channels for CO2 2014, 3, 30.
Separation. J. Mater. Chem. A 2015, 3, 16746−16761. (578) Back, M.; Bauer, M.; Stanjek, H.; Peiffer, S. Sequestration of
(560) Zhang, L.; Wang, R. Salting-out Effect on Facilitated CO2 After Reaction with Alkaline Earth Metal Oxides CaO and MgO.
Transport Membranes for CO2 Separation: From Fluoride Salt to Appl. Geochem. 2011, 26, 1097−1107.
Polyoxometalates. RSC Adv. 2012, 2, 9551−9554. (579) Okunev, a G.; Sharonov, V. E.; Aristov, Y. I.; Parmon, V. N.
(561) Li, Y.; Xin, Q.; Wu, H.; Guo, R.; Tian, Z.; Liu, Y.; Wang, S.; Sorption of Carbon Dioxide from Wet Gases by K2CO3-in-Porous
He, G.; Pan, F.; Jiang, Z. Efficient CO2 Capture by Humidified Matrix: Influence of the Matrix Nature. React. Kinet. Catal. Lett. 2000,
Polymer Electrolyte Membranes with Tunable Water State. Energy 71, 355−362.
Environ. Sci. 2014, 7, 1489−1499. (580) Seo, Y.; Jo, S. H.; Ryu, C. K.; Yi, C. K. Effects of Water Vapor
(562) Li, Y.; Xin, Q.; Wang, S.; Tian, Z.; Wu, H.; Liu, Y.; Jiang, Z. Pretreatment Time and Reaction Temperature on CO2 Capture
Trapping Bound Water within a Polymer Electrolyte Membrane of Characteristics of a Sodium-Based Solid Sorbent in a Bubbling
Calcium Phosphotungstate for Efficient CO2 Capture. Chem. Fluidized-Bed Reactor. Chemosphere 2007, 69, 712−718.
Commun. 2015, 51, 1901−1904. (581) Knuutila, H.; Hessen, E. T.; Kim, I.; Haug-Warberg, T.;
(563) Liao, J.; Wang, Z.; Gao, C.; Li, S.; Qiao, Z.; Wang, M.; Zhao, Svendsen, H. F. Vapor-Liquid Equilibrium in the Sodium Carbonate-
S.; Xie, X.; Wang, J.; Wang, S. Fabrication of High-Performance Sodium Bicarbonate-Water-CO2-System. Chem. Eng. Sci. 2010, 65,
Facilitated Transport Membranes for CO2 Separation. Chem. Sci. 2218−2226.
2014, 5, 2843−2849. (582) Liang, Y.; Harrison, D. P.; Gupta, R. P.; Green, D. A.;
(564) Zhao, W.; He, G.; Zhang, L.; Ju, J.; Dou, H.; Nie, F.; Li, C.; McMichael, W. J. Carbon Dioxide Capture Using Dry Sodium-Based
Liu, H. Effect of Water in Ionic Liquid on the Separation Performance Sorbents. Energy Fuels 2004, 18, 569−575.

BL https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

(583) Dong, W.; Chen, X.; Wu, Y.; Zhao, C.; Liang, C.; Liu, D. (603) Li, C. C.; Cheng, J. Y.; Liu, W. H.; Huang, C. M.; Hsu, H. W.;
Carbonation Characteristics of Dry Sodium-Based Sorbents for CO2 Lin, H. P. Enhancement in Cyclic Stability of the CO2 Adsorption
Capture. Energy Fuels 2012, 26, 6040−6046. Capacity of CaO-Based Sorbents by Hydration for the Calcium
(584) Luo, H.; Kanoh, H. Fundamentals in CO2 Capture of Looping Cycle. J. Taiwan Inst. Chem. Eng. 2014, 45, 227−232.
Na2CO3 under a Moist Condition. J. Energy Chem. 2017, 26, 972− (604) Guo, H.; Yan, S.; Zhao, Y.; Ma, X.; Wang, S. Influence of
983. Water Vapor on Cyclic CO2 Capture Performance in Both
(585) Cai, Y.; Wang, W.; Li, L.; Wang, Z.; Wang, S.; Ding, H.; Carbonation and Decarbonation Stages for Ca-Al Mixed Oxide.
Zhang, Z.; Sun, L.; Wang, W. Effective Capture of Carbon Dioxide Chem. Eng. J. 2019, 359, 542−551.
Using Hydrated Sodium Carbonate Powders. Materials 2018, 11, 183. (605) Manovic, V.; Anthony, E. J. Carbonation of CaO-Based
(586) Hayashi, H.; Taniuchi, J.; Furuyashiki, N.; Sugiyama, S.; Sorbents Enhanced by Steam Addition. Ind. Eng. Chem. Res. 2010, 49,
Hirano, S.; Shigemoto, N.; Nonaka, T. Efficient Recovery of Carbon 9105−9110.
Dioxide from Flue Gases of Coal-Fired Power Plants by Cyclic Fixed- (606) Chen, L.; Qian, N. The Effects of Water Vapor and Coal Ash
Bed Operations over K2CO3-on-Carbon. Ind. Eng. Chem. Res. 1998, on the Carbonation Behavior of CaO-Sorbent Supported by γ-Al2O3
37, 185−191. for CO2 Capture. Fuel Process. Technol. 2018, 177, 200−209.
(587) Shigemoto, N.; Yanagihara, T.; Sugiyama, S.; Hayashi, H. (607) Zhang, W.; Ma, X.; Li, Y.; Zhao, J.; Wang, Z. A Study of the
Material Balance and Energy Consumption for CO2 Recovery from Synergistic Effects of Mn/Steam on CO2 Capture Performance of
Moist Flue Gas Employing K2CO3-on-Activated Carbon and Its CaO by Experiment and DFT Calculation. Greenhouse Gases: Sci.
Evaluation for Practical Adaptation. Energy Fuels 2006, 20, 721−726. Technol. 2019, 9, 409−423.
(588) Lee, S. C.; Chae, H. J.; Choi, B. Y.; Jung, S. Y.; Ryu, C. Y.; (608) Chou, Y. C.; Cheng, J. Y.; Liu, W. H.; Hsu, H. W. Effects of
Park, J. J.; Baek, J. I.; Ryu, C. K.; Kim, J. C. The Effect of Relative Steam Addition during Calcination on Carbonation Behavior in a
Humidity on CO2 Capture Capacity of Potassium-Based Sorbents. Calcination/Carbonation Loop. Chem. Eng. Technol. 2018, 41, 1921−
Korean J. Chem. Eng. 2011, 28, 480−486. 1927.
(589) Luo, H.; Chioyama, H.; Thürmer, S.; Ohba, T.; Kanoh, H. (609) Erans, M.; Nabavi, S. A.; Manović, V. Carbonation of Lime-
Kinetics and Structural Changes in CO2 Capture of K2CO3 under a Based Materials under Ambient Conditions for Direct Air Capture. J.
Moist Condition. Energy Fuels 2015, 29, 4472−4478. Cleaner Prod. 2020, 242, 118330.
(590) Jayakumar, A.; Gomez, A.; Mahinpey, N. Post-Combustion (610) Ć wik, A.; Casanova, I.; Rausis, K.; Koukouzas, N.; Zarȩbska,
CO2 Capture Using Solid K2CO3: Discovering the Carbonation K. Carbonation of High-Calcium Fly Ashes and Its Potential for
Reaction Mechanism. Appl. Energy 2016, 179, 531−543. Carbon Dioxide Removal in Coal Fired Power Plants. J. Cleaner Prod.
(591) Gomez, A.; Jayakumar, A.; Mahinpey, N. Experimental 2018, 202, 1026−1034.
Verification of the Reaction Mechanism of Solid K2CO3 during (611) Yang, P.; Duan, L.; Tang, H.; Cai, T.; Sun, Z. Explaining
Steam-Enhanced Carbonation of CaO Based on First Principles.
Postcombustion CO2 Capture. Ind. Eng. Chem. Res. 2016, 55, 11022−
Greenhouse Gases: Sci. Technol. 2018, 8, 1110−1123.
11028.
(612) Ma, X.; Li, Y.; Zhang, W.; Wang, Z.; Zhao, J. DFT Study of
(592) Wu, S. F.; Beum, T. H.; Yang, J. I.; Kim, J. N. Properties of
CO2 Adsorption across a CaO/Ca12Al14O33 Sorbent in the
Ca-Base CO2 Sorbent Using Ca(OH)2 as Precursor. Ind. Eng. Chem.
Presence of H2O under Calcium Looping Conditions. Chem. Eng. J.
Res. 2007, 46, 7896−7899.
2019, 370, 10−18.
(593) Ding, Y.; Alpay, E. Equilibria and Kinetics of CO2 Adsorption
(613) Wang, N.; Feng, Y.; Guo, X.; Van Duin, A. C. T. Insights into
on Hydrotalcite Adsorbent. Chem. Eng. Sci. 2000, 55, 3461−3474.
the Role of H2O in the Carbonation of CaO Nanoparticle with CO2.
(594) Fagerlund, J.; Highfield, J.; Zevenhoven, R. Kinetics Studies
J. Phys. Chem. C 2018, 122, 21401−21410.
on Wet and Dry Gas-Solid Carbonation of MgO and Mg(OH)2 for (614) He, D.; Ou, Z.; Qin, C.; Deng, T.; Yin, J.; Pu, G.
CO2 Sequestration. RSC Adv. 2012, 2, 10380−10393. Understanding the Catalytic Acceleration Effect of Steam on
(595) Salaudeen, S. A.; Acharya, B.; Dutta, A. CaO-Based CO2 CaCO3 Decomposition by Density Function Theory. Chem. Eng. J.
Sorbents: A Review on Screening, Enhancement, Cyclic Stability, 2020, 379, 122348.
Regeneration and Kinetics Modelling. J. CO2 Util. 2018, 23, 179− (615) Xu, Z.; Wu, Y.; Cai, T.; Chen, X.; Ma, J.; Liu, D.
199. Understanding the CO2 Sorption Mechanisms of the MgO-Doped
(596) Sun, H.; Wu, C.; Shen, B.; Zhang, X.; Zhang, Y.; Huang, J. Na-Based Sorbent at Low Temperatures. Greenhouse Gases: Sci.
Progress in the Development and Application of CaO-Based Technol. 2019, 9, 672−686.
Adsorbents for CO2 Capturea Review. Mater. Today Sustain. (616) Hu, Y.; Guo, Y.; Sun, J.; Li, H.; Liu, W. Progress in MgO
2018, 1−2, 1−27. Sorbents for Cyclic CO2 Capture: A Comprehensive Review. J. Mater.
(597) Blamey, J.; Anthony, E. J.; Wang, J.; Fennell, P. S. The Chem. A 2019, 7, 20103−20120.
Calcium Looping Cycle for Large-Scale CO2 Capture. Prog. Energy (617) Zarghami, S.; Hassanzadeh, A.; Arastoopour, H.; Abbasian, J.
Combust. Sci. 2010, 36, 260−279. Effect of Steam on the Reactivity of MgO-Based Sorbents in
(598) Dean, C. C.; Blamey, J.; Florin, N. H.; Al-Jeboori, M. J.; Precombustion CO2 Capture Processes. Ind. Eng. Chem. Res. 2015,
Fennell, P. S. The Calcium Looping Cycle for CO2 Capture from 54, 8860−8866.
Power Generation, Cement Manufacture and Hydrogen Production. (618) Jiao, X.; Li, H.; Li, L.; Xiao, F.; Zhao, N.; Wei, W. Synthesis
Chem. Eng. Res. Des. 2011, 89, 836−855. and CO2 Capture Properties of Mesoporous MgAl(O) Sorbent. RSC
(599) Hilz, J.; Helbig, M.; Haaf, M.; Daikeler, A.; Ströhle, J.; Epple, Adv. 2014, 4, 47012−47020.
B. Long-Term Pilot Testing of the Carbonate Looping Process in (619) Ram Reddy, M. K.; Xu, Z. P.; Lu, G. Q.; Diniz da Costa, J. C.
1MWth Scale. Fuel 2017, 210, 892−899. D. Layered Double Hydroxides for CO2 Capture: Structure Evolution
(600) Perejón, A.; Romeo, L. M.; Lara, Y.; Lisbona, P.; Martínez, A.; and Regeneration. Ind. Eng. Chem. Res. 2006, 45, 7504−7509.
Valverde, J. M. The Calcium-Looping Technology for CO2 Capture: (620) Ram Reddy, M. K.; Xu, Z. P.; Diniz Da Costa, J. C. Influence
On the Important Roles of Energy Integration and Sorbent Behavior. of Water on High-Temperature CO2 Capture Using Layered Double
Appl. Energy 2016, 162, 787−807. Hydroxide Derivatives. Ind. Eng. Chem. Res. 2008, 47, 2630−2635.
(601) Zhang, L.; Zhang, B.; Yang, Z.; Guo, M. The Role of Water on (621) Hanif, A.; Sun, M.; Shang, S.; Tian, Y.; Yip, A. C. K.; Ok, Y. S.;
the Performance of Calcium Oxide-Based Sorbents for Carbon Yu, I. K. M.; Tsang, D. C. W.; Gu, Q.; Shang, J. Exfoliated Ni-Al LDH
Dioxide Capture: A Review. Energy Technol. 2015, 3, 10−19. 2D Nanosheets for Intermediate Temperature CO2 Capture. J.
(602) Coppola, A.; Palladino, L.; Montagnaro, F.; Scala, F.; Salatino, Hazard. Mater. 2019, 374, 365−371.
P. Reactivation by Steam Hydration of Sorbents for Fluidized-Bed (622) Coenen, K.; Gallucci, F.; Cobden, P.; van Dijk, E.; Hensen, E.;
Calcium Looping. Energy Fuels 2015, 29, 4436−4446. van Sint Annaland, M. Chemisorption Working Capacity and Kinetics

BM https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX
Chemical Reviews pubs.acs.org/CR Review

of CO2 and H2O of Hydrotalcite-Based Adsorbents for Sorption-


Enhanced Water-Gas-Shift Applications. Chem. Eng. J. 2016, 293, 9−
23.
(623) Coenen, K.; Gallucci, F.; Pio, G.; Cobden, P.; van Dijk, E.;
Hensen, E.; van Sint Annaland, M. On the Influence of Steam on the
CO2 Chemisorption Capacity of a Hydrotalcite-Based Adsorbent for
SEWGS Applications. Chem. Eng. J. 2017, 314, 554−569.
(624) Coenen, K.; Gallucci, F.; Hensen, E.; Van Sint Annaland, M.
CO2 and H2O Chemisorption Mechanism on Different Potassium-
Promoted Sorbents for SEWGS Processes. J. CO2 Util. 2018, 25,
180−193.
(625) Coenen, K.; Gallucci, F.; Mezari, B.; Hensen, E.; van Sint
Annaland, M. An In-Situ IR Study on the Adsorption of CO2 and
H2O on Hydrotalcites. J. CO2 Util. 2018, 24, 228−239.
(626) Coenen, K.; Gallucci, F.; Hensen, E.; van Sint Annaland, M.
Kinetic Model for Adsorption and Desorption of H2O and CO2 on
Hydrotalcite-Based Adsorbents. Chem. Eng. J. 2019, 355, 520−531.
(627) Baltrusaitis, J.; Schuttlefield, J. D.; Zeitler, E.; Jensen, J. H.;
Grassian, V. H. Surface Reactions of Carbon Dioxide at the Adsorbed
Water-Oxide Interface. J. Phys. Chem. C 2007, 111, 14870−14880.
(628) Baltrusaitis, J.; Schuttlefield, J.; Zeitler, E.; Grassian, V. H.
Carbon Dioxide Adsorption on Oxide Nanoparticle Surfaces. Chem.
Eng. J. 2011, 170, 471−481.
(629) Omodolor, I. S.; Otor, H. O.; Andonegui, J. A.; Allen, B. J.;
Alba-Rubio, A. C. Dual-Function Materials for CO2Capture and
Conversion: A Review. Ind. Eng. Chem. Res. 2020, 59, 17612−17631.
(630) Duyar, M. S.; Wang, S.; Arellano-Treviño, M. A.; Farrauto, R.
J. CO2 Utilization with a Novel Dual Function Material (DFM) for
Capture and Catalytic Conversion to Synthetic Natural Gas: An
Update. J. CO2 Util. 2016, 15, 65−71.
(631) Al-Mamoori, A.; Rownaghi, A. A.; Rezaei, F. Combined
Capture and Utilization of CO2 for Syngas Production over Dual-
Function Materials. ACS Sustainable Chem. Eng. 2018, 6, 13551−
13561.
(632) Talapaneni, S. N.; Buyukcakir, O.; Je, S. H.; Srinivasan, S.;
Seo, Y.; Polychronopoulou, K.; Coskun, A. Nanoporous Polymers
Incorporating Sterically Confined N-Heterocyclic Carbenes for
Simultaneous CO2 Capture and Conversion at Ambient Pressure.
Chem. Mater. 2015, 27, 6818−6826.
(633) Luo, R.; Chen, M.; Liu, X.; Xu, W.; Li, J.; Liu, B.; Fang, Y.
Recent Advances in CO2capture and Simultaneous Conversion into
Cyclic Carbonates over Porous Organic Polymers Having Accessible
Metal Sites. J. Mater. Chem. A 2020, 8, 18408−18424.
(634) Arellano-Treviño, M. A.; He, Z.; Libby, M. C.; Farrauto, R. J.
Catalysts and Adsorbents for CO2 Capture and Conversion with Dual
Function Materials: Limitations of Ni-Containing DFMs for Flue Gas
Applications. J. CO2 Util. 2019, 31, 143−151.
(635) Arellano-Treviño, M. A.; Kanani, N.; Jeong-Potter, C. W.;
Farrauto, R. J. Bimetallic Catalysts for CO2 Capture and Hydro-
genation at Simulated Flue Gas Conditions. Chem. Eng. J. 2019, 375,
121953.
(636) Wang, J.-Q.; Kong, D.-L.; Chen, J.-Y.; Cai, F.; He, L.-N.
Synthesis of Cyclic Carbonates from Epoxides and Carbon Dioxide
over Silica-Supported Quaternary Ammonium Salts under Super-
critical Conditions. J. Mol. Catal. A: Chem. 2006, 249, 143−148.
(637) Kolle, J. M.; Sayari, A. Substrate Dependence on the Fixation
of CO2 to Cyclic Carbonates over Reusable Porous Hybrid Solids. J.
CO2 Util. 2018, 26, 564−574.

BN https://doi.org/10.1021/acs.chemrev.0c00762
Chem. Rev. XXXX, XXX, XXX−XXX

You might also like