You are on page 1of 11

Accepted Article

Title: Oxidatively-Stable Linear Poly(propylenimine)-Containing


Adsorbents for CO2 Capture from Ultra-Dilute Streams

Authors: Simon H Pang, Ryan P Lively, and Christopher W Jones

This manuscript has been accepted after peer review and appears as an
Accepted Article online prior to editing, proofing, and formal publication
of the final Version of Record (VoR). This work is currently citable by
using the Digital Object Identifier (DOI) given below. The VoR will be
published online in Early View as soon as possible and may be different
to this Accepted Article as a result of editing. Readers should obtain
the VoR from the journal website shown below when it is published
to ensure accuracy of information. The authors are responsible for the
content of this Accepted Article.

To be cited as: ChemSusChem 10.1002/cssc.201800438

Link to VoR: http://dx.doi.org/10.1002/cssc.201800438

A Journal of

www.chemsuschem.org
ChemSusChem 10.1002/cssc.201800438

FULL PAPER
Oxidatively-Stable Linear Poly(propylenimine)-Containing
Adsorbents for CO2 Capture from Ultra-Dilute Streams
Simon H. Pang,[a] Ryan P. Lively,[a] and Christopher W. Jones*[a]

Abstract: Aminopolymer-based solid sorbents have been widely or leaching in the presence of steam.[33,34]
investigated for CO2 capture from dilute streams such as flue gas or For direct air capture, oxidative degradation is a major
ambient air. However, the oxidative stability of the widely-studied concern, due to the abundance of O2 in the air. The oxidation of
aminopolymer, poly(ethylenimine) (PEI), is limited, causing it to lose amines is known to occur through a radical-mediated process,

Accepted Manuscript
its CO2 capture capacity after exposure to oxygen at elevated and though is typically catalyzed by transition metal ions, can
temperatures. Here we demonstrate the use of linear occur in the absence of such ions at elevated temperatures.[35,36]
poly(propylenimine) (PPI), synthesized via a simple cationic ring- The susceptibility of various types of amines to oxidation has
opening polymerization, as a more oxidatively-stable alternative to been studied for molecularly grafted, isolated aminosilanes, with
PEI with high CO2 capacity and amine efficiency. The performance the major finding that primary and tertiary amines are relatively
of linear PPI/SBA-15 composites is investigated over a range of CO2 resistant to oxidative degradation, whereas secondary amines
capture conditions (CO2 partial pressure, adsorption temperature) to and secondary-amine-containing molecules degrade via
examine the tradeoff between adsorption capacity and sorption site formation of imines or amides.[26,27]
accessibility, which may be expected to be more limited in linear In the case of polymeric amines, randomly-branched
polymers relative to the prototypical hyperbranched PEI. Linear poly(ethylenimine) (PEI), which is the best studied
PPI/SBA-15 composites are more efficient at CO2 capture and retain aminopolymer for CO2 capture, is known to readily degrade in
65-83% of their CO2 capacity after exposure to a harsh oxidative oxidizing environments.[37–39] Ahmadalinezhad and Sayari
treatment, compared to 20-40% retention for linear PEI. Additionally, investigated the structures created during oxidation of PEI via
we demonstrate long-term stability of linear PPI sorbents over 50 detailed NMR studies. They suggest that aminopolymers that
adsorption/desorption cycles with no loss in performance. Combined only contain primary amines should be the most resistant to
with other strategies for improving oxidative stability and adsorption degradation, followed by aminopolymers that contain only
kinetics, linear PPI may play a role as a component of stable, solid secondary amines, with highly branched aminopolymers that
adsorbents in commercial applications for CO2 capture. contain primary, secondary, and tertiary amines being the least
stable.[39] This is supported by studies utilizing poly(allyl amine)
(PAA)-based sorbents, which are more resistant to oxidative
degradation than PEI-based sorbents.[38,40]
Introduction However, one of the downsides of aminopolymers that
only contain primary amines is that they will rapidly deactivate in
Research into materials for direct capture of CO2 from dilute CO2-rich atmospheres at elevated temperature via formation of
streams such as flue gas or ambient air has focused on solid urea. A study by Sayari et al. demonstrates that branched PEI
adsorbents due to the potential for reducing the energy required and PAA deactivate via urea formation more rapidly than linear
for material regeneration compared to the benchmark PEI.[31] They propose that the primary amines on branched PEI
technology of aqueous solutions of amines.[1–4] In particular, and PAA react with CO2 and dehydrate to form isocyanate
amines supported in a variety of solids (oxides,[5–15], species, which readily react with another amine to form ureas
polymers,[16–18] metal-organic frameworks,[19–21] for example) that bridge two molecules. However, the secondary amines
have been studied due to their high heats of CO2 adsorption[22– present on linear PEI form cyclic ureas between two neighboring
24]
and the consequently higher capture and separation amines through dehydration of an ammonium carbamate
efficiencies[25] achievable. However, these high heats of intermediate. This pathway is hypothesized to be slower than
adsorption require high temperatures to regenerate the material, the isocyanate pathway, possibly due to secondary amines
and these harsh conditions can cause material degradation. being more sterically hindered.[37,41] Therefore, though linear PEI
Amine-containing sorbents are known to degrade via a is susceptible to oxidative degradation, linear aminopolymers
number of different mechanisms that are accelerated at the high may be desirable for preventing deactivation through urea
temperatures required for sorbent regeneration: oxidative formation.
degradation,[26–29] formation of urea in CO2-rich atmospheres,[30– Several groups have developed strategies for improving
32]
and loss of aminopolymer at high temperature via evaporation oxidative stability of PEI-based sorbents. Srikanth and Chuang
have reported that blending of small aminopolymers with
poly(ethylene glycol) (PEG) improves the oxidative stability of
[a] Dr. S. H. Pang, Prof. R. P. Lively, Prof. C. W. Jones the sorbents by hydrogen bonding with amines.[29,42] This is
School of Chemical & Biomolecular Engineering
proposed to help disperse the aminopolymer and prevent O2
Georgia Institute of Technology
311 Ferst Drive, Atlanta, GA 30332 (USA) from accessing the amine sites, which retards oxidation.
E-mail: cjones@gatech.edu Following this, Choi et al. have functionalized branched PEI with
Supporting information and the ORCID identification numbers for the
1,2-epoxybutane, which incorporates hydrogen bonding groups
authors of this article can be found under directly onto the aminopolymer, to simultaneously suppress urea
https://dx.doi.org/10.1002/cssc.xxxxxxxxx. formation and oxidative degradation of amines by reducing the

This article is protected by copyright. All rights reserved.


ChemSusChem 10.1002/cssc.201800438

FULL PAPER
fraction of primary amines.[43] The same group has extended aqueous NH4OH to remove any residual inorganic salt,
their work to incorporating metal chelating groups on the silica generating linear PPI (Scheme 1). The monomer to initiator ratio
support to help remove ppm levels of residual oxidation- was varied to control the resulting molecular weight of the
catalyzing metal ions that remain from the polymer synthesis.[44] polymer, and the resulting number-averaged molecular weight
Our previous work has shown that increasing the spacing was estimated via 1H NMR by end group analysis (Figures S1-
of the amines in the backbone, by switching from a S4). The ordering of molecular weights was also confirmed via
poly(ethylenimine) to a poly(propylenimine) (PPI) drastically aqueous GPC (Figure S5). As a starting point, linear PPI at
improves the stability against oxidative degradation for small approximately 1000 g/mol was synthesized to mimic the number
molecule aminopolymers.[45] This stability improvement was of polymer repeat units contained in the traditionally-studied
observed for PPI molecules that both contained and did not randomly-branched PEI (800 g/mol). At all molecular weights
contain secondary amines. This observation is in agreement studied, linear PPI was recovered as an off-white solid at room

Accepted Manuscript
with previous work with small molecule amines that suggested temperature; for comparison, the traditionally-studied randomly-
that ethylenimines in oxidizing conditions can undergo chain branched PEI is a viscous liquid at room temperature, and
rearrangement reactions to form six-membered rings, which are commercially-available linear PEI is a solid at room temperature.
easily oxidized to piperazinones. In contrast, propylenimines Composites of linear PPI and mesoporous silica SBA-15
would form either four- or eight-membered rings, which are were prepared by wet impregnation of a methanolic PPI solution;
relatively unfavorable, and so do not deactivate as readily.[46,47] the loading of polymer in the composite was controlled by
However, for all of these small molecule studies, evaporative varying the polymer-to-silica ratio, and the actual loading was
loss of amine during high temperature sorbent regeneration is a confirmed via TGA (Figure S6). For all of the polymer molecular
major concern from both an economic and environmental weights, the textural properties and distribution of polymer within
perspective. the mesopores of SBA-15 was investigated by nitrogen
Based on these previous investigations, in this study we physisorption and the resulting isotherm was analyzed via the
sought to develop an oxidatively-stable aminopolymer-based BJH method to calculate a pore size distribution (Figure 1,
sorbent for capture of CO2 from dilute streams based on Figure S7, Table S1). The nitrogen physisorption isotherm for
poly(propylenimine). To both mitigate urea formation and reduce the bare SBA-15 material is characterized by sharp uptake in the
evaporative loss of aminopolymer during high temperature region of P/P0 < 0.01, corresponding to adsorption within the
regeneration, we synthesized linear PPI at a range of molecular intra-wall micropores of the silica. This is followed by multilayer
weights to probe material stability during CO2 capture from dilute adsorption of N2 and finally capillary condensation within the
conditions. mesopores. These adsorption steps correspond to a population
of micropores with diameter below 2 nm, and a larger population
of mesopores with diameter centered around 7.0 nm.
Results and Discussion At roughly 10-20 wt% linear PPI loading, the micropores
are essentially inaccessible to nitrogen, suggesting that at these
Synthesis of linear PPI was carried out via cationic ring- low polymer loadings, a monolayer of polymer is formed on the
opening polymerization of 2-methyl-5,6-dihydro-4H-1,3-oxazine mesopores walls, which either fills or blocks access to the
(MeOZI) with methyl p-toluenesulfonate (MeOTs) as the initiator. micropores. This is accompanied by a decrease in the average
The resulting poly(MeOZI) was hydrolyzed under acidic pore diameter to approximately 6.5 nm. Further increase in
conditions to remove the protecting groups, and then carefully polymer loading is accompanied by a decrease in the pore
neutralized via addition of NaOH and washed copiously with volume with a gradual decrease in the pore diameter until the

Scheme 1. Synthesis of linear poly(propylenimine) via cationic ring-opening polymerization of 2-methyl-5,6-dihydro-4H-1,3-oxazine.

This article is protected by copyright. All rights reserved.


ChemSusChem 10.1002/cssc.201800438

FULL PAPER
pores are nearly completely filled at about 50 wt%. This behavior As shown in Figure 2, the amine efficiency and CO2
is similar to that observed for composites of randomly-branched capture capacity of all linear PPI sorbents increased as polymer
PEI and SBA-15, which was also studied via small-angle loading in the SBA-15 was increased. In the low polymer loading
neutron scattering, suggesting that, similarly to PEI, linear PPI regime (< 10 mmol N/g SiO2), all molecular weights of polymer
first forms a conformal coating on the mesopores walls, followed appeared to perform similarly for the adsorption conditions used
by forming aggregates of polymer that grow along the axis of the in this study. However, at high polymer loading (> 10 mmol N/g
pore.[48] Notably, there was no apparent difference in the pore- SiO2), the amine efficiency of the polymers appeared to
filling behavior of linear PPI as the molecular weight was varied decrease slightly as the molecular weight of the linear PPI
by almost two orders of magnitude, suggesting that the increased, reaching about 0.16 mmol CO2/mmol N for the lower
impregnation method was able to achieve similarly distributed molecular weight polymers and only 0.13 mmol CO2/mmol N for
linear PPI within the pores of SBA-15 in all cases. the higher molecular weight polymers. In agreement with this

Accepted Manuscript
trend, the corresponding small molecule linear propylenimine,
tripropylenetetramine (TPTA), achieved higher amine
efficiencies than any of the linear PPIs, though it should be
noted that per molecule, TPTA has two primary amines, which
are expected to be the most efficient for CO2 capture.[49]

Figure 2. Amine efficiencies of linear PPI and PEI impregnated in SBA-15


(adsorption in 400 ppm CO2/N2 at 35 °C, 18 h) over a range of amine loading.
Open symbols indicate samples exposed to oxidative treatment (21% O2/N2 at
110 °C, 24 h) prior to CO2 adsorption. Data for TPTA are replotted from ref. [45].
Error bars indicate the standard deviation of three separate adsorption
experiments on different powder samples.

Similar adsorption experiments were performed for


commercially-available linear PEI at two different molecular
Figure 1. (a) Nitrogen physisorption isotherms at 77 K and (b) resulting pore
weights. As observed previously, the amine efficiency for the
size distributions calculated by the BJH method for linear PPI (1000
g/mol)/SBA-15 composites as a function of polymer loading. Open symbols PEI composites was lower than that for the PPI composites,[45]
indicate the desorption branch of the isotherm. Isotherms and pore size likely due to the decrease in the basicity (pKa) of the amines with
distributions for additional linear PPI molecular weights can be found in Figure decreasing length of the hydrocarbon spacer.[50] Under similar
S7.
adsorption conditions, the amine efficiency of linear PEI/silica
composites can be increased to about 0.10 mmol CO2/mmol N
by using a higher pore volume silica,[51] though this value is still
The CO2 adsorption performance of these supported linear
lower than that achievable by linear PPI composites as shown
PPI composites under direct air capture conditions was studied
here.
gravimetrically in a TGA by a temperature swing process.
The best amine efficiency and corresponding CO2 capture
Adsorption was performed at 35 °C with simulated air (400 ppm
capacity for linear PPI/SBA-15 reported here (0.16 mmol
CO2 in N2), and sorbent regeneration was performed at 110 °C
CO2/mmol N and 1.25 mmol CO2/g composite, respectively, for
in UHP N2. Adsorption experiments were performed in triplicate
the sample made with the 1000 g/mol linear PPI) is better than
on different powder samples and the results are plotted in terms
that reported for randomly-branched PEI/SBA-15 composites
of amine efficiency in Figure 2, and CO2 capacity in Figure S8.
(1.05 mmol CO2/g composite) under similar conditions and
amine loadings.[52] In addition to the increased basicity of the

This article is protected by copyright. All rights reserved.


ChemSusChem 10.1002/cssc.201800438

FULL PAPER
amines on linear PPI, we hypothesize that the increase in amine and CH2 asymmetric and symmetric stretching (2940, 2873,
efficiency and CO2 capture capacity for the linear PPI/SBA-15 2815 cm-1). These same stretches are observed for dendritic
materials relative to the randomly-branched PEI/SBA-15 PPI.[53] The stretches observed in the lower wavenumber region
materials is in part due to the absence of tertiary amines, which are attributable to formation of ammonium carbamate pairs
are present on the branched PEI and are essentially inactive for (a/symmetric NH2+ at 1640 and 1480 cm-1, and a/symmetric
CO2 capture under the conditions used here.[49] These values fall COO- at 1540, 1420, and 1380 cm-1),[54,55] due to adsorption of
short of the highest reported capacities of about 2.35 mmol atmospheric CO2. Below this region, the IR spectra for the
CO2/g composite for PEI impregnated in high pore volume composites are dominated by modes from the silica support, but
fumed silicas.[12,14] To test the effect of increased pore volume of spectra of the neat polymer display bands predominantly at
the support, a mesocellular silica foam (MCF) was impregnated 1315 and 1110 cm-1 attributable to CH2 wagging and C-N
with linear PPI (1000 g/mol) at an amine loading of 14 mmol N/g stretching modes as well (Figure S9).

Accepted Manuscript
SiO2. However, this composite performed no better than linear
PPI/SBA-15 composites (0.17 mmol CO2/mmol N, 1.33 mmol
CO2/g composite), suggesting that the morphology of the
polymer layer was more important than the morphology of the
support in the linear PPI case.
The amine efficiencies of linear PPI/SBA-15 composites
after an accelerated oxidative treatment are also shown in
Figure 2. The as-prepared composite materials at high polymer
loading were exposed to a stream of ultrazero grade air at
110 °C for 24 h to simulate potential exposure to air at the
adsorbent regeneration temperature during the temperature
swing process. In all cases, the loading of polymer changed
almost negligibly (< 5% decrease in amine loading) and all of the
samples retained the capability to capture CO2. In particular for
air capture, the 700 g/mol and 1000 g/mol samples retained
about 65% of their amine efficiency, whereas the higher
molecular weight 6700 g/mol and 36000 g/mol samples retained
over 83%. The retention of amine efficiency for sorbents
prepared from linear PPI after exposure to oxidative treatment at
higher concentrations of CO2 (Figure S10f) is similar to that for
other state-of-the-art materials. In the flue gas capture regime
(10-15% CO2), the 1000 g/mol linear PPI composite retains
about 75% of its amine efficiency, comparable to epoxide-
functionalized PEI reported by Choi et al.[43] (retention of
between 56 and 81% amine efficiency depending on extent of
functionalization). In the 1 bar regime, the 1000 g/mol linear PPI
samples retains about 85% of its amine efficiency, which is
better than blends of tetraethylenepentamine (TEPA) and
poly(ethylene glycol) (PEG) reported by Srikanth and Chuang[29]
(retention of between 37 and 56% amine efficiency depending
Figure 3. FTIR spectra of linear PPI and PEI impregnated in SBA-15 before
on the ratio of TEPA to PEG). (black) and after exposure to oxidative treatment (21% O2/N2 at 110 °C, 24 h)
In comparison, the low molecular weight linear PEI sample (red). Spectra were acquired under a flow of UHP N2 and normalized to the Si-
lost over 80% of its amine efficiency and the high molecular O-Si framework bending mode at 460 cm-1.
weight linear PEI lost between 60 and 75% of its amine
efficiency, depending on the polymer loading. These greater
losses in amine efficiency are expected based on our previous After exposure to the oxidizing conditions, the bands for
results with small molecule aminopolymers.[45] The differences in the linear PPI/SBA-15 composites largely remained the same,
loss in CO2 capture capacity and amine efficiency for the though the intensity of the N-H and CH2 stretches decreased
different aminopolymer molecular weights will be discussed slightly, potentially indicating some small loss of organic content.
within the context of mass transport limitations in more detail However, for the linear PEI/SBA-15 composites, an intense
below. band appeared at 1670 cm-1, attributable to formation of C=N or
Fourier transform infrared (FTIR) spectra for the linear C=O groups, consistent with oxidation of
aminopolymer/SBA-15 composites before and after exposure to poly(ethylenimines).[29,37,38,43] The appearance of this band was
the oxidative treatment are shown in Figure 3. Additionally, accompanied by a significant reduction in intensity of the N-H
spectra for the unsupported, neat linear PPI were acquired in and CH2 stretches, as expected. These features can be
attenuated total reflectance (ATR) mode for comparison (Figure correlated with the loss in CO2 capture capability shown in
S9). Prior to the treatment, the major bands present above 2000 Figure 2.
cm-1 correspond to the linear PPI backbone, evident for both the To examine the applicability of linear PPI sorbent materials
composites and the neat polymer – N-H stretching (3240 cm-1) across a range of CO2 adsorption conditions, CO2 adsorption
isotherms were obtained at a range of temperatures (Figure 4a,

This article is protected by copyright. All rights reserved.


ChemSusChem 10.1002/cssc.201800438

FULL PAPER
Figure S10). For all of the linear PPI sorbents, the temperature- resistance for diffusion of CO2 within the bulk of the
dependent behavior depended strongly on the CO2 partial aminopolymer and an inability for CO2 to access all of the amine
pressure regime under consideration. For partial pressures sites. As the temperature is increased, the flexibility and mobility
relevant to direct air capture (PCO2 = 0.3 Torr), the CO2 capacity of the polymer chains increases, as does the diffusivity of CO2.
and amine efficiency decreased as the adsorption temperature In turn, more amine sites become accessible for adsorption, and
increased, consistent with an exothermic chemisorption process thus, the CO2 capacity of the material increases. At 1 bar CO2,
that is thermodynamically-limited. This behavior has been this has been observed for PEI/MCM-41,[5] PEI/KIT-6,[8]
observed for randomly-branched PEI/SBA-15 composites[56] as PEI/SBA-15 and PAA/SBA-15.[40]
well as for amine-grafted sorbents constructed from A similar phenomenon occured at partial pressures
aminosilanes grafted on mesocellular silica foam.[49] relevant to flue gas capture (PCO2 = 76 Torr) – initially the CO2
capacity and amine efficiency increased as temperature

Accepted Manuscript
increased. However, the 1000 g/mol linear PPI/SBA-15
composite reached a maximum amine efficiency at around 55 °C,
after which the amine efficiency decreased. This behavior is
consistent with transitioning from a mass transport-limited
regime where increasing temperature allows access to more
sites for CO2 adsorption to a regime where accessibility of amine
sites is no longer a limiting factor, and the adsorption behavior
can follow the expected thermodynamic response. The
temperature at which this transition occurred appeared to be
dependent on the molecular weight of the polymer; the higher
molecular weight polymers reached a maximum in amine
efficiency around 65 °C (Figure S10). This is likely related to the
decreasing mobility of the polymer chains as molecular weight
increased.
It is expected that this transition from a transport-limited
regime to a thermodynamically-limited regime would occur in
both the air capture and 1 bar CO2 regimes as well, though the
temperature at which the maximum efficiency occurs appears to
be outside of the range of temperatures studied here. The
pressure-dependence of the optimum adsorption temperature
can be rationalized by considering the mechanism for CO2
capture by amines in dry conditions. The formation of
ammonium carbamate ion pairs requires two amines per CO2
captured; this creates a crosslink that can occur within a single
polymer chain, or in between nearby polymer chains. These
crosslinks decrease the effective diffusivity of CO2 by limiting the
mobility of the polymer chains. The density of crosslinks is
directly related to the amount of CO2 captured; in low partial
pressures of CO2, fewer crosslinks are formed and so higher
temperatures are not required to overcome barriers to mass
transport. However, at high partial pressure of CO2, a large
fraction of the polymer is crosslinked, and so higher temperature
Figure 4. (a) CO2 adsorption isotherms, plotted in terms of amine efficiency,
is required to increase the effective diffusivity of CO2 to the point
obtained for the 50 wt% linear PPI (1000 g/mol)/SBA-15 sample from 25 °C to
75 °C. Dashed lines indicate relevant partial pressures for direct air capture where all of the amine sites can become accessible.
and flue gas capture. (b) CO2 adsorption heating/cooling isobar for the 50 wt% To demonstrate the difference in site accessibility, a CO2
linear PPI (1000 g/mol)/SBA-15 sample obtained in 10% CO2/N2. adsorption isobar was collected with 10% CO2/N2 to examine the
temperature-dependent hysteresis (Figure 4b, Figure S11). As
expected from the isotherm studies, the CO2 capacity and amine
In contrast, at pressures near 1 bar, the CO2 capacity and efficiency increased as the adsorption temperature was
amine efficiency increased as temperature increased, reaching a increased from 25 °C to 55 °C, after which the efficiency
value of 2.5 mmol CO2/g composite (Figure S10) at 75 °C for the decreased. Upon cooling the sample stepwise from 95 °C to
1000 g/mol linear PPI with an amine efficiency of 0.32 mmol 25 °C, the CO2 capacity and amine efficiency consistently
CO2/mmol N. This CO2 capacity is slightly lower than the highest increased, as would be the expected thermodynamic response.
value reported for randomly branched PEI/SBA-15 composites This suggests that CO2 was adsorbed on sites that were
at similar polymer loading (2.7 mmol CO2/g composite) at the previously inaccessible when adsorption started from low
same temperature and partial pressure, though the amine temperature, and therefore the material was more easily able to
efficiency reported here is higher due to the lower amine density approach the thermodynamically-limited CO2 capacity of 2.4
of PPI compared to PEI.[40] mmol CO2/g composite at 25 °C (Figure S11). In principle, this
This behavior is not consistent with a thermodynamically-
limited process and has been ascribed to high mass transport

This article is protected by copyright. All rights reserved.


ChemSusChem 10.1002/cssc.201800438

FULL PAPER
capacity could also be reached at low temperature by extending adsorption are somewhat larger than those reported for
the time for adsorption. adsorption of CO2 on other branched PEI/SiO2 composites
The difference in CO2 capacity and amine efficiency at (typically 50-80 kJ/mol).[56,59,60] However, it should be noted that
high polymer loading for the different molecular weights of linear this method of estimating the heats of adsorption tends to
PPI observed in Figure 2 is related to these differences in site underestimate the true value obtained via microcalorimetry,[23,61]
accessibility. There was a clear decrease in the rate of CO2 and thus the values for linear PPI reported here can be taken as
adsorption as molecular weight of the linear PPI was increased, a lower bound. Nonetheless, CO2 appears to adsorb more
as shown in Figure 5, again related to the decreasing mobility of strongly on PPI than PEI, possibly due to the increased basicity
the polymer chains. We hypothesize that this is also the reason of propylenimines compared to ethylenimines.
for the improved resistance to oxidative degradation for the Finally, to investigate the long-term stability of linear PPI
higher molecular weight linear PPI – by decreasing the mobility sorbents, rapid temperature-swing adsorption/desorption

Accepted Manuscript
of the polymer chains, oxygen is not as easily able to penetrate cycles[62,63] for direct air capture were performed. The CO2
the polymer bulk to initiate degradation reactions. This adsorption/desorption cycles are shown in Figure 6a, and the
phenomenon has also been observed in comparison of working amine efficiency for each cycle is shown in Figure 6b.
branched and linear PEI/silica composites -- similarly to linear Due to the rapid decrease in the rate of adsorption seen in
PPI, linear PEI is a solid at room temperature, and thus adsorbs Figure 5, adsorption was limited to one hour, rather than the 18
CO2 more slowly than the more mobile branched PEI, resulting hours required to achieve pseudo-equilibrium. Despite this, 60%
in differences in both the rate of adsorption and CO2 capacity.[57] of the pseudo-equilibrium value was reached with fully reversible
Interestingly, site accessibility and CO2 adsorption kinetics can adsorption; the linear PPI/SBA-15 sorbent was able to maintain
be improved slightly in a larger pore volume silica such as MCF stable operation with a working capacity of 0.76 mmol CO2/g
(Figure S12). However, as mentioned above, it appears that the composite. This suggests that that the sorbent did not deactivate
polymer morphology may be more important than the support due to formation of strongly-bound species like urea, which are
morphology when the polymers are solids under the adsorption less likely to form from linear aminopolymers,[31,37] or
conditions. deteriorate/lose polymer during the high temperature desorption.
Indeed, analysis of the organic content of the composite after
the cycling test indicated that less than 0.3% of the polymer
mass had been lost, suggesting that polymer molecular weights
as low as 1000 g/mol were sufficient for preventing volatilization
of the polymer. The slowly decreasing baseline observed in
Figure 6a, then, is likely due to slow loss of small amounts of
water from the hydrophilic aminopolymer/silica composite, which
is not expected to impact performance significantly.

Figure 5. Dynamic CO2 uptake profiles for high polymer loading samples,
normalized against their pseudoequilibrium value reported in Figure 2.
Adsorption in 400 ppm CO2/N2 at 35 °C.

The CO2 adsorption isotherms in Figure 4a can be


modeled using a dual-site temperature-dependent Toth isotherm
(Supporting Figure S13a), which has been used previously to
model the adsorption of CO2 on supported amines.[58] Using this
model, the zero-coverage heat of adsorption is estimated to be
105 kJ/mol for CO2 adsorption on the 50 wt% linear PPI (1000
g/mol)/SBA-15 sample. The isosteric heats of adsorption as a
Figure 6. (a) CO2 adsorption/desorption working capacity profile for the 50
function of amount of CO2 adsorbed can also be estimated using
wt% linear PPI (1000 g/mol)/SBA-15 sample over 50 consecutive temperature
the isotherm data and the Clausius-Clapeyron equation swing adsorption cycles. Adsorption was performed in 400 ppm CO2/N2 at
(Supporting Figure S13b). Values around 100 kJ/mol are 35 °C for 1 h; desorption was performed in UHP N2 at 110 °C for 10 min. (b)
estimated for loadings of CO2 up to about 1.5 mmol/g, beyond Amine efficiency as a function of the temperature swing adsorption cycle.
which the heat of adsorption decreases. These heats of

This article is protected by copyright. All rights reserved.


ChemSusChem 10.1002/cssc.201800438

FULL PAPER
Conclusions 400 mL deionized water and the precipitate was filtered and washed
copiously with deionized water. The filtered precipitate was dried for 12 h
in an oven at 75 °C and then calcined according to the following program:
This work demonstrates the potential for using linear
heat to 200 °C at 1.2 °C/min, hold at 200 °C 1 h, heat to 550 °C at
poly(propylenimine) (PPI) in CO2 sorbent materials for capture of
1.2 °C/min, hold at 550 °C for 12 h, cool to room temperature. The
CO2 from dilute sources such as ambient air or flue gas. At high resulting white powder was stored in ambient lab conditions.
polymer loading, lower molecular weight species of linear PPI
were more efficient at CO2 capture due to more rapid adsorption Synthesis of 2-methyl-5,6-dihydro-4H-1,3-oxazine
kinetics compared with higher molecular weight species. Under
similar adsorption conditions and sorbent characteristics, linear Synthesis of 2-methyl-5,6-dihydro-4H-1,3-oxazine (MeOZI) was adapted
PPI performed better than the traditionally-studied randomly- from published procedures.[65,66] Anhydrous ZnCl2 (5.5 g, 0.04 mol) was
branched PEI, though it is expected that improvements in the dispersed in anhydrous MeCN (115 mL, 2.2 mol). To this solution, 3-

Accepted Manuscript
adsorption kinetics and performance can be made by changing aminopropanol (150.2 g, 2.0 mol) was added dropwise at room
the textural properties of the support material. Importantly, linear temperature under stirring. The mixture was then purged with dry argon
PPI largely retained its CO2 capture capacity after exposure to and heated to 95 °C with stirring for 48 h, allowing NH3 gas to evolve and
vent from the system. The reaction mixture was cooled to room
an accelerated oxidative treatment, unlike related linear PEI
temperature and excess MeCN and NH3 were removed by rotary
materials. CO2 adsorption isotherm and isobar hysteresis evaporation at 30 °C. MeOZI was isolated by distillation under vacuum
experiments suggest that the significant mass transport (~400 mTorr, 65 °C), cooling the distillate with a dry ice/isopropyl alcohol
limitations imposed by high polymer loading can be overcome by mixture, to afford a clear, colorless liquid which was stored in ambient lab
starting adsorption at elevated temperature to increase the conditions.
1
adsorption site accessibility by improving polymer chain mobility H NMR (400 MHz, CDCl3, δ): 4.144 (t, J = 5.4 Hz, 2H), 3.338 (t, J = 5.8
and the effective diffusivity of CO2, and then reducing the Hz, 2H), 1.879+1.851 (s+q, J = 5.7 Hz, 5H).
13
temperature to increase the thermodynamically-limited capacity. C NMR (100 MHz, CDCl3, δ): 157.85, 65.00, 42.41, 22.06, 21.78.
The results presented here suggest that improvement in
Synthesis of linear poly(MeOZI)
adsorption kinetics will allow PPI-based sorbents to be a
promising, oxidatively-stable, alternative to the traditionally-
Cationic polymerization of MeOZI was adapted from published
studied PEI-based sorbents for use in CO2 capture from dilute
procedures.[67,68] Prior to polymerization, MeOZI and MeCN were dried
sources. over CaH2 by stirring at 65 °C. Dried MeOZI and MeCN was collected
into a Straus flask by distillation under vacuum (~400 mTorr, 65 °C and
30 °C, respectively), cooling the distillate with a dry ice/isopropyl alcohol
Experimental Section mixture. MeOZI and MeCN were stored under dry argon and transferred
into a dry box. 10 g dried MeOZI in the dry box was transferred into a
Chemicals, Materials and Chemical Characterization round bottom flask, to which 50 mL dried MeCN was added. An amount
of methyl p-toluenesulfonate (MeOTs) was added, based on the desired
final molecular weight of the polymer. The round bottom flask was sealed
Chemicals for syntheses reported below were obtained from Sigma-
with a rubber septum and heated to 90 °C with stirring to initiate
Aldrich and Alfa Aesar and used directly without further purification. Ultra-
polymerization. Polymerization was allowed to proceed for 48 h, after
high purity N2, ultra-high purity He, ultra-zero grade air, research grade
which 2.0 equivalents of 1,3-diaminopropane (based on MeOTs) was
CO2, 400 ppm CO2/N2, and 10% CO2/N2 were obtained from Airgas.
1H NMR spectra were recorded on a Bruker AVIII-400 spectrometer (400
added to quench the reaction. MeCN and excess 1,3-diaminopropane
were removed by rotary evaporation at 50 °C to afford a viscous, yellow
MHz). Chemical shifts are reported in parts per million (ppm) either
liquid as crude poly(MeOZI).
referenced to 0.0 ppm for tetramethylsilane or to the center line of a
quintet at 3.312 ppm for methanol-d4. The following abbreviations are
used to explain multiplicities: s = singlet, t = triplet, q = quintet. 13C NMR Synthesis of linear poly(propylenimine)
spectra were recorded on a Bruker AVIII-400 spectrometer (100 MHz),
and were fully decoupled by broad band proton decoupling. Chemical Linear poly(propylenimine) (PPI) was synthesized by acidic hydrolysis of
shifts are reported in ppm referenced to either the center line of a triplet poly(MeOZI), adapted from a previously published procedure.[69] 200 mL
at 77.2 ppm for chloroform-d or to the center line of a septet at 49.15 of 5 M aqueous HCl was added to the crude poly(MeOZI) residue and
ppm for methanol-d4. stirred at 100 °C for 48 h. After cooling to room temperature, the residual
Gel permeation chromatography was performed on a Shimadzu HPLC polymer was concentrated by rotary evaporation at 60 °C, which caused
system with refractive index detector (RID-10A) and TSKgel G3000- the polymer to precipitate from solution. The concentrated polymer
G6000-PWxl-CP Guard, G3000-PWxl-CP, and G5000-PWxl-CP columns. solution was cooled to 0 °C in an ice bath and carefully brought to pH 13
Samples were prepared and flowed in an eluent phase of 0.4 M acetic using concentrated aqueous NaOH, which caused to polymer to first
acid with 0.3 M NaNO3 at 1 mL/min. become fully dissolved, and then precipitate as an off-white powder.
Linear PPI was recovered by centrifugation (14000 rpm, 10 min) and
Synthesis of SBA-15 washed copiously with 10% NH4OH to remove any residual inorganic salt.
The final polymer was dissolved in MeOH and dried under high vacuum
to afford an off-white solid (~75% yield based on 10 g monomer). The
SBA-15 was synthesized according to our previously published
absence of residual inorganic salt was verified by TGA. The number-
procedure.[64] 24.0 g of Pluronic P-123 block copolymer
averaged molecular weight was estimated by 1H NMR by end group
((EO)20(PO)70(EO)20) was dissolved in 636 g of deionized water and 120
analysis, Figures S1-S4 (400 MHz, methanol-d4, δ): 3.625 (t, -
mL of 12.1 M HCl. The components were stirred vigorously for 3 h, until
CH2CH2CH2NH2, 2H), 2.612 (t, -CH2CH2CH2NH-, 4nH), 2.361 (s, CH3-,
everything dissolved. 46.6 g of tetraethyl orthosilicate (TEOS) was added
3H), 1.709 (q, -CH2CH2CH2NH-, 2nH).
dropwise to the mixture and stirred at 40 °C for 20 h, during which time a
white precipitate formed. The solution was then heated to 100 °C and
held for 24 h in the absence of stirring. The reaction was quenched with Linear PPI/SBA-15 composite preparation

This article is protected by copyright. All rights reserved.


ChemSusChem 10.1002/cssc.201800438

FULL PAPER
SBA-15 was impregnated with linear PPI by wet impregnation. SBA-15 were degassed under dynamic vacuum on the instrument (< 5 mTorr) at
was dried overnight under vacuum (< 20 mTorr) at 110 °C. The desired 110 °C for 12 h. The free space measurement was performed prior to
amount of linear PPI was dissolved in 15 mL methanol and added to the analysis followed by a series of 180 sec leak tests to ensure that the
desired amount of SBA-15. The mixture was allowed to stir at room helium from the free space measurement had been completely
temperature for at least 6 h. Methanol was removed by rotary evacuated from the sample. An equilibration interval of 60 sec was used
evaporation at room temperature. The resulting powder was dried for pressures less than 20 Torr and 30 sec for all other pressures. For a
overnight under vacuum (< 20 mTorr) at room temperature. The resulting given composite, all isotherms were collected using the same sample to
dried powder composites were stored in ambient lab conditions. An avoid introducing weighing error, and the order in which the isotherms
identical procedure was used to prepare linear PEI/SBA-15 composites, were collected was randomized to avoid potential temperature-hysteresis.
except these powder composites were stored in dry Ar to prevent For the 1000 g/mol linear PPI sample, the 35 °C isotherm measurement
unwanted oxidation. was repeated after all other isotherms had been collected, to ensure that
no degradation or loss of aminopolymer had occurred (Figure S10f).
Physical Characterization Adsorption isotherms for N2 and O2 at 35 °C were obtained similarly,

Accepted Manuscript
demonstrating that these gases adsorb only through weak physisorption,
whereas CO2 adsorbs via strong chemisorption (Figure S14). Therefore
Nitrogen adsorption isotherms were obtained on a MicrotracBEL
the presence of N2 or O2 during the mixed gas CO2 adsorption process
BELSORP-max at 77 K. Prior to analysis, the samples were degassed
(such as in the gravimetric uptake experiments) is not expected to impact
under vacuum on a BELPREP-vac II below 10-2 kPa for 12 h at 110 °C.
the amount of CO2 adsorbed.
The free space measurement was performed after analysis. Surface area
was analyzed by applying BET theory in the range of P/P0 from 0.05 to
0.2, total pore volume was calculated based on the total amount of N2 Accelerated Oxidative Treatment
adsorbed at P/P0 of 0.95, and pore size distributions were calculated by
the BJH method with Broekhoff-de Boer thickness curve, implemented in Linear aminopolymer/SBA-15 composites were exposed to an
the MicroActive software package by Micromeritics. accelerated oxidative treatment in a TA Instruments Q500 TGA. The as-
Organic content of the samples was estimated on a Netzsch STA409PG prepared composites were heated to 110 °C under a flow of ultra-zero
TGA. Mass loss from 125 to 900 °C under a flow of nitrogen-diluted air grade air and held for 24 h.
was recorded and normalized by the residual mass at 900 °C.
Fourier transform infrared (FTIR) spectra were recorded in transmission
mode in a flow of N2 on a Thermo Nicolet iS10 with an MCT detector.
Approximately 5 mg of the sample was diluted by 110 mg KBr and
Acknowledgements
pelletized for analysis.
Funding for this work was provided as a part of UNCAGE-ME,
Gravimetric CO2 Adsorption an Energy Frontier Research Center funded by the U.S.
Department of Energy, Office of Science, Basic Energy
Pseudo-equilibrium CO2 adsorption capacities were measured Sciences, under award no. DE-SC0012577, and by Global
gravimetrically on a TA Instruments Q500 TGA. The samples were Thermostat, LLC. We also thank Dr. Michele Sarazen for help
pretreated by heating to 110 °C at a ramp rate of 10 °C/min under a flow with the GPC experiments.
of N2 and held for 2 h. The samples were cooled to 35 °C and
equilibrated at this analysis temperature for 1 h. Subsequently, the gas
flow was switched to a premixed gas containing 400 ppm CO2/N2 for 18 h Keywords: adsorption • amines • polymers • carbon storage •
and the rate of mass change was less than 5 × 10-6 %/min. The mass mesoporous materials
gain was recorded and normalized by the dry mass of the sample.
Pseudo-equilibrium capacities were obtained in triplicate and the [1] S. Choi, J. H. Drese, C. W. Jones, ChemSusChem 2009, 2, 796–854.
reported error bars indicate the standard deviation of these [2] Q. Wang, J. Luo, Z. Zhong, A. Borgna, Energy Environ. Sci. 2011, 4,
measurements. 42–55.
The CO2 adsorption isobar was obtained similarly, but the sample was [3] A. Goeppert, M. Czaun, G. K. Surya Prakash, G. A. Olah, Energy
not treated at high temperature in between adsorption points. Following Environ. Sci. 2012, 5, 7833.
pretreatment and equilibration at 25 °C, the gas flow was switched to a [4] E. S. Sanz-Pérez, C. R. Murdock, S. A. Didas, C. W. Jones, Chem.
premixed gas containing 10% CO2/N2 for at least 6 h, or until the rate of Rev. 2016, 116, 11840–11876.
mass change was less than 5 × 10-6 %/min. Following this, the [5] X. Xu, C. Song, J. M. Andrésen, B. G. Miller, A. W. Scaroni, Energy
temperature was increased to 35 °C at a rate of 10 °C/min in the flow of Fuels 2002, 16, 1463–1469.
premixed gas and again held for at least 6 h, or until the rate of mass [6] X. Xu, C. Song, J. M. Andrésen, B. G. Miller, A. W. Scaroni,
change was less than 5 × 10-6 %/min. This procedure was repeated up to Microporous Mesoporous Mater. 2003, 62, 29–45.
95 °C to collect the heating curve for the isobar. Following the 95 °C [7] R. S. Franchi, P. J. E. Harlick, A. Sayari, Ind. Eng. Chem. Res. 2005,
measurement, the temperature was decreased to 85 °C, and the same 44, 8007–8013.
procedure was repeated down to 25 °C to collect the cooling curve for [8] W.-J. Son, J.-S. Choi, W.-S. Ahn, Microporous Mesoporous Mater.
the isobar. 2008, 113, 31–40.
Cyclic adsorption/desorption experiments were performed similarly, with [9] Y. Belmabkhout, R. Serna-Guerrero, A. Sayari, Chem. Eng. Sci. 2010,
sorbent regeneration in between each adsorption cycle. The sample was 65, 3695–3698.
initially pretreated at 110 °C for 2 h, and then equilibrated at 35 °C for 1 h. [10] A. Goeppert, M. Czaun, R. B. May, G. K. S. Prakash, G. A. Olah, S. R.
Cycles of adsorption for 1 h and desorption for 10 min were then Narayanan, J. Am. Chem. Soc. 2011, 133, 20164–20167.
performed, allowing for 10 min of equilibration time at the adsorption [11] Z. Bacsik, N. Ahlsten, A. Ziadi, G. Zhao, A. E. Garcia-Bennett, B.
temperature to determine the starting mass for adsorption for each cycle. Martín-Matute, N. Hedin, Langmuir 2011, 27, 11118–11128.
[12] S. Choi, M. L. Gray, C. W. Jones, ChemSusChem 2011, 4, 628–635.
Volumetric CO2 Adsorption [13] S. Choi, J. H. Drese, P. M. Eisenberger, C. W. Jones, Environ. Sci.
Technol. 2011, 45, 2420–2427.
Pseudo-equilibrium CO2 adsorption isotherms were measured [14] A. Goeppert, H. Zhang, M. Czaun, R. B. May, G. K. S. Prakash, G. A.
volumetrically on a Micromeritics ASAP 2020. Prior to analysis, samples Olah, S. R. Narayanan, ChemSusChem 2014, 7, 1386–1397.

This article is protected by copyright. All rights reserved.


ChemSusChem 10.1002/cssc.201800438

FULL PAPER
[15] A. Sayari, Q. Liu, P. Mishra, ChemSusChem 2016, 5, 1–9. [44] K. Min, W. Choi, C. Kim, M. Choi, Nat. Commun. 2018, 9, 726.
[16] S. Satyapal, T. Filburn, J. Trela, J. Strange, Energy Fuels 2001, 15, [45] S. H. Pang, L.-C. Lee, M. A. Sakwa-Novak, R. P. Lively, C. W. Jones, J.
250–255. Am. Chem. Soc. 2017, 139, 3627–3630.
[17] S. H. Pang, M. L. Jue, J. Leisen, C. W. Jones, R. P. Lively, ACS Macro [46] H. Lepaumier, D. Picq, P.-L. Carrette, Ind. Eng. Chem. Res. 2009, 48,
Lett. 2015, 4, 1415–1419. 9068–9075.
[18] C. Xu, Z. Bacsik, N. Hedin, J. Mater. Chem. A 2015, 3, 16229–16234. [47] H. Lepaumier, S. Martin, D. Picq, B. Delfort, P.-L. Carrette, Ind. Eng.
[19] A. R. Millward, O. M. Yaghi, J. Am. Chem. Soc. 2005, 127, 17998– Chem. Res. 2010, 49, 4553–4560.
17999. [48] A. Holewinski, M. A. Sakwa-Novak, C. W. Jones, J. Am. Chem. Soc.
[20] K. Sumida, D. L. Rogow, J. A. Mason, T. M. McDonald, E. D. Bloch, Z. 2015, 137, 11749–11759.
R. Herm, T.-H. Bae, J. R. Long, Chem. Rev. 2012, 112, 724–781. [49] S. A. Didas, A. R. Kulkarni, D. S. Sholl, C. W. Jones, ChemSusChem
[21] T. M. McDonald, W. R. Lee, J. A. Mason, B. M. Wiers, C. S. Hong, J. R. 2012, 5, 2058–2064.
Long, J. Am. Chem. Soc. 2012, 134, 7056–7065. [50] M. Borkovec, G. J. M. Koper, J. Phys. Chem. 1994, 98, 6038–6045.
[22] C. Knöfel, C. Martin, V. Hornebecq, P. L. Llewellyn, J. Phys. Chem. C [51] W. Chaikittisilp, R. Khunsupat, T. T. Chen, C. W. Jones, Ind. Eng.

Accepted Manuscript
2009, 113, 21726–21734. Chem. Res. 2011, 50, 14203–14210.
[23] M. A. Alkhabbaz, P. Bollini, G. S. Foo, C. Sievers, C. W. Jones, J. Am. [52] W. Chaikittisilp, H. J. Kim, C. W. Jones, Energy Fuels 2011, 25, 5528–
Chem. Soc. 2014, 136, 13170–13173. 5537.
[24] C.-J. Yoo, L.-C. Lee, C. W. Jones, Langmuir 2015, 31, 13350–13360. [53] S. S. Abkenar, R. M. A. Malek, Cellulose 2012, 19, 1701–1714.
[25] R. P. Lively, M. J. Realff, AIChE J. 2016, 62, 3699–3705. [54] M. W. Hahn, M. Steib, A. Jentys, J. A. Lercher, J. Phys. Chem. C 2015,
[26] P. Bollini, S. Choi, J. H. Drese, C. W. Jones, Energy Fuels 2011, 25, 119, 4126–4135.
2416–2425. [55] G. S. Foo, J. J. Lee, C.-H. Chen, S. E. Hayes, C. Sievers, C. W. Jones,
[27] A. Heydari-Gorji, Y. Belmabkhout, A. Sayari, Microporous Mesoporous ChemSusChem 2016, 1–12.
Mater. 2011, 145, 146–149. [56] N. Gargiulo, A. Peluso, P. Aprea, F. Pepe, D. Caputo, J. Chem. Eng.
[28] G. Calleja, R. Sanz, A. Arencibia, E. S. Sanz-Pérez, Top. Catal. 2011, Data 2014, 59, 896–902.
54, 135–145. [57] K. Li, J. Jiang, F. Yan, S. Tian, X. Chen, Appl. Energy 2014, 136, 750–
[29] C. S. Srikanth, S. S. C. Chuang, ChemSusChem 2012, 5, 1435–1442. 755.
[30] T. C. Drage, A. Arenillas, K. M. Smith, C. E. Snape, Microporous [58] R. Serna-Guerrero, A. Sayari, Chem. Eng. J. 2010, 161, 182–190.
Mesoporous Mater. 2008, 116, 504–512. [59] X. Wang, C. Song, Catal. Today 2012, 194, 44–52.
[31] A. Sayari, A. Heydari-Gorji, Y. Yang, J. Am. Chem. Soc. 2012, 134, [60] H. Zhang, A. Goeppert, G. K. S. Prakash, G. Olah, RSC Adv. 2015, 5,
13834–13842. 52550–52562.
[32] S. A. Didas, R. Zhu, N. A. Brunelli, D. S. Sholl, C. W. Jones, J. Phys. [61] M. E. Potter, S. H. Pang, C. W. Jones, Langmuir 2017, 33, 117–124.
Chem. C 2014, 118, 12302–12311. [62] R. P. Lively, R. R. Chance, B. T. Kelley, H. W. Deckman, J. H. Drese,
[33] S. Hammache, J. S. Hoffman, M. L. Gray, D. J. Fauth, B. H. Howard, H. C. W. Jones, W. J. Koros, Ind. Eng. Chem. Res. 2009, 48, 7314–7324.
W. Pennline, Energy Fuels 2013, 27, 6899–6905. [63] Y. Fan, R. P. Lively, Y. Labreche, F. Rezaei, W. J. Koros, C. W. Jones,
[34] M. A. Sakwa-Novak, C. W. Jones, ACS Appl. Mater. Interfaces 2014, 6, Int. J. Greenh. Gas Control 2014, 21, 61–71.
9245–9255. [64] E. G. Moschetta, M. A. Sakwa-Novak, J. L. Greenfield, C. W. Jones,
[35] S. Chi, G. T. Rochelle, Ind. Eng. Chem. Res. 2002, 41, 4178–4186. Langmuir 2015, 31, 2218–2227.
[36] A. K. Voice, G. T. Rochelle, Ind. Eng. Chem. Res. 2014, 53, 16222– [65] J. Lee, K. Lee, H. Kim, Bull. Korean Chem. Soc. 1996, 17, 115--116.
16228. [66] M. M. Bloksma, R. M. Paulus, H. P. C. van Kuringen, F. van der
[37] A. Heydari-Gorji, A. Sayari, Ind. Eng. Chem. Res. 2012, 51, 6887–6894. Woerdt, H. M. L. Lambermont-Thijs, U. S. Schubert, R. Hoogenboom,
[38] S. Bali, T. T. Chen, W. Chaikittisilp, C. W. Jones, Energy Fuels 2013, Macromol. Rapid Commun. 2012, 33, 92–96.
27, 1547–1554. [67] T. Saegusa, S. Kobayashi, Y. Nagura, Macromolecules 1974, 7, 265–
[39] A. Ahmadalinezhad, A. Sayari, Phys. Chem. Chem. Phys. 2014, 16, 272.
1529–1535. [68] S. Sinnwell, H. Ritter, Macromol. Rapid Commun. 2006, 27, 1335–
[40] D. Wang, X. Wang, C. Song, ChemPhysChem 2017, 18, 3163–3173. 1340.
[41] H. Lepaumier, D. Picq, P. Carrette, Ind. Eng. Chem. Res. 2009, 48, [69] L. Hu, R. Frech, D. T. Glatzhofer, R. Mason, S. S. York, Solid State
9061–9067. Ionics 2008, 179, 401–408.
[42] C. S. Srikanth, S. S. C. Chuang, J. Phys. Chem. C 2013, 117, 9196–
9205.
[43] W. Choi, K. Min, C. Kim, Y. S. Ko, J. W. Jeon, H. Seo, Y.-K. Park, M.
Choi, Nat. Commun. 2016, 7, 12640.

This article is protected by copyright. All rights reserved.


ChemSusChem 10.1002/cssc.201800438

FULL PAPER
Entry for the Table of Contents (Please choose one layout)

Layout 1:

FULL PAPER
HaPPI to Capture CO2: S. H. Pang, R. P. Lively, C. W. Jones*
Poly(propylenimine)-containing solid
adsorbents are demonstrated for CO2 Page No. – Page No.

Accepted Manuscript
capture from ambient air. These
Oxidatively-Stable Linear
materials are more resistant to
Poly(propylenimine)-Containing
oxidation than their poly(ethylenimine)
Adsorbents for CO2 Capture from
counterparts and show stable
Ultra-Dilute Streams
adsorption/desorption performance
over many cycles.

This article is protected by copyright. All rights reserved.

You might also like