Next Article in Journal
Middle Latitude Geomagnetic Disturbances Caused by Hall and Pedersen Current Circuits Driven by Prompt Penetration Electric Fields
Next Article in Special Issue
Sr1-xKxFeO3 Perovskite Catalysts with Enhanced RWGS Reactivity for CO2 Hydrogenation to Light Olefins
Previous Article in Journal
Association between Ambient Air Pollutants and Pneumonia in Wuhan, China, 2014–2017
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Amine-Modified Biochar for the Efficient Adsorption of Carbon Dioxide in Flue Gas

1
State Key Laboratory of High–Efficiency Utilization of Coal and Green Chemical Engineering, Ningxia University, Yinchuan 750021, China
2
Department of Chemistry and Chemical Engineering, North Minzu University, Yinchuan 750021, China
3
Department of Chemistry, University College London, London WC1H 0AJ, UK
4
Department of Chemistry and Chemical Engineering, Weifang University, Weifang 261061, China
*
Author to whom correspondence should be addressed.
Atmosphere 2022, 13(4), 579; https://doi.org/10.3390/atmos13040579
Submission received: 22 February 2022 / Revised: 25 March 2022 / Accepted: 31 March 2022 / Published: 4 April 2022
(This article belongs to the Special Issue CO2 Sequestration, Capture and Utilization)

Abstract

:
Biochar-based carbonaceous adsorbents are gaining interest due to their high availability, ease of modification, and low cost; however, they show limited adsorption of CO2 in flue gas due to common textural properties. In this study, TEPA-modified biochar was used to prepare a solid amine adsorbent for the efficient capture of CO2 in flue gas. First, the porous biochar was prepared with FeCl3, Mg(NO3)2, and H2O (g) as activators and walnut shells as carbon sources. Next, the biochar was modified with TEPA to obtain a solid amine adsorbent. Porous texture properties and sample surface functional groups were characterized, and we measured the adsorption CO2 of the amine-modified biochar in a breakthrough adsorption device. Results showed that biochar has a large specific surface area (744.38 m2 g−1), a total pore volume of 1.41 cm3 g−1, and a high mesoporous volume ratio (82.7%). The high pore volume provided a more efficient support space for loading tetraethylenepentamine (TEPA). The adsorbent had an excellent CO2 adsorption capacity, corresponding to 2.82 mmol g−1, which increased to 3.31 mmol g−1 and kept water resistance at 10% H2O (g) simulated flue gas (SFG). The FTIR analysis showed that H2O (g) inhibited urea production after cyclic adsorption. Therefore, solid amine adsorbent created by amine-modified biochar has potential advantages in its application for capturing CO2 in SFG.

1. Introduction

CO2 emissions caused by fossil energy are important causes of global warming and the greenhouse effect [1]. Adsorption is the most promising method for reducing CO2 emissions from SFG because of its low corrosiveness and low energy consumption. At the same time, it does not require significant modifications to the considerable existing equipment [2,3,4]. The excellent adsorbent is the core of efficient CO2 adsorption. Solid amine adsorbents remove CO2 from SFG because of their multiple active sites, high adsorption capacity, broad application temperature range, and eco-friendliness [5,6,7,8]. The carrier is the core and essential for synthesizing the solid amine adsorbent. Commonly, carriers can be divided into noncarbon and carbon materials [7]. Noncarbon carriers include alkali-metal materials, metal oxides, metal–organic frameworks (MOFs), and molecular sieves [5,6,7]. Carbon material carriers include graphene oxide (GO), activated carbon fibers, carbon nanotubes, and ordered mesoporous carbon. Although these materials have excellent pore structure parameters, complex preparation methods and high production cost limit their large-scale production.
Compared with the abovementioned carriers, biochar has been extensively used in various fields because of its broad availability, low cost, well-strung pores, and ample surface functional groups. Biochar has a large specific surface area (SBET), total pore volume (VT), and pore diameter (Pd); hence, it is suitable as a carrier. By introducing different functional groups, adsorbents with different functions can be easily obtained [4,9]. Nevertheless, biochar obtained using traditional methods has multiple limitations; hence, more studies have been conducted with the aim of improving the structure of biochar [10,11,12]. Usually, chemical and physical activation are employed to improve the structural properties of biochar, such as the SBET and VT. KOH [13], H3PO4 [14], K2CO3 [15], HNO3 [16], O2 [17], ZnCl2 [18], FeCl3 [19], MgCl2 [20], and steam [21] had been used as activators to optimize the structure of biochar. Although the usage of KOH, NaOH, and other strong acid and alkali activators can yield a suitable SBET, the pore diameter is generally small. Furthermore, the loading of organic amine leads to the plugging of these pores, which is not conducive to CO2 diffusion and adsorption. Moreover, disposable activators (NaOH, H3PO4, K2CO3, HNO3) are highly corrosive and difficult to recycle. Their large-scale application would cause considerable damage to the environment. H2O (g) is considered to be an eco-friendly activator for creating new pores and expanding the pore diameter. However, if only H2O (g) is used as the activator, it is difficult to control the activation rate and pore structure of biochar [22,23,24]. Furthermore, metal salts activation was used to prepare biochar. Using the metal salts as activating agents to prepare biochar accelerates the pyrolysis of biomass, promotes graphitization, and creates pores for biochar. Liu et al. [20] obtained mesoporous biochar with an SBET of 421.4 m2 g−1 using high-temperature activation and pyrolysis of FeCl3 preloaded on the biomass surface. The coactivated biochar with two metal salts was more effective than the single metal salts biochar. Guo et al. [25] used MgCl2 and FeCl3 loading on the biomass surface and obtained biochar with SBET, Smeso, and VT of 664.55 m2 g−1, 92.76 m2 g−1, and 0.33 cm3 g−1, respectively. Yan et al. [18] prepared porous biochar with ZnCl2. SBET and Vmicro reached 852.41 m2 g−1 and 0.086 cm3 g−1, respectively. Metal salts are used as activators that could prepare biochar with more micropores. In addition, Idrees et al. [13] prepared biochar with high microporosity using KOH-activated peanut shells, and the adsorption capacity of CO2 was 5.32, 4.24, and 1.21 mmol g−1 at 0, 25, 30 °C, P = 1 bar, respectively. Microporous biochar has been used for CO2 adsorption at low temperatures and sewage treatment. However, the adsorption capacity of biochar for CO2 in flue gas is lower, and there are few studies on the realization of efficient adsorption of CO2 with biochar under high-temperature conditions. In order to achieve the effective adsorption of CO2 in SFG, it is necessary to research biochar modification.
Walnuts of world production are over 3.7 million t, and China is the largest producer, with 1.06 million tons. After walnut kernels are used for food purposes, the shell remains as a by-product or waste, although walnut shell biomass is also a renewable source of raw material [26]. Walnut shell has a low ash content and a high-quality natural structure, which is conducive to the formation of a developed pore structure and is an excellent raw material for the preparation of biochar.
In this study, FeCl3, Mg(NO3)2, and H2O (g) simultaneously activated and pyrolyzed walnut shells to prepare more mesoporous biochar. Different proportions of TEPA modified biochar to obtain solid amine adsorbents; subsequently, biochar and the adsorbent’s structural properties and thermal stability were characterized. The adsorption capacity and cycling stability of the adsorbents were measured under different SFG. Moreover, the adsorption and deactivation mechanisms of the adsorbents were examined.

2. Experimental Section

2.1. Materials

Walnut shells were obtained from a local market (Yinchuan, Ningxia, China); FeCl3·6H2O and Mg(NO3)2·6H2O (AR) were purchased from Sigma-Aldrich (St. Louis, MO, USA); analytically pure HCl (AR, hydrochloric acid 36.0 wt.%) was obtained from Sinopharm Chemical Reagent; analytically pure absolute ethanol (AR, 99.7 wt.%) was obtained from Damao Chemical Reagent Factory, Tianjin, China; high-purity CO2 and N2, (99.999%) were obtained from Ningxia Guangli Gas Co, Yinchuan, China. The elemental and industrial analyses of the walnut shells are shown in Table 1.

2.2. Synthesis of the Amine-Functionalized Sorbent

2.2.1. Preparation of Biochar

First, walnut shells were rinsed with water and dried at 100 °C. Then, they were crushed and sieved to form 50–100 mesh-sized particles. FeCl3·6H2O (3 g) and Mg(NO3)2·6H2O (6 g) were placed in 500 mL of deionized water and stirred until the solution was dispersed; 10 g of biomass as the carbon source was then added to the metal salts solution for impregnation, and the solution was then evaporated to dryness by stirring at 150 °C, and 400 r min−1. In a tube furnace, one-step pyrolysis and carbonization were performed in an inert atmosphere (H2O 0.1 mL min−1, N2 flow 200 mL min−1) at 900 °C for 100 min. After carbonization, H2O (g) was switched to pure N2 and cooled to a low temperature. The carbonization biochar was washed with 0.3 M HCl for 6 h to remove the metal oxides and a small amount of ash and then repeatedly washed using distilled water to neutralize it. Finally, the biochar was obtained by drying. Samples were named RBC (without activation), HBC (H2O (g) activated), FMBC (FeCl3, and Mg(NO3)2 coactivated) and CBC (FeCl3, Mg(NO3)2, and H2O (g) coactivated).

2.2.2. Preparation of Amine-Functionalized Biochar

TEPA was uniformly dispersed in anhydrous ethanol after the addition of the CBC. The solution was stirred with a heating plate at 35 °C and 200 r min−1 until it was completely evaporated. Finally, samples were dried (80 °C dry for 10 h under vacuum) to obtain the TEPA-loaded adsorbent. Samples were named CBC-X-TEPA, where X (X = 20, 30, 40, 50, and 60) refers to the mass fraction of TEPA.

2.3. Characterization

The ultimate analysis of samples was characterized by the elemental analyzer (EA 3000, Euro Vector, Pavia, Italy). The textural properties of the samples were measured using the method of low-temperature N2 adsorption–desorption with a rapid surface area and porosity analyzer (Autosorb-iQ, Quantachrome Instruments, Boynton Beach, FL, USA). The morphology features of samples were observed using scanning electron microscopy (SEM Quanta 400, FEI company, Hillsboro, OR, USA). The crystalline phase of samples was tested by X-ray diffractometer (XRD, Bruker, Germany) at Cu radiation, 40 kV, 40 mA, and 2θ range of 3°–85°. The samples graphitization was analyzed by Raman spectra (DXR2xi Raman, Thermo Fisher Scientific, Waltham, Massachusetts, U.S. wavelength range from 500 to 3500 cm−1). The functional groups on the surface of the samples were analyzed by using FTIR (TENSOR II, Bruker, Germany). Thermogravimetric analysis (TGA STA449 F3, NETZSCH-Gerätebau GmbH, Selb, Germany) was employed to analyze the thermal stability of samples.

2.4. CO2 Adsorption

CO2 sorption on adsorbent was performed in a breakthrough adsorption setup that is shown in Figure 1. The adsorption column (with length 40 cm and inner diameter 1 cm) was filled with adsorbent 1 g. In SFG (15% CO2/85% N2), the sorption CO2 capacity of adsorbent with different TEPA loaded was tested to select the best. Subsequently, CO2 sorption of the best adsorbent was performed in SFG (15% CO2/75% N2/10% H2O). The specific steps of the adsorption experiment were as follows: N2 (30 mL/min) was aerated in the adsorption bed, and the temperature was slowly heated to 105 °C and kept for 1 h. Then, the temperature of the adsorption bed was lowered to a constant temperature of 60 °C; the adsorption experiment was started by introducing SFG into the adsorption bed. When the outlet and inlet gases had the same CO2 concentration, the experiment was stopped. For the recycle experiment, the absorbent was exposed to SFG (30 mL min−1) at 60 °C to adsorb CO2. When the adsorption reaction ended, the high-purity N2 (30 mL min−1) was aerated in the adsorption bed, and the temperature was heated to 105 °C and maintained for 1 h. The test was repeated ten times. The adsorption capacity of the adsorbent was calculated using the following equation:
q = 1 M × [ ( 0 t Q × C 0 C 1 C ) ] × T 0 T × 1 V m
where q represents the adsorption capacity of CO2, mmol g−1; M is the quality of the CBC-X-TEPA, g; t is adsorption time, min. Q is the total SFG flow, cm3 min−1; C0 and C are the input and output CO2 concentrations; Vol.%; t is the adsorption time, min; T0 represents absolute zero, 273 K; T is adsorption temperature, K; and Vm is the standard molar volume, 22.4 L mol−1.
Amine (TEPA) efficiency was employed to investigate the utilization efficiency of organic amine loaded into biochar and the adsorption capacity of the adsorbent for CO2. The calculation formula of amine efficiency was as follows:
T E P A   e f f i c i e n c y = n CO 2 m a m i n e
where nCO2 is the saturated adsorption capacity of the adsorbent, mmol g−1; mamine refers to the quality of amine, g.

3. Results and Discussion

3.1. Biochar Preparation Mechanism

FeCl3, Mg(NO3)2, and H2O (g) played important roles in biomass pyrolysis. In the heating process, FeCl3·6H2O pyrolysis and gradual transformation into Fe2O3 and Fe3O4 are shown in Equations (3)–(6) [25]. Mg (NO3)2·6H2O decomposed into MgO after dehydration in the pyrolysis reaction process. The reactions are shown in Equations (7) and (8) [24]. MgO and Fe3O4 formed the MgO and FeO solid solution during pyrolysis. H2O (g) reacted with the surface carbon of biochar generated after biomass pyrolysis and then gradually diffused into the internal pores of biochar as the reaction proceeded, and the reactions formulae of H2O (g) and C are shown in Equations (9)–(11) [23].
The XRD patterns were employed to investigate the metal-conversion process during the preparation of biochar. The specific reaction is shown in Figure S1. During the initial 10 min pyrolysis stage, FeCl3·6H2O and Mg(NO3)2·6H2O first pyrolyzed into Fe2O3/Fe3O4 and MgO. Fe2O3/Fe3O4 reduced to FeO and Fe as shown in Equations (12) and (13) [19,25], then they sintered and agglomerated as MgO(0.91)FeO(0.09). When the reaction proceeded to 20 min and 40 min, MgO(0.91)FeO(0.09) transformed into MgO(0.77)FeO(0.23). At 60 min, H2O (g) introduction made the FeO of MgO(0.77)FeO(0.23) completely oxidize to Fe3O4; the reactions are shown in Equations (14) and (15) [27]. In addition, the sintered metal oxides separated from each other. The reaction mechanism is shown in Figure 2. Since the metal oxides were distributed on the biochar surface after pyrolysis, they could prevent the H2O (g) from entering the pores and hence prevent it from overreacting and collapsing the internal pore. In addition, the carbon peak in the biochar gradually weakened, which is attributed to high temperatures forming a more disordered structure [22].
FeCl 3 · 6 H 2 O FeCl 3 · 3 H 2 O + 3 H 2 O
FeCl 3 · 3 H 2 O Fe ( OH ) 3 + 3 HCl
Fe ( OH ) 3 FeO ( OH ) Fe 2 O 3
3 Fe 2 O 3 + 4 H 2 ( CO ,   C ) 2 Fe 3 O 4 + 4 H 2 O ( CO 2 ,   CO )
Mg ( NO 3 ) 2 · 6 H 2 O Mg ( NO 3 ) 2 + 6 H 2 O
Mg ( NO 3 ) 2 MgO + 2 NO 2 + 1 2 O 2
3 Fe + 4 H 2 O ( g ) Fe 3 O 4 + 4 H 2
C + H 2 O ( g ) CO + H 2
CO + H 2 O ( g ) CO 2 + H 2
Fe 3 O 4 + H 2 ( CO ,   C ) 3 FeO + H 2 O ( CO 2 ,   CO )
Fe 3 O 4 + 4 H 2 ( CO ,   C ) 3 Fe + 4 H 2 O ( CO 2 ,   CO )
3 FeO + H 2 O ( g ) Fe 3 O 4 + H 2
CO 2 + C 2 CO

3.2. Sample Characterization

The SEM characterization of samples was carried out to investigate the effect of the preparation method on the structure and morphology. Figure 3a shows the RBC without metal salts, and H2O (g) activation exhibited macropores, with an undeveloped pore structure and no lamellar structure. Figure 3b shows results after being coactivated with metal salts and H2O (g); the surface of CBC possessed distributed metal particles, which were produced by FeCl3 and Mg(NO3)2. Figure 3c shows results after the removal of these metal particles. The CBC surface showed a stacked lamellar structure with multiple large pores, which was different from what was roughly observed with irregular crakes, plates, and particles reported by Guo et al. [25]. The large pores on the surface were partly produced by the spillage of the volatile fraction during thermal decomposition and partly formed by the removal of large-sized metal particles sintered on the biochar surface. Compared with RBC, the surface structure and morphology of the CBC had undergone tremendous changes, which showed that metal salts and H2O (g) had a great influence on the formation of biochar. This could be more conducive to TEPA loading and CO2 adsorption on the surface of biochar. Figure 3d is an SEM image of the adsorbent that shows the stacked lamellar structure and surface pores on the CBC did not change significantly after the TEPA loading. The presence of pores of different sizes on the adsorbent surface could provide large passages, accelerating CO2 diffusion into the internal part of the adsorbent.
XRD is one of the best techniques to characterize the effect of the preparation method on the crystalline biochar phase. The XRD patterns of samples are shown in Figure 4a. The RBC and HBC had broad and weak diffraction peaks at 2θ = 25.8° and 42.8°, corresponding to the graphite (002) and (100) crystal face diffraction of amorphous carbon [28,29,30,31]. The intensity of the carbon peak gradually strengthened for FMBC (compared with RBC) after being added to FeCl3 and Mg(NO3)2. This confirmed that the addition of metal salts formed a large and orderly crystal structure. CBC had a more intense carbon peak than FMBC, which was attributed to FeO being continuously oxidized to Fe3O4 by H2O (g), making the reaction between carbon and Fe3O4 continue. At the same time, it increased the degree of graphitization and the thermal conductivity of biochar and sped up the CO2 adsorption and desorption processes of biochar. The XRD peak intensity of carbon weakened for CBC-50-TEPA (compared with CBC). This indicated TEPA was impregnated into the pores of CBC. In addition, TEPA impregnation induced a structure in which the edges or sides of a stacked, layered, and connected structure were destroyed so that the XRD pattern peak intensity weakened [30]. After TEPA incorporation, there was a bulging peak in front of the (002) crystal face peak, and the half peak width widened. This confirmed that the amine impregnation prevented the restacking of a part of the CBC’s lamellar structure [31].
Figure 4b shows the Raman spectra of the samples. The RBC had a weak G peak at 1570 cm−1, a D peak at 1340 cm−1, and no 2D peaks at 2670 cm−1. The HBC had intensity G and D peaks. However, the 2D peak weakened. FMBC had high-intensity D and G peaks and a 2D peak. CBC spectra had both high-intensity G and D peaks and weak 2D peaks. The symmetrical stretching vibration of the sp2 carbon atom in the aromatic ring produced the D peak, which represents the purity and defects of the graphite structure [32]. The tensile vibration between the carbon atoms of sp2 caused the appearance of the G peak. The double resonance transition of two phonons of carbon atoms with opposite momentum produced 2D peaks, reflecting the number of layers of carbon [33]. The Raman spectra of CBC-50-TEPA were similar to that of the CBC, indicating that the structure was unchanged after TEPA modification, which agreed with SEM results. The ratio of D to G (ID: IG) is an essential parameter for investigating carbon graphitization. The ID: IG of the RBC, HBC, FMBC, CBC, and CBC-50-TEPA were calculated as 2.62, 2.42, 1.86, 1.65, and 1.46, respectively.
Figure 5a shows adsorption–desorption isotherms of samples. According to the IUPAC classification, hysteresis loops appeared in the adsorption–desorption isotherm so that mesopores were present in all samples. Figure 5b shows that the pores of the CBC were mainly mesopores, primarily distributed between 4 and 20 nm, whereas few were micropores. There were primarily micropores and some mesopores regarding the RBC, HBC, and FMBC. The TEPA blocked micropores in the adsorbent w, but some large mesopores were retained.
Table 2 lists the specific pore properties of the RBC, HBC, FMBC, CBC, and CBC-50-TEPA. Comparing CBC with other biochars, SBET slightly increased. However, VT, Smeso, and Vmeso significantly increased, which indicated the high efficiency of combining H2O (g) with bimetal salts to change the pore structure. According to Sun et al. [34], the pore channels of biochar collapsed at high temperatures, thereby increasing both the VT and Pd of biochar, whereas the corresponding SBET and the number of pores decreased. However, in this study, while effectively increasing the VT and Pd of CBC, the SBET was also increased. This was due to the new pores created by metal salts and H2O (g), and it removed the Fe3O4, while MgO particles formed the stacked lamellar structure. The SBET and VT of CBC had a positive effect on the CO2 adsorption capacity of the solid amine adsorbent, and the Smeso and Vmeso are more vital than other pores [35]. In addition, compared to the reported results [11,12,13,14,15,16,17], when biochar had a large specific surface area, it also had a small pore volume and average pore size. In the process of amine modification, the micropores are easily blocked by organic amine molecules. Therefore, microporous biochar is not suitable for organic amine modification.
Carbon content in CBC was extremely high, whereas the N, H, and O contents were low (Table 3). During pyrolysis, a dehydration reaction reduced the H: C and O: C ratios, and the breakdown of macromolecular organic nutrients into smaller molecules occurred, which spilled over along with volatile fractions and decreased the N content in the CBC at 900 °C [36]. In addition, the N content of CBC-50-TEPA (7.34%) was approximately 28.6 times that of CBC (0.47%), which indicated that a mass of amine was attached to the CBC during functionalization. This further confirmed the effectiveness of TEPA as an ample origin of N (37%) with high amino density. According to the N content of elemental analysis, the actual loading of TEPA was calculated as 35.0% less than 50.0%, which was because, during TEPA immersion, some pores of the CBC were blocked, preventing the amine from continuing to diffuse deeply into the interior. In addition, the O content in the adsorbent increased because of the adsorption of CO2 and H2O by the samples.
The FTIR spectra of CBC and CBC-50-TEPA are shown in Figure S2. Regarding CBC, the peak at 1585 cm−1 was the C=C vibration peak [37,38]. There were no other oxygen-and nitrogen-containing groups, indicating low N and O contents, which agreed with the elemental analysis. After TEPA modification, the stretching vibration peak of primary amine (−NH2) appeared at 3292 cm−1. Peaks at 2821 and 2930 cm−1 were CH2 stretching vibrations, whereas C–N stretching vibrations were observed at 1308 cm−1 [39,40]. The newly added N groups in the CBC showed that TEPA was successfully loaded.
Figure 6 shows the CBC and CBC-50-TEPA TG analyses. Figure 6a shows that the CBC had almost no weight loss in the range of 30–500 °C, which is attributed to the organic matter in the biochar having been volatilized and no organic matter left for decomposition at high temperatures. So, the CBC has excellent thermal stability at high temperatures. There are two weight-loss stages in the curve of the CBC-50-TEPA. The results showed that CBC-50-TEPA only lost 4.7% of its weight within 150 °C. Therefore, it is suitable for CO2 capture in SFG (50–100 °C). The DTG curves of CBC-50-TEPA and CBC are shown in Figure 6b,c. CBC-50-TEPA has two weight-loss peaks through 38–500 °C and has a wide temperature range of weight loss. CBC-50-TEPA weight loss in the range of 38–105 °C is mainly due to CO2 and H2O heat volatilization, and weight loss within 105–500 °C is caused by pyrolysis of TEPA. CBC DTG curve has one weight loss peak at 95–103 °C. The excellent heat resistance of the biochar and aminated biochar makes it potentially applicable to CO2 adsorption in SFG.

3.3. CO2 Adsorption–Desorption

The sorption of CO2 was performed on a breakthrough adsorption setup. In dry SFG containing 15% CO2/N2 85%, the adsorption capacity of CO2 was tested for CBC-X-TEPA adsorbent (at 101 kPa, 60 °C). Figure 7a,b shows breakthrough adsorption curves and CO2 adsorption capacities of the adsorbent, which gradually increased as the TEPA loading increased from 20% to 50%; the adsorption time extended, as shown in Figure 7a, and reached a maximum value of 2.82 mmol g−1 at 50% TEPA, as shown in Figure 7b. The large pore volume of the CBC provided a large space for the loading of TEPA, which could allow a sufficient amount of amine to penetrate the pore and provide more adsorption sites for CO2 adsorption.
The adsorption capacity started to decrease at 60% TEPA loading, which was because the pores of the adsorbent were blocked so that CO2 only reacted with the amine-based active sites loaded on the surface of the CBC. In contrast, the organic amines immersed in the pores could not function and reduced CO2 adsorption capacity. When the amine efficiency was 8.05 mmol g−1, the adsorption capacity of the adsorbent began to decrease, and the adsorbent reached the optimal load of the amine group.
The adsorption capacity of biochar for CO2 at 0 °C and 60 °C was also tested. At low temperatures (0 °C, pure CO2), biochar with high microporosity has a high physical adsorption capacity for CO2, 4.17 mmol g−1. At 60 °C, the capacity of biochar with high microporosity to adsorb CO2 is low, only 0.76 mmol g−1, and micropores are no longer the dominant factor for CO2 adsorption. The results of pore-size analysis showed that the micropores in the biochar after TEPA modification were greatly reduced, and more mesopores and macropores were retained. At 60 °C, the CO2 adsorption experiment results show that the adsorbent has a high CO2 adsorption capacity, mainly due to the chemical reaction between the amine group and CO2, and the adsorption at this time is mainly chemical adsorption. It can be seen that the effect of micropores on chemisorption is not significant. In general, in the adsorption of CO2 at 60 °C, the adsorbent CBC-50-TEPA was mainly used for chemical adsorption. At the same time, there was also some physical adsorption caused by biochar CBC.
In addition, the adsorption capacity of CBC-50-TEPA was tested using 10% H2O (g) SFG [15% CO2/75% N2/10% H2O (g)]. The results are shown in Figure 8. Remarkably, it reached 3.31 mmol g−1, 0.49 mmol g−1 higher than in the dry SFG, as shown in Figure 8a. TEPA efficiency increased from 8.05 to 9.4 mmol g−1, as shown in Figure 8b, since the presence of H2O (g) changed the reaction mechanism of the amine with CO2. The reactions of the amino group with CO2 in the dry SFG are shown in Equations (16)–(18). When one CO2 molecule was consumed, two amino functional groups were also consumed [41,42]. In the 10% H2O (g) SFG, the reactions of the amino group with CO2 are shown in Equations (19) and (20) [42,43]. One CO2 molecule was absorbed, and one amino functional group was expended. Therefore, H2O (g) increased the adsorption efficiency of the amino-functional group and the adsorption capacity of CO2.
2 R 1 NH 2 + CO 2 ( R 1 NH 3 + ) ( R 1 NHCOO )  
2 R 1 R 2 NH 2 + CO 2 ( R 1 R 2 NH 3 + ) ( R 1 R 2 NHCOO )
R 1 NH 2 + R 1 R 2 NH + CO 2 ( R 1 NH 3 + ) ( R 1 R 2 NHCOO )
R 1 NH 2 + CO 2 + H 2 O ( R 1 NH 3 + ) ( HCO 3 )
R 1 R 2 NH + CO 2 + H 2 O ( R 1 R 2 NH 2 + ) ( HCO 3 )
The CO2 adsorption capacity of CBC-50-TEPA was compared with those of reported solid amine adsorbents in Table 4. CBC-50-TEPA exhibited an adsorption capacity (3.31 mmol g−1) equivalent to GO (graphene oxide), MCM-41, CNT, and MOFS functionalized by TEPA. This proved that the solid amine adsorbent with biochar as the carrier is also a kind of adsorbent, with potential advantages for removing CO2 from FG.

3.4. Cyclic Adsorption Experiment

The cycle stability determines the service life of the adsorbent. Figure 9 shows the cycle experiments (adsorption temperature: 60 °C, desorption temperature: 100 °C) of CBC-50-TEPA in the dry and 10% H2O (g) SFG. The experimental results show that the adsorption capacity of adsorbents decreased from the initial 2.82 to 0.86 mmol g−1 after ten cycles in dry SFG. It was not easy to regenerate the adsorption of the product just by changing the temperature. Under dry conditions, amine-group adsorption of CO2 generated extremely stable adsorption products of urea functional groups [43,45]. Whenever H2O (g) was present, the adsorption capacity of the adsorbent decreased by only 0.08 mmol g−1 from the initial 3.31 after ten cycles, which was almost unchanged. H2O (g) caused the reaction products between amino groups and CO2 to be transformed into carbamates [44,46]. It could be effectively regenerated by increasing the temperature, thus keeping the capacity of the adsorbent stable.
Figure 10 shows the FTIR spectra of the fresh adsorbent and the adsorbent circulating in the 10% H2O (g) SFG and dry SFG. A new absorption peak appeared at 1658 cm−1 after circulating in 10% H2O (g) SFG, which was the carbonyl peak in urea. Meanwhile, the N–H and C–H vibration peaks were located at 1570 cm−1 and 1460 cm−1, whose intensity weakened compared with the fresh adsorbent [47,48,49]. However, the adsorbent recycled in dry SFG had a lower peak at 1570 cm−1; the C–H vibration peak at 1460 cm−1 almost disappeared. The carbonyl vibration peak at 1658 cm−1 was further strengthened. FTIR spectra show that in dry SFG, CBC-50-TEPA formed many urea groups after ten cycles.
There are two forms of urea. The open-chain urea is attributed to the dehydration condensation of CO2 with the different amines in two molecules. The cyclic urea is attributed to the reaction of CO2 with two amines in the molecules [46,50]. The results indicated that 10% H2O (g) SFG could delay the CBC-50-TEPA formatted urea, thereby maintaining the stability of the adsorbent.

4. Conclusions

We developed an eco-friendly and cost-effective method to prepare biochar (CBC). At the same time, FeCl3, Mg(NO3)2, and H2O (g) activated and increased the ratio of biochar mesopores through a continuous redox reaction. The large VT (1.41 cm3 g−1) and SBET (744.38 m2 g−1) of biochar provided immersion space and channels for the loading of TEPA. The solid amine adsorbent (CBC-50-TEPA) prepared with biochar as a carrier exhibited good thermal stability and a high CO2 adsorption capacity (3.31 mmol g−1) in 10% H2O (g) SFG. The H2O (g) could delay the generation of urea groups, thus allowing the adsorbent to maintain high stability for ten cycles.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/atmos13040579/s1, Supporting information includes the XRD spectra of metal oxide during the reaction process, the FTIR spectra of the CBC and CBC-50-TEPA, and the formation process of the open-chain and cyclic urea. Figure S1. XRD spectra of Metal oxide during the reaction process, Figure S2. FTIR spectra of the CBC and CBC-50-TEPA.

Author Contributions

W.T.: conceived and designed the research; performed experimental testing; analyzed the data, and wrote the paper. Y.W.: contributed to the design of the experimental scheme and data analysis. J.H.: assisted with manuscript revision. T.G.: assisted with the experimental data analysis and modified the paper. X.W.: assisted with manuscript revision. X.X.: contributed to BET experiment and data analysis. Q.G.: provided constructive suggestions, manuscript revision, supervision, and funding support. All authors have read and agreed to the published version of the manuscript.

Funding

The Joint Funds of the National Natural Science Foundation of China (grant number U20A20124); the Major Project of the Key Research and Development Program of Ningxia Province, China (grant number 2018BCE01002); The National First-rate Discipline Construction Project of Ningxia (grant number NXYLXK2017A04); The National Natural Science Foundation of China (grant number 22108208); the Foundation of State Key Laboratory of High-efficiency Utilization of Coal and Green Chemical Engineering (grant number 2022–K27).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data that support the findings of this study are available from the corresponding author, Q.G., upon reasonable request.

Acknowledgments

The authors gratefully acknowledge the financial support from the Joint Funds of the National Natural Science Foundation of China (grant no. U20A20124), the Major Project of the Key Research and Development Program of Ningxia Province, China (2018BCE01002), the National First-rate Discipline Construction Project of Ningxia (NXYLXK2017A04), the National Natural Science Foundation of China (22108208), and the Foundation of State Key Laboratory of High-efficiency Utilization of Coal and Green Chemical Engineering (2022–K27).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wang, J.; Song, C.; Yuan, R. CO2 emissions from electricity generation in China during 1997–2040: The roles of energy transition and thermal power generation efficiency. Sci. Total Environ. 2021, 773, 145026. [Google Scholar] [CrossRef]
  2. Osman, A.I.; Hefny, M.; Abdel Maksoud, M.I.A.; Elgarahy, A.M.; Rooney, D.W. Recent advances in carbon capture storage and utilisation technologies: A review. Environ. Chem. Lett. 2021, 19, 797–849. [Google Scholar] [CrossRef]
  3. Yang, Z.X.; Guo, X.F.; Zhang, G.J.; Xu, Y. One–pot synthesis of high N–doped porous carbons derived from a N–rich oil palm biomass residue in low temperature for CO2 capture. Int. J. Energy Res. 2020, 44, 4875–4887. [Google Scholar] [CrossRef]
  4. Quan, C.; Su, R.R.; Gao, N.B. Preparation of activated biomass carbon from pine sawdust for supercapacitor and CO2 capture. Int. J. Energy Res. 2020, 44, 4335–4351. [Google Scholar] [CrossRef]
  5. Wang, X.; Chen, L.L.; Guo, Q.J. Development of hybrid amine–functionalized MCM–41 sorbents for CO2 capture. Chem. Eng. J. 2015, 260, 573–581. [Google Scholar] [CrossRef]
  6. Yang, L.Z.; Yan, L.T.; Wang, Y.; Liu, Z.; He, J.X.; Fu, Q.J.; Liu, D.D.; Gu, X.; Dai, P.C.; Li, L.J.; et al. Adsorption Site Selective Occupation Strategy within a Metal-Organic Framework for Highly Efficient Sieving Acetylene from Carbon Dioxide. Angew. Chem. Int. Ed. 2021, 60, 4570–4574. [Google Scholar] [CrossRef]
  7. Sharma, H.; Dhir, A. Capture of carbon dioxide using solid carbonaceous and non–carbonaceous adsorbents: A review. Environ. Chem. Lett. 2021, 19, 851–873. [Google Scholar] [CrossRef]
  8. Qi, Y.F.; Hu, X.D.; Liu, Y.Z.; Sun, D.S.; Li, R.H.; Zhu, H.T. Highly efficient and reversible absorption of SO2 by hydroxyl ammonium ionic liquids at low partial pressure. J. Chem. Technol. Biotechnol. 2019, 94, 3325–3332. [Google Scholar] [CrossRef]
  9. Dissanayake, P.D.; You, S.; Igalavithana, A.D.; Xia, Y.; Bhatnagar, A.; Gupta, S.; Kua, H.W.; Kim, S.; Kwon, J.; Tsang, D.C.W.; et al. Biochar–based adsorbents for carbon dioxide capture: A critical review. Renew. Sustain. Energy Rev. 2020, 119, 109582. [Google Scholar] [CrossRef]
  10. Yuan, T.; He, W.J.; Yin, G.J.; Xu, S.A. Comparison of bio–chars formation derived from fast and slow pyrolysis of walnut shell. Fuel 2020, 261, 116450. [Google Scholar] [CrossRef]
  11. Chatterjee, R.; Sajjadi, B.; Chen, W.; Mattern, D.L.; Hammer, N.; Raman, V.; Dorris, A. Effect of Pyrolysis Temperature on PhysicoChemical Properties and Acoustic–Based Amination of Biochar for Efficient CO2 Adsorption. Front. Energy Res. 2020, 218, 741–748. [Google Scholar] [CrossRef]
  12. Shahkarami, S.; Azargohar, R.; Dalai, A.K.; Soltan, J. Breakthrough CO2 adsorption in bio-based activated carbons. J. Environ. Sci. 2015, 34, 68–76. [Google Scholar] [CrossRef] [PubMed]
  13. Idrees, M.; Rangari, V.; Jeelani, S. Sustainable packaging waste-derived activated carbon for carbon dioxide capture. J. CO2 Util. 2018, 26, 380–387. [Google Scholar] [CrossRef]
  14. Villota, E.M.; Lei, H.W.; Qian, M.; Yang, Z.X.; Villota, S.M.A.; Zhang, Y.Y.; Yadavalli, G. Optimizing Microwave-Assisted Pyrolysis of Phosphoric Acid-Activated Biomass: Impact of Concentration on Heating Rate and Carbonization Time. ACS Sustain. Chem. Eng. 2018, 6, 1318–1326. [Google Scholar] [CrossRef]
  15. Demir, M.; Doguscu, M. Preparation of Porous Carbons Using NaOH, K2CO3, Na2CO3 and Na2S2O3 Activating Agents and Their Supercapacitor Application: A Comparative Study. Chemistryselect 2022, 7, 1–9. [Google Scholar] [CrossRef]
  16. Gao, B.; Wang, Y.; Huang, L.; Liu, S.M. Study on the performance of HNO3-modified biochar for enhanced medium temperature anaerobic digestion of food waste. Waste Manag. 2021, 135, 338–346. [Google Scholar] [CrossRef] [PubMed]
  17. Liu, S.S.; Wu, G.; Syed-Hassan, S.S.A.; Li, B.; Hu, X.; Zhou, J.B.; Huang, Y.; Zhang, S.; Zhang, H. Catalytic pyrolysis of pine wood over char-supported Fe: Bio-oil upgrading and catalyst regeneration by CO2/H2O. Fuel 2022, 307, 121778. [Google Scholar] [CrossRef]
  18. Yan, L.L.; Liu, Y.; Zhang, Y.D.; Liu, S.; Wang, C.X.; Chen, W.T.; Liu, C.; Chen, Z.L.; Zhang, Y. ZnCl2 modified biochar derived from aerobic granular sludge for developed microporosity and enhanced adsorption to tetracycline. Bioresour. Technol. 2020, 297, 122381. [Google Scholar] [CrossRef] [PubMed]
  19. Liu, W.J.; Tian, K.; He, Y.R.; Jiang, H.; Yu, H.Q. High–Yield Harvest of Nanofibers/Mesoporous Carbon Composite by Pyrolysis of Waste Biomass and Its Application for High Durability Electrochemical Energy Storage. Environ. Sci. Technol. 2014, 48, 13951–13959. [Google Scholar] [CrossRef]
  20. Chen, S.S.; Cao, Y.; Tsang, D.C.W.; Tessonnier, J.P.; Shang, J.; Hou, D.Y.; Shen, Z.T.; Zhang, S.C.; Ok, Y.S.; Wu, K.C.W. Effective Dispersion of MgO Nanostructure on Biochar Support as a Basic Catalyst for Glucose Isomerization. ACS Sustain. Chem. Eng. 2020, 8, 6990–7001. [Google Scholar] [CrossRef]
  21. Diedhiou, A.; Ndiaye, L.; Bensakhria, A.; Sock, O. Thermochemical conversion of cashew nut shells, palm nut shells and peanut shells char with CO2 and/or steam to aliment a clay brick firing unit. Renew. Energy 2019, 142, 581–590. [Google Scholar] [CrossRef]
  22. Fu, J.P.; Zhang, J.R.; Jin, C.J.; Wang, Z.Q.; Wang, T.; Cheng, X.X.; Ma, C.Y. Effects of temperature, oxygen and steam on pore structure characteristics of coconut husk activated carbon powders prepared by one–step rapid pyrolysis activation process. Bioresour. Technol. 2020, 310, 123413. [Google Scholar] [CrossRef] [PubMed]
  23. Feng, D.D.; Zhang, Y.; Zhao, Y.J.; Sun, S.Z.; Wu, J.Q.; Tan, H.P. Mechanism of in–situ dynamic catalysis and selective deactivation of H2O–activated biochar for biomass tar reforming. Fuel 2020, 279, 118450. [Google Scholar] [CrossRef]
  24. Lahijani, P.; Mohammadi, M.; Mohamed, A.R. Metal incorporated biochar as a potential adsorbent for high capacity CO2 capture at ambient condition. J. CO2 Util. 2018, 26, 281–293. [Google Scholar] [CrossRef]
  25. Guo, F.Q.; Jiang, X.C.; Li, X.L.; Jia, X.P.; Liang, S.; Qian, L. Synthesis of MgO/Fe3O4 nanoparticles embedded activated carbon from biomass for high–efficient adsorption of malachite green. Mater. Chem. Phys. 2020, 240, 122240. [Google Scholar] [CrossRef]
  26. Jovičić, N.; Antonović, A.; Matin, A.; Antolović, S.; Kalambura, S.; Krička, T. Biomass Valorization of Walnut Shell for Liquefaction Efficiency. Energies 2022, 15, 495. [Google Scholar] [CrossRef]
  27. Fan, L.S.; Li, F.X.; Ramkumar, S. Utilization of chemical looping strategy in coal gasification processes. Particuology 2008, 6, 131–142. [Google Scholar] [CrossRef]
  28. Keiluweit, M.; Nico, P.S.; Johnson, M.G.; Kleber, M. Dynamic Molecular Structure of Plant Biomass–Derived Black Carbon (Biochar). Environ. Sci. Technol. 2010, 44, 1247–1253. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Kim, K.H.; Kim, J.; Cho, T.; Choi, J.W. Influence of pyrolysis temperature on physicochemical properties of biochar obtained from the fast pyrolysis of pitch pine (Pinus rigida). Bioresour. Technol. 2012, 118, 158–162. [Google Scholar] [CrossRef] [PubMed]
  30. Sui, Z.Y.; Cui, Y.; Zhu, J.H.; Han, B.H. Preparation of Three–Dimensional Graphene Oxide–Polyethylenimine Porous Materials as Dye and Gas Adsorbents. ACS Appl. Mater. Interfaces 2013, 5, 9172–9179. [Google Scholar] [CrossRef] [PubMed]
  31. Cai, D.Y.; Song, M. Preparation of fully exfoliated graphite oxide nanoplatelets in organic solvents. J. Mater. Chem. A 2007, 17, 3678. [Google Scholar] [CrossRef]
  32. Liu, Y.M.; Sajjadi, B.; Chen, W.Y.; Chatterjee, R.J. Ultrasound–assisted amine functionalized graphene oxide for enhanced CO2 adsorption. Fuel 2019, 247, 10–18. [Google Scholar] [CrossRef]
  33. Tuinstra, F.; Koenig, J.L. Raman Spectrum of Graphite. J. Chem. Phys. 1970, 53, 1126–1130. [Google Scholar] [CrossRef] [Green Version]
  34. Sun, L.; Tian, C.G.; Wang, L.; Zou, J.L.; Mu, G.; Fu, H.G. Magnetically separable porous graphitic carbon with large surface area as excellent adsorbents for metal ions and dye. J. Mater. Chem. A 2011, 21, 7232. [Google Scholar] [CrossRef]
  35. Gibson, J.A.A.; Gromov, A.V.; Brandani, S.; Campbell, E.E.B. The effect of pore structure on the CO2 adsorption efficiency of polyamine impregnated porous carbons. Microporous Mesoporous Mater. 2015, 208, 129–139. [Google Scholar] [CrossRef] [Green Version]
  36. Guo, Y.; Pan, X.; Zhu, Q.; Ma, J.; Guo, Q. Effect of Biomass-Based Catalyst in Walnut Shell/Polypropylene to BTX. Fine Chem. Eng. 2021, 3, 11–28. [Google Scholar] [CrossRef]
  37. Alabadi, A.; Razzaque, S.; Yang, Y.; Chen, S.; Tan, B. Highly porous activated carbon materials from carbonized biomass with high CO2 capturing capacity. Chem. Eng. J. 2015, 281, 606–612. [Google Scholar] [CrossRef]
  38. Karakaş, C.; Özçimen, D.; İnan, B. Potential use of olive stone biochar as a hydroponic growing medium. J. Anal. Appl. Pyrol. 2017, 125, 17–23. [Google Scholar] [CrossRef]
  39. Kong, Q.P.; Wei, J.Y.; Hu, Y.; Wei, C.H. Fabrication of terminal amino hyperbranched polymer modified graphene oxide and its prominent adsorption performance towards Cr(VI). J. Hazard. Mater. 2019, 363, 161–169. [Google Scholar] [CrossRef]
  40. Dao, D.S.; Yamada, H.; Yogo, K. Enhancement of CO2 Adsorption/Desorption Properties of Solid Sorbents Using Tetraethylenepentamine/Diethanolamine Blends. ACS Omega 2020, 5, 23533–23541. [Google Scholar] [CrossRef] [PubMed]
  41. Sardo, M.; Afonso, R.; Juźków, J.; Pacheco, M.; Bordonhos, M.; Pinto, M.L.; Gomes, J.R.B.; Mafra, L. Unravelling moisture–induced CO2 chemisorption mechanisms in amine–modified sorbents at the molecular scale. J. Mater. Chem. A 2021, 9, 5542–5555. [Google Scholar] [CrossRef] [PubMed]
  42. Wang, Y.X.; Guo, T.; Hu, X.D.; Hao, J.; Guo, Q.J. Mechanism and kinetics of CO2 adsorption for TEPA- impregnated hierarchical mesoporous carbon in the presence of water vapor. Powder Technol. 2020, 368, 227–236. [Google Scholar] [CrossRef]
  43. Numaguchi, R.; Fujiki, J.; Yamada, H.; Chowdhury, A.; Kida, K.; Goto, K.; Okumura, T.; Yoshizawa, K.; Yogo, K. Development of Post–combustion CO2 Capture System Using Amine–impregnated Solid Sorbent. Energy Procedia 2017, 114, 2304–2312. [Google Scholar] [CrossRef]
  44. Wang, Y.X.; Hu, X.D.; Guo, T.; Tian, W.G.; Hao, J.; Guo, Q.J. The competitive adsorption mechanism of CO2, H2O and O2 on a solid amine adsorbent. Chem. Eng. J. 2021, 416, 129007. [Google Scholar] [CrossRef]
  45. Choe, J.H.; Kang, D.W.; Kang, M.; Kim, H.; Park, J.R.; Kim, D.W.; Hong, C.S. Revealing an unusual temperature–dependent CO2 adsorption trend and selective CO2 uptake over water vapors in a polyamine–appended metal–organic framework. Mater. Chem. Front. 2019, 3, 2759–2767. [Google Scholar] [CrossRef]
  46. Zhao, X.L.; Cui, Q.; Wang, B.D.; Yan, X.L.; Singh, S.; Zhang, F.; Gao, X.; Li, Y.L. Recent progress of amine modified sorbents for capturing CO2 from flue gas. Chin. J. Chem. Eng. 2018, 26, 2292–2302. [Google Scholar] [CrossRef]
  47. Ma, L.; Bai, R.; Hu, G.; Chen, R.; Hu, X.; Dai, W.; Dacosta, H.F.M.; Fan, M. Capturing CO2 with Amine–Impregnated Titanium Oxides. Energy Fuels 2013, 27, 5433–5439. [Google Scholar] [CrossRef]
  48. Hladnik, L.; Vicente, F.A.; Novak, U.; Grilc, M.; Likozar, B. Solubility assessment of lignin monomeric compounds and organosolv lignin in deep eutectic solvents using in situ Fourier–transform infrared spectroscopy. Ind. Crop. Prod. 2021, 164, 113359. [Google Scholar] [CrossRef]
  49. Li, K.M.; Jiang, J.G.; Chen, X.J.; Gao, Y.C.; Yan, F.; Tian, S.C. Research on Urea Linkages Formation of Amine Functional Adsorbents During CO2 Capture Process: Two Key Factors Analysis, Temperature and Moisture. J. Phys. Chem. C 2016, 120, 25892–25902. [Google Scholar] [CrossRef]
  50. Zhao, J.; Deng, S.; Zhao, L.; Yuan, X.Z.; Du, Z.Y.; Li, S.J.; Chen, L.J.; Wu, K.L. Understanding the effect of H2O on CO2 adsorption capture: Mechanism explanation, quantitative approach and application. Sustain. Energy Fuels 2020, 4, 5970–5986. [Google Scholar] [CrossRef]
Figure 1. Schematic of CO2 breakthrough adsorption setup.
Figure 1. Schematic of CO2 breakthrough adsorption setup.
Atmosphere 13 00579 g001
Figure 2. The biochar preparation mechanism.
Figure 2. The biochar preparation mechanism.
Atmosphere 13 00579 g002
Figure 3. SEM images of (a) RBC, (b) CBC before removal of metal particles, (c) CBC after removal of metal particles, and (d) CBC-50-TEPA.
Figure 3. SEM images of (a) RBC, (b) CBC before removal of metal particles, (c) CBC after removal of metal particles, and (d) CBC-50-TEPA.
Atmosphere 13 00579 g003
Figure 4. XRD spectra (a), Raman spectra of biochar from different activation methods, and CBC-50-TEPA (b).
Figure 4. XRD spectra (a), Raman spectra of biochar from different activation methods, and CBC-50-TEPA (b).
Atmosphere 13 00579 g004
Figure 5. (a) Nitrogen adsorption−desorption isotherms and (b) pore diameter of biochar from different activation methods and CBC-50-TEPA.
Figure 5. (a) Nitrogen adsorption−desorption isotherms and (b) pore diameter of biochar from different activation methods and CBC-50-TEPA.
Atmosphere 13 00579 g005
Figure 6. TG curves of weight losses and weight loss rates of pyrolysis and catalytic pyrolysis: (a) TGA CBC and CBC-50-TEPA, (b) derivative TG (DTG) of CBC-50-TEPA, and (c) DTG of CBC.
Figure 6. TG curves of weight losses and weight loss rates of pyrolysis and catalytic pyrolysis: (a) TGA CBC and CBC-50-TEPA, (b) derivative TG (DTG) of CBC-50-TEPA, and (c) DTG of CBC.
Atmosphere 13 00579 g006
Figure 7. Breakthrough adsorption curves of CO2 (a), and CO2 adsorption capacity and TEPA efficiencies of CBC-X-TEPA (b).
Figure 7. Breakthrough adsorption curves of CO2 (a), and CO2 adsorption capacity and TEPA efficiencies of CBC-X-TEPA (b).
Atmosphere 13 00579 g007
Figure 8. (a) Breakthrough adsorption curves and (b) adsorption capacity of CBC-50-TEPA in different SFGs.
Figure 8. (a) Breakthrough adsorption curves and (b) adsorption capacity of CBC-50-TEPA in different SFGs.
Atmosphere 13 00579 g008
Figure 9. Ten cycles of adsorption capacity of CBC-50-TEPA in different SFGs.
Figure 9. Ten cycles of adsorption capacity of CBC-50-TEPA in different SFGs.
Atmosphere 13 00579 g009
Figure 10. FTIR spectra of the CBC-50-TEPA after 10 cycles in different SFGs.
Figure 10. FTIR spectra of the CBC-50-TEPA after 10 cycles in different SFGs.
Atmosphere 13 00579 g010
Table 1. Industrial analysis and elemental analysis of walnut shell.
Table 1. Industrial analysis and elemental analysis of walnut shell.
SamplesElemental Analysis (wt%, Ad)Industrial Analysis (wt%, Ad)
Walnut shellCHO aNSMAVFC
49.945.8643.960.200.0436.211.8681.2310.69
Note: a indicates the use of differential subtraction; Ad: air-dried basis. M: moisture; A: ash; V: volatile; FC: fixed-carbon.
Table 2. Porous texture properties of biochar from different activation methods and CBC-50-TEPA.
Table 2. Porous texture properties of biochar from different activation methods and CBC-50-TEPA.
SamplesSBET/(m2 g−1) aVT (cm3 g−1) bSmeso/(m2 g−1) cVmeso (cm3 g−1) dDA/(nm) eRef.
RBC667.420.3380.150.091.98present work
HBC998.500.73239.090.312.91present work
FMBC500.180.45206.760.313.61present work
CBC744.381.41425.471.237.57present work
CBC-50-TEPA23.670.1216.440.1119.39present work
US-MS 7005880.21Chatterjee et al. [11]
Steam AC8400.553.77Sepideh et al. [12]
CHC-O6S20333.810.153.77Fu et al. [22]
H3PO4-81.211725.71.541.533.72Villota et al. [14]
C-Na2CO37080.3Demir et al. [15]
BCO25.3Gao et al. [16]
Char-Fe517.960.332.61Liu et al. [17]
a SBET was obtained by the BET method. b VT at p/p0 = 0.99. c Mesoporous surface area was acquired by the BJH method. d Mesoporous pore volume was acquired by the BJH method. e Average pore diameter.
Table 3. Elemental analysis of CBC and CBC-50-TEPA (wt%).
Table 3. Elemental analysis of CBC and CBC-50-TEPA (wt%).
SampleCHON
CBC94.5090.6600.1880.469
BC-50-TEPA71.5323.8934.4297.34
Table 4. Comparison of adsorption capacities of CBC-50-TEPA in past research.
Table 4. Comparison of adsorption capacities of CBC-50-TEPA in past research.
AdsorbentActivating AgentTemp. °CCO2 Partial Pressure (bar)Adsorption Capacities (mmol g−1)Ref.
CBCTEPA600.153.31Present work
GOUS-TEPA700.11.20Liu et al. [32]
MCM–41TEPA700.152.45Wang et al. [5]
MOFTEPA600.152.00Quan et al. [4]
CNTPEI600.154.75Wang et al. [44]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Tian, W.; Wang, Y.; Hao, J.; Guo, T.; Wang, X.; Xiang, X.; Guo, Q. Amine-Modified Biochar for the Efficient Adsorption of Carbon Dioxide in Flue Gas. Atmosphere 2022, 13, 579. https://doi.org/10.3390/atmos13040579

AMA Style

Tian W, Wang Y, Hao J, Guo T, Wang X, Xiang X, Guo Q. Amine-Modified Biochar for the Efficient Adsorption of Carbon Dioxide in Flue Gas. Atmosphere. 2022; 13(4):579. https://doi.org/10.3390/atmos13040579

Chicago/Turabian Style

Tian, Wengang, Yanxia Wang, Jian Hao, Tuo Guo, Xia Wang, Xiaoju Xiang, and Qingjie Guo. 2022. "Amine-Modified Biochar for the Efficient Adsorption of Carbon Dioxide in Flue Gas" Atmosphere 13, no. 4: 579. https://doi.org/10.3390/atmos13040579

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop