Download as pdf or txt
Download as pdf or txt
You are on page 1of 39

Journal Pre-proof

Versatile surface modification of aerogels by click chemistry as an approach


to generate model systems for CO2 adsorption features in amine-containing
organosilica

Nele Klinkenberg, Alexander Klaiber, Magdalena Müller, Sebastian Polarz


PII: S1387-1811(19)30738-3
DOI: https://doi.org/10.1016/j.micromeso.2019.109879
Reference: MICMAT 109879

To appear in: Microporous and Mesoporous Materials

Received Date: 22 September 2019


Revised Date: 6 November 2019
Accepted Date: 7 November 2019

Please cite this article as: N. Klinkenberg, A. Klaiber, M. Müller, S. Polarz, Versatile surface modification
of aerogels by click chemistry as an approach to generate model systems for CO2 adsorption
features in amine-containing organosilica, Microporous and Mesoporous Materials (2019), doi: https://
doi.org/10.1016/j.micromeso.2019.109879.

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition
of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of
record. This version will undergo additional copyediting, typesetting and review before it is published
in its final form, but we are providing this version to give early visibility of the article. Please note that,
during the production process, errors may be discovered which could affect the content, and all legal
disclaimers that apply to the journal pertain.

© 2019 Published by Elsevier Inc.


Versatile Surface Modification of Aerogels by

Click Chemistry as an Approach to Generate

Model Systems for CO2 Adsorption Features

in Amine-Containing Organosilica

Nele Klinkenberg, Alexander Klaiber, Magdalena Müller, Sebastian Polarz*

Department of Chemistry, University of Konstanz, Konstanz D-78457, Germany

*Corresponding Author: Prof. Dr. Sebastian Polarz, sebastian.polarz@uni-konstanz.de

Abstract

The conversion of waste into valuable products is most appealing in the case of CO2, a

molecule which is produced in mass by our society and industries. Because its atmospheric

concentration correlates to climate change and the green-house effect, major efforts are on

the way to reduce the emission of CO2. One promising strategy is the separation of CO2

from the gas-phase (e.g. flue gases) by solid-adsorbents containing amine moieties. The

synthesis of tailor-made adsorbents with changing surface properties remains a challenge.

This work presents a click chemistry approach that enables the easy modification of

organosilica materials with functional groups that can be used as model systems to study

1
the influence of surface chemistry on CO2 adsorption. As an example, the modification of

the materials with primary amines is discussed in detail but furthermore the approach offers

the possibility to tailor the surface properties using any desired functional group. The

increased affinity of the used copper catalyst introduced some difficulties but we were able

to remove all remains of copper. With this approach, we were able to synthesize materials

with different degrees of functionalization up to 80 %. This approach for the development

of new carbon capture model systems offers high functionalization combined with the

flexibility of a post-functionalization approach. Thus, surface chemistry can be tailored to

study the influence of surface chemistry on CO2 adsorption. As an example for the model

character of our materials, we could show that the heat of adsorption can be tuned by

systematically varying the degree of amine functionalization.

Keywords: amine-functionalized materials; organosilica, aerogels; functional gradients;

carbon capture model systems; CO2 adsorption; click chemistry

1. INTRODUCTION

Despite many efforts to lower the carbon dioxide emission, the atmospheric CO2

concentration has reached a new record of 410 ppm in 2019.[1] As long as coal, gas and oil

remain the major resources for fuel and energy, carbon capture represents one of the only

options to lower the CO2 emission. On an industrial scale, aqueous monoethanolamine

(MEA) solutions are mainly used in Post-Combustion carbon Capture, which allows

2
selective capture of CO2.[2, 3] However, the regeneration of the MEA solution is still the

biggest problem.[2] The necessary thermal regeneration procedure to release the chemically

bound CO2 requires a substantial amount of energy, which consequently lowers the overall

efficiency of the entire plant.[4] Therefore, due to their lower heat capacity and easier

regeneration solid adsorbents have become a major research focus.

The most promising solid adsorbents are amine-grafted porous materials[5-7], metal

organic frameworks (MOFs)[8-10], covalent organic/triazine frameworks

(COFs/CTFs)[11-13], Zeolites[14] and carbon materials[15]. For the application as solid

adsorbents high capacity, selectivity, reversibility and fast kinetics are important.[3] To

maximize capacity high surface areas as well as large micropore volumes are necessary.

With surface areas as high as ~ 6000 m²/g and large micropore volumes, MOFs are the

state-of-the-art materials regarding capacity.[9, 11, 14] While high capacities are important

also the selectivity towards CO2 is crucial as in typical applications of carbon capture

materials, such as capture from flue gas, only ~15 % of the gas mixture is CO2 while the

main component is N2. Additionally, trace gases such as NO2, SO2 or O2 are present.[3, 16]

Therefore, selective capture of CO2 is important. Selective capture of CO2 can be achieved

by introducing CO2-philic groups such as amine groups into the material. In this field,

amine-functionalized silica materials are widely discussed.[3, 5] A high affinity (heat of

adsorption) of the amine groups towards CO2 leads to selective uptake of CO2 over other

gases. While high affinity is important for selective uptake, more energy is needed for

regeneration. Therefore, the ideal material for an efficient and economic, reversible CO2

capture combines a high affinity towards CO2 with an easy release of CO2. According to a

3
recent review article published in 2017, the ideal adsorption enthalpy is in the range of

35-50 kJ/mol.[3] To optimize materials to meet these requirements, a deeper understanding

of the interaction of CO2 with the materials surface is necessary.

Investigations regarding the interactions of CO2 with the material surface include the

identification of the binding species found in these materials. Here, FTIR and NMR studies

have identified the most important binding species such as carbamate, carbamic acid and

bicarbonate.[17-23] Also, there are some publications that discuss the influence of

parameters such as the amine linker length[24], amine content or type[22]. A first example

studying the influence of the surrounding environment of the amine group, active in carbon

capture is the publication by Danon et al.[20] They studied the influence of surface silanol

groups on the adsorption process. Still, further research is necessary to fully understand for

example the influence of amine density and the surrounding environment on the affinity or

selectivity of a material.[3, 6, 25] For that, new model systems are needed that exhibit

defined surface chemistry that can be systematically altered to study these effects on the

CO2 adsorption. In this publication, we introduce such a model system.

A good model system should combine the possibility to vary the degree of

functionalization, as well as the functional group itself. High densities of functional groups

should be realized essentially meaning high degrees of functionalization. In addition, the

introduction of neighboring groups should be possible to study their influence on the CO2

adsorption. Classical approaches to functionalize silica materials are grafting, co-

condensation or the use of bridging silsesquioxane precursors.[26] Due to comercial

availability of the presursors, co-condensation or grafting are most commonly used in

4
literature to obtain amine-functionalized materials. As for both methods the degree of

organic modification is fairly low, the use of bridging silsesquioxane precursors becomes of

interest. These precursors consist of two silica centers that are connected by an organic

entity (RO)3Si-R-Si(OR)3 leading to materials with a degree of functionalization of 100 %.

The maximized density of functional organic groups in these materials makes them

interesting for the use as carbon capture model systems. The synthesis of such

silsesquioxane precursors is demanding making it a tedious process varying the organic

function. Therefore, our group has developed an azide-containing silsesquioxane precursor

(1,3-bis-tri-isopropoxysilyl-phenyl-5-azide) enabling further post-functionalization by

using click chemistry.[27] Using click chemistry to introduce further functionality

represents an elegant approach, as the materials surface properties can be easily changed by

using different click reagents without changing the structure of the network itself. This

way, high degrees of functionalization (up to 100 %) can be combined with the high

flexibility of a post-functionalization. This makes this an extremely versatile process for the

development of materials with varying surface properties highly suitable as carbon capture

model systems. According to a recent review by Patel et al., the combination of CO2-philic

groups immobilized in larger pore systems to increase diffusion represent ideal materials to

ensure both selective capture with fast kinetics.[3] Therefore, we use an aerogel

architecture that combines a high internal surface area with pores that are typically in the

macroporous region. Aerogels are also considered as promising carbon capture

materials.[28, 29]

5
The current paper focusses on organosilica aerogels with an inherent multimodal pore

structure as model systems for CO2 adsorption. We used a highly versatile click-chemistry

approach for the development of new carbon-capture model systems by introducing amine-

functionalities into the material. The described approach offers a high flexibility towards

changing the degree of functionalization as well as changing the functionality itself and its

distribution as shown in Scheme 1. We also showed that we could systematically change

the degree of functionalization within one material yielding a gradient in the surface

functionalization. Finally, the model character of the obtained materials was demonstrated

by showing that by varying the degree of functionalization we could change the heat of

adsorption towards CO2.

2. EXPERIMENTAL

2.1. Synthesis

2.1.1. General Remarks.

1,3-bis-tri-isopropoxysilyl-phenyl-5-azide as well as the aerogel materials were synthesized

according to literature procedure.[27] Copper(I)-tetrakis(acetonitrile) hexafluorophosphate

and Boc-propargylamine and Propargylamine were purchased by SigmaAldrich and TCI

and used without purification.

2.1.2. Material Synthesis (AzSil (1)).

117.2 µL (0.22 mmol, 116 mg) 1,3-bis-tri-isopropoxysilyl-phenyl-5-azide were dissolved in

1 mL ethanol (p.a.). 50 µL aqueous 1M HCl were added under stirring and the solution was

heated to 60 °C overnight. Then the reaction was cooled to room temperature, 50 µL

aqueous NH3 (25 wt%) were added and the solution was stirred for one minute. The

6
reaction solution was filled into a 3 mL syringe and sealed. After 15 minutes, gelation did

occur as can be seen by the turbidity. After aging for 22 h at room temperature, the

monolith was removed from the syringe and soaked in acetone. Acetone was replaced once

a day for five times.

2.1.3. Post-functionalization via click chemistry (AmSil (3a)).

85 mg (0.23 mmol, 1 eq) copper(I) tetrakis(acetonitrile) hexafluorophosphate were

dissolved in 10 mL acetone p.a.. The degree of functionalization was varied by addition of

different equivalents (0.5 – 2 eq) of Boc-propargylamine. The AzSil (1) (0.22 mmol)

monolith was added to the reaction mixture. The monolith was kept in place by a perforated

syringe. The reaction was stirred at room temperature for 48 h. Afterwards, the reactions

solution was exchanged against acetone and the monolith was washed three times with

acetone while waiting 1 day in between washing step. The monolith was put into a 4 M

HCl in acetone solution for two days. This solution was exchanged once after 24 h. After

another washing step with acetone the monolith was placed in a solution of 500 µL NH3

(25 %) in 10 mL acetone for 24 h. This step was repeated twice. Then, the monolith was

put into acetone for 24 h. This step was again repeated three times before the material was

dried supercritically.

2.1.4. Radial gradient generation.

Radial gradients are established by fully dipping the monolith into the reaction solution as

for homogeneously functionalized materials. To achieve the gradient architecture 3 eq of

alkyne are used as well as much shorter reaction times (12-24 h).

2.1.5. Vertical gradient generation.

7
The solvent inside the monolith was exchanged against ethanol using several washing

steps. For the embedding of the monolith inside the shrinking tube, a procedure was used

developed by Spitzbarth et al..[30] The monolith was placed in the middle of a polyolefine

shrinking tube, cut to the right length, into a 10 mL beaker filled with ethanol. The beaker

was placed into a pressure cooker filled with 500 mL of ethanol and heated to 90 °C

(2 bar). After the right temperature and pressure was reached, the heating plate was turned

off and the cooker was cooled to room temperature. The monolith, now embedded in the

shrinking tube, was removed from the pressure cooker and the solvent was exchanged with

acetone, which was used for the functionalization reaction. Again, the same procedure was

applied, as described 2.3. Instead of holding the monolith in place with a syringe, the

shrinking tube is attached to a glass stopper and the reaction is performed in a 25 mL round

bottom flask by only dipping the lower part of the monolith into the reaction solution. After

the completion of the reaction the monolith was removed from the shrinking tube by

dipping it into dichloromethane for 15 minutes. Thereby, the shrinking tube was swollen

and the monolith was removed. The solvent was exchanged against acetone.

2.1.6. Supercritical drying.

The monoliths were placed into the sample holder filled with acetone. The holder was

placed into the dryer chamber and closed. The chamber was filled with liquid CO2

(p ~ 60 bar). The acetone was exchanged with liquid CO2 by removing the acetone through

the release valve. After complete removal of acetone from the chamber, it was heated to

42 °C (p ~ 90 bar). The CO2, which is supercritical under these conditions, was released

over a period of 4-5 hours. After complete release of CO2, the dried monoliths were

removed from the dryer chamber.

8
2.2. Characterization.

ATR-IR Spectroscopy. IR spectra were recorded using a Perkin Elmer Spectrum 100

spectrometer with an ATR unit. All spectra were background corrected and normalized to

the Si-O-Si vibration at 1044 cm -1. Physisorption Measurements. N2 physisorption

measurements were performed with the Micromeritics ASAP 2020 at -196 °C. Prior to

analysis the samples were degassed at 85 °C for 720 min. CO2 adsorption measurements.

The CO2 adsorption isotherms were determined with the Micromeritics ASAP 2020 at 30

°C, 40 °C and 50 °C. The Micromeritics ISO Controller was used to maintain the desired

temperature. Prior to analysis the samples were outgassed at 85°C for at least three hours.

The equilibration interval used was 10 s. The measurements were corrected by additional

blank tube measurements under identical measurement conditions. The heat of adsorption

was determined using the MicroActive Plugin (4.06) by Micromeritics. The program

determines the heat of adsorption by applying the Clausius-Clapeyron-Equation (ln =



+ , C = unknown constant) using isotherms at 30 °C, 40°C and 50°C to determine the


slope of ln as a function of 1/RT. Thermogravimetric Analysis (TGA). TGA


measurements were performed under O2 using a Netzsch STA F3 Jupiter with a heating rate

of 5 K/min until 300 °C and 10 K/min for 300-1000 °C. Magic-angle spinning (MAS)-

NMR. NMR spectra were recorded using a Bruker DRX 400 spectrometer. Spectra were

recorded with a spinning frequency of 10000 Hz and a relaxation delay of 120 s. Scanning

electron microscopy (SEM). SEM- images were obtained with a Zeiss Auriga CrossBeam

microscope. Ninhydrin Assay. 1.5 mL of a Ninhydrin solution (69 mM in EtOH) were

9
added to 3.2 mg of sample. The mixture was heated to 100 °C for 10 min and cooled to

room temperature. 1.5 mL EtOH were added and a UV-Vis spectrum was recorded of the

supernatant. The supernatant was further diluted with EtOH due to high initial absorbance.

UV-Vis spectroscopy. UV-Vis spectra were recorded using a Thermo Scientific Genesys

10S UV-Vis Spectrophotometer. XPS. XPS measurements were conducted with an

Omicron Multiprobe ultra-high vacuum system (base pressure of 10−11 mbar) with Al Kα

radiation (hν = 1486.7 eV) from a dual-anode X-ray source. The sample (powder pressed

into a pellet) was fixated with silver lacquer on a molybdenum sample holder.

3. RESULTS AND DISCUSSION

3.1. Reference-State: The Phenylazide Organosilica Aerogel Monoliths.

The organosilica materials were prepared according to a method, which was modified

slightly compared to ref. [27]. Because the unmodified phenylazide material AzSil (1) is an

important reference state, the corresponding characterization data are summarized in

Electronic Supplementary Information (ESI) Figure S1.

3.2. Copper-free organosilica materials containing amine groups.

For establishing amine-containing organosilica aerogels, the copper-catalyzed click

reaction was performed using propargyl amine (2a); see Scheme 1. After the reaction, we

saw the resulting materials have become dark green in color (see Figure 1). The green

color stems for a broad absorption signal between 800-600 nm (Figure 1d) according to

diffuse reflectance UV-VIS spectroscopy. Because the target materials containing amine

groups AmSil (3a) should be colorless, the coloration represents an indication for CuII

10
coordinating to the NH2 groups of the materials. The presence of paramagnetic (d9) CuII -

centers can be confirmed by electron paramagnetic resonance spectroscopy (EPR) (Figure

1e) and energy-dispersive X-ray spectroscopy (EDX) (Figure 1f). The EPR spectrum

shows the characteristic signal for Cu2+ with a g⊥ = 2.0464 and g∥ = 2.3210 and a hyperfine

splitting constant A∥ = 196 G similar to materials functionalized with copper that have been

reported previously.[31] The small signal with a g value of 1.9834 that can be observed for

AzSil (1) and AmSil (3a) as well is probably due to material defects in the solid state.

The described scenario is unfavorable for future binding of CO2 to these amine groups,

because the lone pair of the amine groups is blocked (Figure 1a). Thus, a strategy for

avoiding copper coordination is necessary. Therefore, we tested, if the use of a tert-

butoxycarbonyl (boc) protected propargyl amine (2b) for the click reaction solves the

problem. EPR spectroscopy, EDX and XPS (ESI Figure S4c) confirm, there is no copper

present anymore, and the resulting material 3b is now free of color. The success of the

deprotection of the amine groups (3a'' → 3a) was investigated using infrared (IR)

spectroscopy and 13C- solid-state magic angle spinning nuclear magnetic resonance (MAS-

NMR) spectroscopy (Figure 2a). The aromatic signals in material AzSil (1) (138-125 ppm)

are shifted to 169-147 ppm because of the electron withdrawing effect of the triazol ring,

whose carbon signals are in the same region. The signals characteristic for the boc-group

(180 ppm (COOR), 103 ppm (CCH3), 49 ppm (CCH3) have disappeared after deprotection

and the signal for the free amine appears (58 ppm).[32] According to literature[33, 34],

traces of copper can lead to degradation reactions of amine groups under aerobic

conditions. Therefore, we thoroughly checked that the amine groups are stable in our

11
material. A Ninhydrin-Assay clearly shows the presence of accessible and active amine

groups. The latter conclusion is supported by 1H-NMR and XPS studies. The additional

data are summarized in ESI Figure 4. The characteristic IR signals of the boc-group can be

found at 1685, 1515 cm-1 and 834 cm-1 (Figure 2b).[35] The reduction in intensity of these

signals is also in agreement with a successful deprotection reaction. Thermogravimetric

analysis (TGA) was performed as well shown in ESI Figure S2. The residual mass at T =

800°C is SiO2, and one can deduce the content of organic constituents from the degree of

total mass-loss (∆mtot). The TGA trace of AmSil (3a) shows less organic content compared

to the boc-protected variant (3a''), which is expected. Therefore, also the TGA data are

consistent with the successful removal of the boc group.

The quantification of the TGA traces of the starting material AzSil (1) (∆mtot = -51.7%)

compared to AmSil (3a) (∆mtot = -61.8%) is consistent with a degree of conversion x = 0.83

of all azide groups to amine functions via click chemistry (see also Scheme 1).

Additionally, the degree of conversion was verified using hpdec-MAS-NMR spectroscopy.

Through integration of the corresponding signals, (see ESI Figure S3) a degree of

conversion x = 0.81 was determined which is in good agreement with the quantification

using TGA. However, the most important tool is for sure FT-IR spectroscopy, respectively

the analysis of the signal intensities. Because all materials contain the same number of

siloxane bonds (Si-O-Si), it is possible to reference the spectra to the band at 1050 cm-1.

Then, the highest intensity of the azide vibration at 2111 cm-1 stands for a degree of

functionalization x = 0 (= AzSil (1)) and a potential intensity of zero would stand for a

degree of functionalization degree of x = 1. Then, we assume the intensity of the azide

12
vibration is a linear function of x. We conclude from IR spectra (ESI Figure S3), that the

maximum degree of functionalization is x = 0.80±0.05, which consistent with the value

obtained by the other methods (see above). Furthermore, one can obtain spatially resolved

information by the application of IR-microscopy techniques (Figure 2c). The measurement

shows the homogeneous modification of the material with amine functions.

The advantage of the click chemistry approach is that it is highly flexible allowing the

incorporation of any linker carrying an alkyne group. In addition to the primary amine with

a C1-Spacer (2a), we could show that primary amines with a C3-chain (2b) as well as

secondary (2c) and tertiary amines (2d) could be incorporated (ESI Fig.S5).

Materials with different amine content were synthesized by varying the DOF. The variation

of DOF can be observed by IR-spectroscopy as the intensity of the azide vibration (2112

cm-1) decreases thus indicating the increase of the DOF (see Figure 3a). Simultaneously,

the band corresponding the amine vibration (1622 cm-1) is increased.[35] For physical

characterization of the material N2- physisorption measurements were conducted which can

be seen in Figure 3b. Upon functionalization, the overall pore volume is increased as can

be seen by the higher adsorbed amounts for pressures of p/p0 close to 1. This can be

explained by stabilization of the macropores during the supercritical drying process. The

increased organic amount stabilizes the overall network thus leading to less collapse of the

pore structure. This effect has already been described by Schachtschneider et al.[27] For

AmSil with the highest DOF (x=0.8, dark red) the adsorbed amount of N2 in the

microporous region (p/p0 < 0.2) is decreased. This effect is probably due to pore blockage

of the smaller pores.[36] With BET surface areas typically in the range of 500-800 m2/g,

13
the materials (see ESI Table S1) are characterized with high surface areas also after

functionalization. Figure 3c shows an exemplary SEM micrograph of AmSil (x = 0.54)

where the porous structure of the material can be nicely seen. The high surface area as well

as high porosity makes these materials excellent candidates for carbon dioxide capture.

3.4. Creation of Functional Gradients. The use of this post-functionalization approach

enables not only the homogeneous functionalization of the materials but also a graded

functionalization. By only dipping one side of the monolith into the reaction solution and

letting the reaction solution diffuse into the material gradients can be established.[27]

While it is necessary to remove the copper for later applications it can be used as a probe to

visualize the obtained gradient. Therefore, one can already see the establishment of the

gradient in a vertical (Figure 4b) or radial manner (Figure 4c) by the gradient in green

color. Additionally, EDX-mapping using copper as a probe can be used to visualize the

gradient (Figure 4a). Again, using the intensity of the azide vibration the gradient can also

be directly imaged by IR-Microscopy. For that, IR-spectra are obtained in reflection along

the vertical axis of the monolith (Figure 4e, Inset). As for the ATR-IR also in the reflection

spectra the decrease of the azide vibration can be observed (Figure 4d). Figure 4e shows

the decrease in intensity of the azide vibration along the vertical axis of the monolith while

Figure 4d shows the reflection spectra of two example positons showing the decreasing

intensity of the IR band for the azide vibration. The use of the shrinking tube for the

establishment of the vertical gradient ensures that diffusion can only take place vertically

and therefore no radial gradient is established in this case (see ESI Figure S6).

3.5. CO2 Adsorption Studies.

14
To test our material as model systems for CO2 adsorption, homogeneously functionalized

AmSil (3a’) with varying DOF were used as a proof of concept. Volumetric CO2

adsorption studies at T = 30 °C, 40 °C and 50 °C were performed. Figure 5a shows the

CO2 adsorption isotherms at T = 30 °C of AmSil (3a’) with varying DOF from x = 0.23

(yellow) to x = 78 (dark red) (CO2 isotherms for T = 40 °C and T = 50 °C can be found in

the ESI Figure S7). For AmSil (3a’) with higher amine content (dark red, DOF x = 0.78)

higher capacities are obtained. For higher temperatures, the overall capacity is decreased

(ESI Figure S7). For T = 30 °C capacities at 920 mmHg range between 1-1.4 µmol/m2

(respectively ≈ 1mmol CO2/gadsorbent; ≈ 44mg CO2/gadsorbent). First, we compared the uptake

capacities to other amine-functionalized organosilica materials known in literature.

According to summarizing tables published for instance by Chen et al.[17], Zhou et al.[37]

or Gargiulo et al.[38] adsorption capacities vary from 0.4-3 mmol CO2/gadsorbent. Therefore,

one can conclude the performance of our materials regarding CO2 storage capacity is only

average and, thus, is not interesting for an industrial application. However, we would like to

emphasize, maximization of the CO2 capacity was never our goal, as our materials cannot

compete with microporous solids such as MOFs and COFs with much higher internal

surface area.

One can also see the CO2 adsorption isotherm shows a Henry-type behavior. Therefore, like

said before, our materials should not be considered for concrete carbon capture applications

on an industrial scale. Different to other materials reported in the literature, the Henry-type

behavior is a consequence of the absence of micropores. While the absence of micropores

reduces the total surface area and, thus also the uptake capacities, we think it is an

15
advantage for a model material, because one can now study the effects of the surface

chemistry uninfluenced by superposition to deviations in structural effects like a difference

in microporosity. How differences in microporosity can influence the adsorption behavior

of CO2 in particular, has already been discussed in several articles.[39-41] While for

application, capacity is of course an important factor to consider, for the use as model

systems we are more interested in the influence of surface chemistry on the selectivity and

affinity towards CO2. Figure 5a additionally shows the adsorption of nitrogen at T = 30 °C.

It can be seen that for all AmSil (3a’) materials the adsorption of nitrogen under identical

conditions is negligible thus meaning that the proposed materials exhibit a high selectivity

towards CO2. Another important parameter to consider is the isosteric heat of adsorption

(Qist) representing the affinity of the material towards CO2. Qist can be determined by

measuring isotherms at different temperature and subsequent calculation with the Clausius-

Clayperon Equation.[42] Figure 5b shows the dependence of Qist on the adsorbed quantity

of CO2 (nads) for AmSil (3a’) with DOF x = 0.23, x = 0.36, x = 0.54 and x = 0.78. Qist is

highest for low nads and decreases towards higher nads. This is believed to be due to

increased steric hindrances due to already formed binding species on the surface.[22, 43]

With increasing DOF Qist is increased showing that an increased amine content leads to an

increased affinity towards CO2 which has been also reported in literature.[43] Therefore,

the use of this click chemistry approach enables the targeted synthesis of materials with

different interaction strengths towards CO2, which is important considering that the ideal

adsorbent is not necessarily the one with the highest capacity or heat of adsorption. As

described by Patel et al. the ideal heat of adsorption should be between 35-50 kJ/mol.[3]

Previous reports also show that adsorption enthalpies drop with increasing surface

16
coverage.[18, 44, 45] The values range from 120-60 KJ/mol for quasi-empty surfaces to 40-

20 KJ/mol for higher surface coverages. As can be seen in Figure 5b AmSil (3a’) the

values reported by us are in a comparable region. However, we believe it is also important

to see, the adsorption enthalpy does not only depend on surface coverage (as expected), but

also on the surface density of the amine groups (Fig. 5c). The adsorption enthalpy can be

increased up to a factor of 250%, if there are other amine groups (x = 0.8) in vicinity. The

effect is less pronounced for higher CO2 coverages, but can still be clearly identified. The

latter results show, there is clearly a neighboring group effect through amine groups in

vicinity on CO2 adsorption enthalpy. The described click chemistry approach therefore

presents a new approach to generate model systems to be able to better understand the

influence of surface chemistry on the adsorption process. Additionally, the possibility of

generating gradients presented in this work offers the possibility of not only being able to

generate materials with targeted Qist but also to generate materials with a gradient in Qist

within one material as it was shown that Qist varies with varying amine content.

4. CONCLUSIONS

In this work, we showed a versatile synthesis route for the synthesis of amine-

functionalized organosilica aerogels via a click chemistry approach. Due to the high affinity

of amines to copper, remaining copper was found in the materials. The use of boc-protected

amines led to the complete removal of the remaining copper, and the presence of the amine

groups was proven by several independent methods. The CO2 uptake capacities was in the

expected range compared to other amine-functionalized materials known in the literature,

but definitely not extraordinarily high. Thus, the advantage of our materials is a different

17
one. In particular, because the materials can be prepared without a significant amount of

micropores, one can study effects of the surface chemistry on CO2 adsorption behavior

without (major) interference of structural factors. The developed synthesis approach offers

the possibility of easily varying the DOF of the material making this a valuable strategy to

obtain materials suitable as model systems to study the effect of amine content, amine type

as well as amine density on the CO2 adsorption. Via this route a maximum degree of

functionalization of up to 80 % can be achieved which is much higher than the DOF of

commonly used amine-functionalized silica materials obtained by grafting or co-

condensation. This click chemistry approach enables not only homogeneous

functionalization but also functionalization in a gradient manner making vertical as well as

radial gradients possible. CO2 adsorption measurements show that AmSil (3a’) changing

the DOF the isosteric heat of adsorption Qist can be changed. This leads to the possibility to

synthesize materials with targeted adsorption strength towards CO2 making the AmSil (3a’)

materials ideal model systems to further study and optimize the CO2 uptake and release.

The investigation of gradients in functionalization and thereby also in heat of adsorption

within one material represents a challenging task and should be tackled in future research.

Additionally, our approach offers the possibility to introduce a second neighboring group,

which means the environment of the amine can be altered. We hope the adjustment of the

amine-group environment by using appropriate neighboring groups will enable to tune the

selectivity for CO2 in mixtures with methane or water vapor.

Author Contributions

18
The manuscript was written by Nele Klinkenberg and Sebastian Polarz. All authors have

given approval to the final version of the manuscript.

Acknowledgments

We thank the SFB1214 and the IRTG for support. We thank the Electronmicroscopy Center

and the Nanolab of the University of Konstanz for providing the measurement facilities.

We thank Tobias Lemke for many fruitful scientific discussions and proof-reading of the

manuscript. We thank Ilona Wimmer for the conduction of XPS measurements. We thank

Julian Steinbrecher for his contribution to material synthesis.

19
REFERENCES

[1] J.P.L. NASA, Global Climate Change- Vital Signs of the Planet, in: H. Shaftel (Ed.)

2019.

[2] G.T. Rochelle, Amine scrubbing for CO2 capture, Science, 325 (2009) 1652-1654.

[3] H.A. Patel, J. Byun, C.T. Yavuz, Carbon Dioxide Capture Adsorbents: Chemistry and

Methods, ChemSusChem, 10 (2017) 1303-1317.

[4] D.M. D'Alessandro, B. Smit, J.R. Long, Carbon Dioxide Capture: Prospects for New

Materials, Angew. Chem. Int. Ed., 49 (2010) 6058-6082.

[5] C. Chen, S. Zhang, K.H. Row, W.S. Ahn, Amine-silica composites for CO2 capture: A

short review, J. Energy Chem., 26 (2017) 868-880.

[6] A. Alonso, J. Moral-Vico, A. Abo Markeb, M. Busquets-Fité, D. Komilis, V. Puntes, A.

Sánchez, X. Font, Critical review of existing nanomaterial adsorbents to capture carbon

dioxide and methane, Sci. Total Environ., 595 (2017) 51-62.

[7] M. Jahandar Lashaki, S. Khiavi, A. Sayari, Stability of amine-functionalized CO2

adsorbents: a multifaceted puzzle, Chem. Soc. Rev., 48 (2019) 3320-3405.

[8] C.A. Trickett, A. Helal, B.A. Al-Maythalony, Z.H. Yamani, K.E. Cordova, O.M. Yaghi,

The chemistry of metal–organic frameworks for CO2 capture, regeneration and conversion,

Nat. Rev. Mater., 2 (2017) 17045.

[9] M. Ding, R.W. Flaig, H.-L. Jiang, O.M. Yaghi, Carbon capture and conversion using

metal–organic frameworks and MOF-based materials, Chem. Soc. Rev., 48 (2019) 2783-

2828.

20
[10] J. Yu, L.-H. Xie, J.-R. Li, Y. Ma, J.M. Seminario, P.B. Balbuena, CO2 Capture and

Separations Using MOFs: Computational and Experimental Studies, Chem. Rev., 117

(2017) 9674-9754.

[11] A. Modak, S. Jana, Advancement in porous adsorbents for post-combustion CO2

capture, Microporous Mesoporous Mater., 276 (2019) 107-132.

[12] J. Ozdemir, I. Mosleh, M. Abolhassani, L.F. Greenlee, R.R. Beitle, M.H. Beyzavi,

Covalent Organic Frameworks for the Capture, Fixation, or Reduction of CO2,

Front.Energy Res., 7 (2019).

[13] H. Wang, D. Jiang, D. Huang, G. Zeng, P. Xu, C. Lai, M. Chen, M. Cheng, C. Zhang,

Z. Wang, Covalent triazine frameworks for carbon dioxide capture, J. Mater. Chem. A, 7

(2019) 22848-22870.

[14] M. Abu Ghalia, Y. Dahman, Development and Evaluation of Zeolites and Metal–

Organic Frameworks for Carbon Dioxide Separation and Capture, Energy Technol., 5

(2017) 356-372.

[15] R. Ben-Mansour, M.A. Habib, O.E. Bamidele, M. Basha, N.A.A. Qasem, A.

Peedikakkal, T. Laoui, M. Ali, Carbon capture by physical adsorption: Materials,

experimental investigations and numerical modeling and simulations – A review, Appl.

Energy, 161 (2016) 225-255.

[16] C.H. Yu, C.H. Huang, C.S. Tan, A Review of CO2 Capture by Absorption and

Adsorption, Aerosol Air Qual. Res., 12 (2012) 745-769.

[17] G.S. Foo, J.J. Lee, C.-H. Chen, S.E. Hayes, C. Sievers, C.W. Jones, Elucidation of

Surface Species through in Situ FTIR Spectroscopy of Carbon Dioxide Adsorption on

Amine-Grafted SBA-15, ChemSusChem, 10 (2017) 266-276.

21
[18] C. Knöfel, C. Martin, V. Hornebecq, P.L. Llewellyn, Study of Carbon Dioxide

Adsorption on Mesoporous Aminopropylsilane-Functionalized Silica and Titania

Combining Microcalorimetry and in Situ Infrared Spectroscopy, J. Phys. Chem. C, 113

(2009) 21726-21734.

[19] L. Mafra, T. Čendak, S. Schneider, P.V. Wiper, J. Pires, J.R.B. Gomes, M.L. Pinto,

Structure of Chemisorbed CO2 Species in Amine-Functionalized Mesoporous Silicas

Studied by Solid-State NMR and Computer Modeling, J. Am. Chem. Soc., 139 (2017) 389-

408.

[20] A. Danon, P.C. Stair, E. Weitz, FTIR Study of CO2 Adsorption on Amine-Grafted

SBA-15: Elucidation of Adsorbed Species, J. Phys. Chem. C, 115 (2011) 11540-11549.

[21] A. Sayari, Y. Belmabkhout, Stabilization of Amine-Containing CO2 Adsorbents:

Dramatic Effect of Water Vapor, J. Am. Chem. Soc., 132 (2010) 6312-+.

[22] M.W. Hahn, J. Jelic, E. Berger, K. Reuter, A. Jentys, J.A. Lercher, Role of Amine

Functionality for CO2 Chemisorption on Silica, J. Phys. Chem. B, (2015).

[23] Z. Bacsik, N. Ahlsten, A. Ziadi, G. Zhao, A.E. Garcia-Bennett, B. Martín-Matute, N.

Hedin, Mechanisms and Kinetics for Sorption of CO2 on Bicontinuous Mesoporous Silica

Modified with n-Propylamine, Langmuir, 27 (2011) 11118-11128.

[24] N.A. Brunelli, S.A. Didas, K. Venkatasubbaiah, C.W. Jones, Tuning Cooperativity by

Controlling the Linker Length of Silica-Supported Amines in Catalysis and CO2 Capture,

J. Am. Chem. Soc., 134 (2012) 13950-13953.

[25] B. Sreenivasulu, I. Sreedhar, P. Suresh, K.V. Raghavan, Development Trends in

Porous Adsorbents for Carbon Capture, Environ. Sci. Technol., 49 (2015) 12641-12661.

22
[26] F. Hoffmann, M. Cornelius, J. Morell, M. Froba, Silica-based mesoporous organic-

inorganic hybrid materials, Angew. Chem. Int. Ed., 45 (2006) 3216-3251.

[27] A. Schachtschneider, M. Wessig, M. Spitzbarth, A. Donner, C. Fischer, M. Drescher,

S. Polarz, Directional MaterialsNanoporous Organosilica Monoliths with Multiple

Gradients Prepared Using Click Chemistry, Angew. Chem. Int. Ed., 54 (2015) 10465-

10469.

[28] S. Cui, W.W. Cheng, X.D. Shen, M.H. Fan, A. Russell, Z.W. Wu, X.B. Yi,

Mesoporous amine-modified SiO2 aerogel: a potential CO2 sorbent, Energy Environ. Sci., 4

(2011) 2070-2074.

[29] J.E. Amonette, J. Matyas, Functionalized silica aerogels for gas-phase purification,

sensing, and catalysis: A review, Microporous Mesoporous Mater., 250 (2017) 100-119.

[30] M. Spitzbarth, T. Lemke, M. Drescher, In Situ Monitoring of Diffusion of Guest

Molecules in Porous Media Using Electron Paramagnetic Resonance Imaging, (2016)

e54335.

[31] M. Halder, P. Bhanja, M.M. Islam, A. Bhaumik, S.M. Islam, Chiral copper-salen

complex grafted over functionalized mesoporous silica as an efficient catalyst for

asymmetric Henry reactions and synthesis of the potent drug (R)-isoproterenol, New J.

Chem., 42 (2018) 11896-11904.

[32] D. Canet, NMR-Konzepte und Methoden, InterEditions, Paris, 1991.

[33] J. Kim, G. Golime, H.Y. Kim, K. Oh, Copper(II)-Catalyzed Aerobic Oxidation of

Amines: Divergent Reaction Pathways by Solvent Control to Imines and Nitriles, Asian J.

Org. Chem., 8 (2019) 1674-1679.

23
[34] J. Kim, S.S. Stahl, Cu/Nitroxyl-Catalyzed Aerobic Oxidation of Primary Amines into

Nitriles at Room Temperature, ACS Catalysis, 3 (2013) 1652-1656.

[35] B.H. Stuart, Organic Molecules, Infrared Spectroscopy: Fundamentals and

Applications, John Wiley & Sons, Ltd2005, pp. 71-93.

[36] S. Storck, H. Bretinger, W.F. Maier, Characterization of micro- and mesoporous solids

by physisorption methods and pore-size analysis, Appl. Catal., A: General, 174 (1998) 137-

146.

[37] K. Zhou, S. Chaemchuen, F. Verpoort, Alternative materials in technologies for

Biogas upgrading via CO2 capture, Renew. Sust. Energ. Rev., 79 (2017) 1414-1441.

[38] N. Gargiulo, F. Pepe, D. Caputo, CO2 Adsorption by Functionalized Nanoporous

Materials: A Review, J. Nanosci. Nanotechnol., 14 (2014) 1811-1822.

[39] D. Lozano-Castelló, D. Cazorla-Amorós, A. Linares-Solano, Usefulness of CO2

adsorption at 273 K for the characterization of porous carbons, Carbon, 42 (2004) 1233-

1242.

[40] M. Oschatz, M. Antonietti, A search for selectivity to enable CO2 capture with porous

adsorbents, Energy Environ. Sci., 11 (2018) 57-70.

[41] L. Yu, M. Kanezashi, H. Nagasawa, T. Tsuru, Fabrication and CO2 permeation

properties of amine-silica membranes using a variety of amine types, J. Membrane Sci.,

541 (2017) 447-456.

[42] P.A. Webb, An Introduction To The Physical Characterization of Materials by

Mercury Intrusion Porosimetry with Emphasis On Reduction And Presentation of

Experimental Data, Micromeritics Instrument Cooperation, Norcross, Georgia, 2001.

24
[43] M.A. Alkhabbaz, P. Bollini, G.S. Foo, C. Sievers, C.W. Jones, Important Roles of

Enthalpic and Entropic Contributions to CO2 Capture from Simulated Flue Gas and

Ambient Air Using Mesoporous Silica Grafted Amines, J. Am. Chem. Soc., 136 (2014)

13170-13173.

[44] E. Berger, M.W. Hahn, T. Przybilla, B. Winter, E. Spiecker, A. Jentys, J.A. Lercher,

Impact of solvents and surfactants on the self-assembly of nanostructured amine

functionalized silica spheres for CO2 capture, J. Energy Chem.

[45] K.S. Sánchez-Zambrano, L. Lima Duarte, D.A. Soares Maia, E. Vilarrasa-García, M.

Bastos-Neto, E. Rodríguez-Castellón, D.C. Silva de Azevedo, CO₂ Capture with

Mesoporous Silicas Modified with Amines by Double Functionalization: Assessment of

Adsorption/Desorption Cycles, Materials (Basel), 11 (2018) 887.

25
FIGURE CAPTIONS

Scheme 1. Overview of the amine-containing materials and their synthesis by copper-

catalyzed click chemistry.

Figure 1. (a) Reaction scheme leading to copper-free amine modified organosilica aerogels

(AmSil). Photographs of the copper-containing 3a' (b) and copper-free 3a'' materials (c).

UV-Vis spectra in diffuse reflectance (e), EPR spectra (f) and EDX spectra of AzSil (1)

(blue), (3a') (green) and (3a'') (red).

Figure 2. 13C-MAS-NMR (a) and FT-IR spectra (b) of the AzSil (1) as a reference (blue),

the boc-protected material 3a'' (grey) and the material with free amine groups 3a (red). (c)

IR microscopy measurements of the spatial intensity of the azide vibration; scalebar =

200µm.

Figure 3: IR-spectra (a), N2-isotherms (b) of AzSil (black) and AmSil with varying DOF x

= 0.23 – 0.8, yellow → red). (c) SEM micrograph of AmSil with x = 0.54 (light red). Scale

bar 2µm.

26
Figure 4: (a) SEM-Micrograph of AmSil (3a’) with overlaying EDX mapping (silicon:

blue, copper: green). Scale bar 1mm. Photographs of gradient materials with a vertical (b)

and a radial (c) gradient. (d) Exemplary IR spectra (colors in correspondence to 4e)

measured in reflection mode. (e) Intensity of the azide vibration depending on the position

in the material derived from IR-microscopy in reflection mode. Inset showing the image of

the material with the measurement positions. Scale bar 200 µm.

Figure 5: (a) CO2 (squares) and N2 (triangles) adsorption isotherms at 30 °C, (b) Isosteric

heat of adsorption Qist depending on the quantity adsorbed and depending on the DOF x of

AmSil (3a’) with DOF x = 0.23 (yellow), x = 0.36 (orange) and x = 0.54 (light red) and x =

0.78 (dark red). CO2 adsorption isotherms for T = 30 °C and 40 °C are shown in ESI

Figure S7. (c) Change of adsorption enthalpy referenced to Qist (DOF = 0.23) with

increasing surface amine concentration for low surface coverace (nads → 0; red points) and

higher surface coverage (nads = 0.1; blue points).

27
FIGURES

Scheme 1

28
Figure 1.

29
Figure 2.

30
Figure 3.

31
Figure 4.

32
Figure 5.

33
TOC.

34
Highlights:
• Versatile click-chemistry approach with degrees of functionalization up to 80 %
• Offers opportunity to generate vertical and radial gradients in functionalization
• Isosteric heat of adsorption can be varied systematically with amine concentration
• Approach enables generation of materials with gradients in heat of adsorption

You might also like