2014 - CO2 Capture Using Biochar Produced From Sugarcane Bagasse and Hickory Wood

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 27

Accepted Manuscript

Carbon dioxide capture using biochar produced from sugarcane bagasse and
hickory wood

Anne Elise Creamer, Bin Gao, Ming Zhang

PII: S1385-8947(14)00394-5
DOI: http://dx.doi.org/10.1016/j.cej.2014.03.105
Reference: CEJ 11964

To appear in: Chemical Engineering Journal

Received Date: 20 February 2014


Revised Date: 24 March 2014
Accepted Date: 26 March 2014

Please cite this article as: A.E. Creamer, B. Gao, M. Zhang, Carbon dioxide capture using biochar produced from
sugarcane bagasse and hickory wood, Chemical Engineering Journal (2014), doi: http://dx.doi.org/10.1016/j.cej.
2014.03.105

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. 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.
Carbon dioxide capture using biochar produced from sugarcane bagasse and hickory wood

Anne Elise Creamer, Bin Gao*, Ming Zhang

Department of Agricultural and Biological Engineering, University of Florida, Gainesville,

FL32611

______________
*
Corresponding author, phone: (352) 392-1864 ext. 285, email: bg55@ufl.edu

1
Abstract

Anthropogenic CO2 emissions continue to climb due to increases in global energy

demand. Because power plants are the largest stationary source of anthropogenic CO2,

developing effective and affordable post combustion CO2 capture technology has attracted

substantial research attention. This study assessed the adsorption of CO2 onto biochar, a low-cost

adsorbent that can be produced from waste biomass through low-temperature pyrolysis.

Sugarcane bagasse (BG) and hickory wood (HW) feedstock converted into biochar at 300, 450,

and 600 oC. Sorption of CO2 on each of the resulted biochar was measured by monitoring its

weight changes in CO2 at 25 or 75 oC. The biochar was found to be effective for CO2 capture and

the adsorption process could be described by the second-order kinetics model. In general, the

biochars produced at higher temperature had better CO2 capture performance. The BG biochar

produced at 600 oC showed the most adsorption of CO2 (73.55 mg g-1 at 25 oC). However, even

when the feedstock was exposed to only 300 oC pyrolysis, the biochar was still able to capture

more than 35 mg g-1 CO2 at 25 oC. Experimental results suggest that CO2 weakly bound to the

surface of the biochar through physisorption, so surface area was a significant determinant of

CO2 adsorption (Pearson’s r = 0.82); nevertheless, the presence of nitrogenous groups also

played a role, when the surface area was sufficient. Biochar’s porous structure and unique

surface properties enable it to be an efficient CO2 adsorbent, while being sustainable and

inexpensive.

Keywords: biochar; carbon dioxide; CO2 capture; low-cost adsorbent; surface adsorption

2
1. Introduction

Carbon dioxide (CO2) is the primary greenhouse gas (GHG) produced by human

activities, which are the main cause for climate change [1]. In order to lower GHG emissions

enough to stabilize global temperature, it is necessary to nearly eliminate anthropogenic CO2. It

is estimated that between 2010 and 2060, fossil fuel combustion would emit 282-701 gigatonnes

of CO2, assuming the existing global infrastructure does not change [2]. Nevertheless,

atmospheric CO2 is reversible; if emissions are reduced enough, global levels will decrease

because oceans continually take up anthropogenic CO2 [2]. A great deal of research attention

thus has been given to designing materials to capture CO2, particularly with respect to reducing

power and industry CO2 emissions commonly produced by large, stationary exhaust stacks [3, 4].

Post-combustion capture methods, which separate CO2 from flue gas after energy generation, are

considered the most promising technology to satisfy the need for large-scale CO2 reduction [5-7].

The main challenges of this approach are in the separation of gas, high flow rate, and low partial

pressure, which often increase the cost of carbon capture, making it economically unfavorable [1,

8].

The technique of “wet-scrubbing” has been employed for the separation of post-

combustion flue gas for over 50 years. This process uses amines or other alkaline solvents, to

chemically bind CO2 (i.e., chemical absorption) [6, 9]. Although 80-95% of CO2 can be

recovered with this practice, there are both (1) large heat and energy requirements for solvent

regeneration and chemical stripping and (2) potential adverse environmental impacts due to

solvent degradation and reaction emissions, in some processes [2]. Compared to chemical

absorption, solid adsorbents are more energy efficient and mainly utilize physical adsorption

processes to capture CO2 [3, 10]. Many porous sorbents with high surface area, such as activated

3
carbon (AC), zeolite, and metal-organic frameworks (MOFs), are good capture materials with

relatively high CO2 uptake ability [7, 11]. Most of these sorbents are expensive for large-scale

applications, so need to be regenerated. Because of the strong interaction between the CO2

molecules with the “active” sites of the sorbents, this regeneration process may be ineffective

and/or costly [12, 13].

Although AC can be derived from various renewable and inexpensive biomass feedstock,

its production requires either physical or chemical activation processes involving high pyrolysis

temperatures (up to 950 oC) and/or chemical additions (e.g., ZnCl2) that may impose both

economic and environmental burdens [14, 15]. Additional modifications, such as surface

functionalization with amines and metal oxides, are applied to further engineer the AC to

improve its performance to remove CO2 from flue gas [16, 17]. These modifications, however,

may require additional capital investment as well as environmental burdens to produce the

carbon-based CO2 adsorbent.

Similar to AC, another porous carbon framework called biochar has received attention for

its ability to remove contaminants, sequester carbon in soils, and concurrently improve soil

quality [18, 19]. Biochar is more environmentally friendly and at least 10x cheaper than AC and

other common CO2 capture materials because it is often produced from various waste biomass

residues through one-step slow pyrolysis at relatively low temperature without sophisticated

equipment [19, 20]. The pyrolysis process, which involves heating under oxygen-starving

conditions, produces this pyrogenic form of carbon that is less susceptible to

degradation/mineralization [19, 21]. In regards to supply, biochar can be generated from waste

biomass and agricultural/forest residue, so the source is abundant. Although its ability to adsorb

gaseous pollutants is still unclear, biochar has unique surface properties that give it excellent

4
potential for CO2 capture because not only does it tend to be polar and hydrophilic, but it also

has an extensive porous structure with comparatively high surface area-to-weight ratios [22-24].

If it can be used as an alternative, low-cost adsorbent for CO2 capture with reasonable efficiency,

no regeneration may be needed because spent biochar can potentially be applied directly to the

soils to improve soil quality.

The overarching objective of this work was to determine biochar’s potential as a low-cost

CO2 capture material. Six types of biochars were prepared from two commonly used feedstock

materials through slow pyrolysis at 300, 450, 600 oC and were assessed for their sorption of CO2.

Specific objectives were to (1) measure and model the adsorption kinetics of CO2 onto the

biochars; (2) determine the effect of temperature on the adsorption process; and (3) elucidate the

governing adsorption mechanisms of CO2 onto the biochars.

2. Experimental

2.1. Biochar production

Two commonly used biomass materials, sugarcane bagasse (BG) and hickory wood

(HW), were used as feedstock for biochar production. The raw materials were oven dried and

converted into biochar through slow pyrolysis in a furnace (Olympic 1823HE) in a N2

environment at temperatures of 300, 450, and 600 oC. The six biochar samples are labeled as

BG300, BG450, BG600, HW300, HW450, and HW600, based on the feedstock type and

pyrolysis temperature. All samples were then crushed and sieved yielding a 0.5-1 mm size

fraction. After rinsing with DI water several times for cleaning, the resulting biochar samples

were oven-dried and sealed in containers. Detailed information about biochar production

procedures was reported in Yao et al. [25, 26].

5
2.2. Characterization

The basic physicochemical properties of the biochar samples used in this work have been

determined previously following the standard methods [25, 27]. In brief, elemental composition

was determined with a CHN Elemental Analyzer (Carlo-Erba NA-1500) and inductively-coupled

plasma atomic emission spectroscopic (ICP-AES) analyzer. The oxygen (O) content of each

sample was calculated by weight difference between the sample and other major elements.

Biochar surface areas were measured with a Quantachrome Autosorb-1 analyzer using the N2

adsorption method.

In addition, Fourier Transform Infra-Red (FTIR) analysis was used to detect the surface

functional groups on the samples. The biochar samples were further ground into finer powder

and then directly mounted on the diamond base of a Nicolet 6700 FTIR (Thermo Scientific) for

the FTIR analysis.

2.3. CO2 capture

Adsorption kinetics was determined by using a Thermogravimetric Analysis (TGA)

instrument at a flow rate of 50 ml/min following the procedures of Shafeeyan et al. [28]. CO2

capture has been shown to be more effective under high pressures [29], so this low flow should

provide an underestimate of the total capacity of the biochar for CO2 adsorption. About 22 mg of

biochar sample was placed on the sample holder of the TGA. Nitrogen gas was fed through the

sample chamber at a temperature of 120 oC for one hour to purge the volatile components (e.g.,

moisture, solvents, and residual monomers) out of the sample. The volumetric flow rate during

the purification step was the same as the flow rate during the adsorption steps. Next, the

6
temperature was cooled to the experimental temperature of 75 or 25 oC and CO2 was fed into the

analyzer for over one hour. Mass measurements were initialized once the experimental

temperature is reached (N2 adsorption on the sample was considered negligible). A blank run

was performed at all temperature variations under identical test conditions. This baseline

considers instrument and buoyancy influences to ensure correct mass change values.

2.4. Effect of adsorption temperature

To determine how the adsorption of CO2 varies with temperature, the total change in

mass of BG600 due to CO2 adsorption was measured using the adsorption procedure discussed in

2.3. A 22 mg sample was first degassed with nitrogen at 120 oC and CO2 was fed through at the

terminal temperatures of 25, 30, 40, 75, 90, and 100 oC. To assess how CO2 was desorbed

depending on a gradual temperature increase, after a 22 mg sample of BG600 was degassed, the

temperature was then dropped to 25 oC and CO2 was fed through the analyzer. The temperature

was then elevated from 25 oC (80 minutes) to 75 oC (40 minutes) and then to 120 oC (60 minutes)

and the change in mass due to CO2 adsorption was evaluated.

2.5. Statistical Analysis

Statistical analysis was performed to determine the correlations between the adsorption of

CO2 and the surface properties of the biochar using both the Spearman rank and Pearson tests.

3. Results and Discussion

3.1. Biochar properties

7
All the biochar samples used in this work were rich in carbon (> 69%), oxygen (> 11%),

and hydrogen (> 2%), but contained relatively low amount of nitrogen (0.17%-0.92%), which is

common for chars converted from agricultural and forest residuals [30]. Part of the oxygen,

hydrogen, and nitrogen elements in the biochar samples are likely in functional groups (e.g.,

carboxyl, hydroxyl, and amine) on the carbon surfaces [31, 32]. In comparison with the samples

produced at other temperatures, the two 450 oC biochar samples had higher carbon but lower

oxygen content. As a result, they also had the lowest O/C and the O/H values (Table 1). The

three BG biochars had a higher level of nitrogen than their corresponding HW biochars (Table

1), suggesting that BG biochars may contain more nitrogenous functional groups, such as amine

groups, on their surfaces. Because CO2 has a strong affinity towards amine functional groups [3,

16], BG biochar is potentially a better CO2 capture material than the HW biochar.

Results of surface area analysis show that when the biomass underwent higher

temperature pyrolysis, the surface area increased (Table 1), which is consistent with the literature

that high pyrolysis temperature would create more micropores [30]. The surface area of HW600

(401 m2/g) and BG600 (388 m2/g) was much higher for the biochars prepared at lower

temperatures (Table 1). Previous studies showed that adsorbents with high surface area may have

large CO2 capture capacity [1, 3]. Thus, it is anticipated that HW600 and BG600 would have

better CO2 capture ability than the other biochar samples.

The FTIR analysis confirms the abundance of oxygen, hydrogen, and nitrogen containing

functional groups on biochar surfaces (Figure 1). Alkane (wave number, ~2910), and cyclic

alkene (~1570) groups were identified on BG300 and BG450 samples, along with a strong peak

that could represent sulfur, amine, or ester groups (~1195). This strong peak (~1195) and cyclic

alkene groups (~1570) were similarly found on HW300 and HW450 samples. Compared with

8
the biochars converted at lower temperature, the HW600 and BG600 lost several surface

functional groups, particularly the strong, unclear, peak (~1195), due to the high temperature

pyrolysis. Previous studies have suggested that high temperature pyrolysis can introduce

dehydration and decarboxylation reactions to reduce the hydrogen and oxygen containing

functional groups on carbon surfaces [33-35]. Nevertheless, there could be an increase in

nitrogenous groups (2000-2400) in the two 600 oC biochar samples (Figure 1).

3.2. Adsorption of CO2

The adsorption of CO2 onto the six biochar samples at two different temperatures (i.e., 25

and 75 oC) show similar kinetic behaviors (Figure 2). Adsorption was very fast at the beginning,

slowed down after about 10 minutes, and finally reached a plateau after 60 minutes, indicating it

was approaching equilibrium. Comparison of the adsorption kinetic curves show that the

adsorption of CO2 onto the biochars increased with the pyrolysis temperature, with HW600 and

BG600 having the highest CO2 adsorption rates at the equilibrium. For all the biochars tested,

adsorption decreased when the reaction temperature increased from 25 to 75 oC, suggesting this

was an exothermic process. The findings were consistent with previous studies of CO2

adsorption on AC and other carbon-based adsorbents [7, 36]. Although the HW600 had larger

surface area than the BG600 (Table 1), its CO2 adsorption curve of was slightly lower than that

of the BG600 for both 25 and 75 oC adsorption conditions (Figure 2). This result suggests that, in

addition to surface area, other factors, such as interactions with surface functional groups, could

also play important roles in controlling the adsorption of CO2 on the biochars. As shown in the

FTIR analysis, there were nitrogenous groups on the surface of the biochars. Those weakly basic

and polar functional groups would induce a multipole on CO2, an acidic gas. Previous studies,

9
such as Zhang et al [37], revealed that the presence of nitrogenous groups tends to increase the

CO2 adsorption capacity in activated carbon.

The adsorption kinetics could be best described using the pseudo-second order model:

dC s
= k 2 (C e − C s ) 2 (1)
dt

where Cs (mg·g-1) and Ce (mg·g-1) are the CO2 adsorption concentrations at time t and at

equilibrium, respectively, and k2 (g·mg-1·s-1) is the second-order adsorption rate constants [38].

The model simulations matched the experimental data very well (Figure 2) and the coefficient of

determination (R2) values were larger than 0.988 for all the conditions tested (Table 2). This

result suggested that the adsorption kinetics of CO2 on the biochars could be controlled by more

than one rate-determining step or mechanism [39].

The second-order adsorption rate constant (k2) tends to decrease with increasing pyrolysis

temperature in the Bagasse samples for both 25 and 75 oC adsorption temperatures (Table 2).

The initial kinetic rate constant (k0), which is the product of k2 and Ce, also decreased with

increasing pyrolysis temperature in the Bagasse samples.

The equilibrium adsorption concentration (Ce) represents the maximum capacity for the

biochar to adsorb CO2 under the tested temperature and pressure conditions (2). The biochars

showed relatively high adsorption to CO2 for both 25 oC (34.48-73.55 mg g-1) and 75 oC (11.15-

43.67 mg g-1). These values are comparable to many of the previously studied CO2 capture

materials, particularly carbon-based adsorbents. For example, Shafeeyan et al. [28] showed that

commercial palm shell-based AC has a CO2 capture capacity of about 48 mg g-1 under similar

experimental conditions. Rashidi et al. [40] found that AC prepared from Coconut fiber has a

CO2 capture capacity of about 20 and 80 mg g-1 at 100 and 25 oC, respectively. Correlation

analyses revealed that there was no detectable correlation between the Ce and the elemental

10
composition of the biochars (Table 3). Both the Spearman (ρ=0.94, p=0.02) and the Pearson

(r=0.84, p=0.02) tests, however, indicate that there is a strong positive correlation, which is

statistically significant, between Ce and biochar surface area, confirming that surface area may

control the adsorption of CO2 [1, 3].

3.3. Mechanisms

The experimental and modeling results suggest that the adsorption of CO2 onto biochar

was mainly controlled by physisorption, which is a weak interaction arising from intermolecular

forces (e.g. Van der Waals forces). Although physisorption of gas molecules on solid surface can

be temperature dependent as it approaches equilibrium, it does not require activation energy [41].

CO2 has a strong quadrupole moment, but no dipole; it can interact with the biochar surface

through polar bonds on either end of its linear shape. Both dispersion and induction contribute to

the attraction of CO2 to carbon surface, depending on the surface property [42]. Recently,

D’Alessandro et al. [1] stressed the importance of high internal surface area when choosing

adsorption materials for CO2 capture. Findings from this work showed that the CO2 capture

ability of most tested biochar samples also increased with the surface area. The only discrepancy

is that HW600 adsorbed less CO2 than BG600, although the Hickory biochar had slightly larger

surface area. This probably can be attributed to the fact that there were more nitrogenous

functional groups on the bagasse biochar, which could promote the physisorption of CO2 on

biochar surfaces because of the strong interaction between the acidic CO2 and basic nitrogenous

surface functional groups [37]. As shown in this work, at any of the three respective pyrolysis

temperatures, bagasse samples had a greater proopoertion of nitrogen than hickory samples and

thus showed higher CO2 adsorption, respectively.

11
To provide further evidence to the mechanism of physical adsorption, the heat of

adsorption of CO2 onto the biochar was determined integrating the adsorption signal in J/s from

the TGA over a pre-defined time at 25 oC as described in Pires et al. [43]. Physical adsorption

typically produces a heat of adsorption between 5 and 40 kJ mol-1 ; whereas, chemical adsorption

is typically between 40 and 800 kJ mol-1 [44]. The heat of adsorption values obtained in this

study ranged from 7.47 to 32.19 kJ mol-1 (Table 2), confirming physisorption controlled the

adsorption of CO2 onto the biochars. Because of the physical adsorption mechanism, surface area

was a significant determinant of CO2 adsorption in this study; nevertheless, the presence of

nitrogenous groups also played a role, when the surface area was sufficient.

3.5. Temperature Dependence

When CO2 was adsorbed to equilibrium at 25 oC, the BG600 could be regenerated by

92% by raising the adsorption temperature to 120 oC (Figure 3). This relatively low temperature

desorption was a further evidence for the mechanism of physisorption. Furthermore, when the

adsorption experiment was performed at 25, 30, 40, 75, 90, 100, and 110 oC, the CO2 adsorption

ability of the BG600 decreased linearly (R2=.93) with the temperatures (Figure 4). These results

are consistent with the findings from previous studies [3, 37], which identifies a strong negative

correlation between adsorption temperature and CO2 adsorption capacity. Nevertheless, the

BG600 showed strong sorption of CO2 at normal atmospheric/soil temperatures (> 50 mg g-1).

This introduces the question of whether spent biochar could be applied directly to soils without

substantial CO2 release.

4. Conclusions

12
Biochar is environmentally friendly, as it is used as amendment to improve soil quality

and to sequester carbon; is low-cost, as production does not require a large energy input; and is

sustainable, as the feedstock can be made from renewable and waste biomass materials.

Experimental data show that biochar could also be a promising framework for CO2 capture with

adsorption capacity comparable to many commonly used, expensive adsorbents. Because the

adsorption of CO2 was mainly through physisorption, biochars with larger surface area generally

showed better adsorption ability, although the presence of nitrogenous groups also promoted the

adsorption. The results also suggest that, under normal soil/atmospheric temperatures, biochar

could retain the bulk of captured CO2 and thus spent biochar could be applied directly as a soil

amendment and carbon storage. Because of all these advantages, biochar is a promising CO2

capture material that can be used to reduce anthropogenic CO2 emission and to mitigate global

warming.

Acknowledgments

This research was partially supported by the NSF through Grant CBET-1054405.

13
References

[1] D.M. D'Alessandro, B. Smit, J.R. Long, Carbon dioxide capture: Prospects for new materials,

Angew. Chem. Int. Edit. 49 (2010) 6058-6082.

[2] S. Davis, K. Caldeira, H. Matthews, Future CO2 emissions and climate change from existing

energy infrastructure, Science 329 (2010) 1330-1333.

[3] C.H. Yu, C.H. Huang, C.S. Tan, A review of CO2 capture by absorption and adsorption,

Aerosol and Air Quality Research 12 (2012) 745-769.

[4] A. Kaithwas, M. Prasad, A. Kulshreshtha, S. Verma, Industrial wastes derived solid

adsorbents for CO2 capture: A mini review, Chem. Eng. Res. Des. 90 (2012) 1632-1641.

[5] M. Olivares-Marin, M. Maroto-Valer, Development of adsorbents for CO2 capture from

waste materials: A review, Greenhouse Gases-Science and Technology 2 (2012) 20-35.

[6] A.A. Olajire, CO2 capture and separation technologies for end-of-pipe applications - A

review, Energy 35 (2010) 2610-2628.

[7] A. Samanta, A. Zhao, G.K.H. Shimizu, P. Sarkar, R. Gupta, Post-combustion CO2 capture

using solid sorbents: A review, Ind. Eng. Chem. Res. 51 (2012) 1438-1463.

[8] M. Zhao, A.I. Minett, A.T. Harris, A review of techno-economic models for the retrofitting

of conventional pulverised-coal power plants for post-combustion capture (PCC) of CO2, Energy

& Environmental Science 6 (2013) 25-40.

[9] M. Wang, A. Lawal, P. Stephenson, J. Sidders, C. Ramshaw, Post-combustion CO2 capture

with chemical absorption: A state-of-the-art review, Chem. Eng. Res. Des. 89 (2011) 1609-1624.

[10] N.P. Wickramaratne, M. Jaroniec, Activated carbon spheres for CO2 adsorption, ACS Appl.

Mater. Interfaces 5 (2013) 1849-1855.

14
[11] E.R. Parnham, R.E. Morris, Ionothermal synthesis of zeolites, metal-organic frameworks,

and inorganic-organic hybrids, Accounts. Chem. Res. 40 (2007) 1005-1013.

[12] E.S. Kikkinides, R.T. Yang, S.H. Cho, Concentration and recovery of CO2 from flue-gas by

pressure swing adsorption, Ind. Eng. Chem. Res. 32 (1993) 2714-2720.

[13] G.D. Pirngruber, F. Guillou, A. Gomez, M. Clausse, A theoretical analysis of the energy

consumption of post-combustion CO2 capture processes by temperature swing adsorption using

solid sorbents, International Journal of Greenhouse Gas Control 14 (2013) 74-83.

[14] W. Heschel, E. Klose, On the suitability of agricultural by-products for the manufacture of

granular activated carbon, Fuel 74 (1995) 1786-1791.

[15] M.T.O. Jonker, M.P.W. Suijkerbuijk, H. Schmitt, T.L. Sinnige, Ecotoxicological effects of

activated carbon addition to sediments, Environ. Sci. Technol. 43 (2009) 5959-5966.

[16] A. Houshmand, W.M.A.W. Daud, M.S. Shafeeyan, Exploring potential methods for

anchoring amine groups on the surface of activated carbon for CO2 adsorption, Separ. Sci.

Technol. 46 (2011) 1098-1112.

[17] G.L. Drisko, C. Aquino, P.H.M. Feron, R.A. Caruso, S. Harrisson, V. Luca, One-pot

preparation and CO2 adsorption modeling of porous carbon, metal oxide, and hybrid beads, ACS

Appl. Mater. Interfaces 5 (2013) 5009-5014.

[18] S. Jeffery, F.G.A. Verheijen, M. van der Velde, A.C. Bastos, A quantitative review of the

effects of biochar application to soils on crop productivity using meta-analysis, Agr. Ecosyst.

Environ. 144 (2011) 175-187.

[19] J. Lehmann, J. Gaunt, M. Rondon, Bio-char sequestration in terrestrial ecosystems – A

review, Mitigation Adapt. Strat. Global. Change (2006) 403–427.

15
[20] P. Oleszczuk, S.E. Hale, J. Lehmann, G. Cornelissen, Activated carbon and biochar

amendments decrease pore-water concentrations of polycyclic aromatic hydrocarbons (PAHs) in

sewage sludge, Bioresource Technol. 111 (2012) 84-91.

[21] A.R. Zimmerman, B. Gao, M.Y. Ahn, Positive and negative carbon mineralization priming

effects among a variety of biochar-amended soils, Soil Biol. Biochem. 43 (2011) 1169-1179.

[22] D. Mohana, A. Sarswata, Y.S. Okb, C.U.P. Jr., Organic and inorganic contaminants removal

from water with biochar, a renewable, low cost and sustainable adsorbent - A critical review,

Bioresource Technol. (2014) doi: 10.1016/j.biortech.2014.1001.1120.

[23] M. Ahmad, A.U. Rajapaksha, J.E. Lim, M. Zhang, N. Bolan, D. Mohan, M. Vithanage, S.S.

Lee, Y.S. Ok, Biochar as a sorbent for contaminant management in soil and water: A review,

Chemosphere 99 (2014) 19-33.

[24] L. Beesley, E. Moreno-Jimenez, J.L. Gomez-Eyles, E. Harris, B. Robinson, T. Sizmur, A

review of biochars' potential role in the remediation, revegetation and restoration of

contaminated soils, Environ. Pollut. 159 (2011) 3269-3282.

[25] Y. Yao, B. Gao, M. Zhang, M. Inyang, A.R. Zimmerman, Effect of biochar amendment on

sorption and leaching of nitrate, ammonium, and phosphate in a sandy soil, Chemosphere 89

(2012) 1467-1471.

[26] Y. Yao, B. Gao, M. Inyang, A.R. Zimmerman, X. Cao, P. Pullammanappallil, L. Yang,

Biochar derived from anaerobically digested sugar beet tailings: Characterization and phosphate

removal potential, Bioresource Technol. 102 (2011) 6273-6278.

[27] Y. Sun, B. Gao, Y. Yao, J. Fang, M. Zhang, Y. Zhou, H. Chen, L. Yang, Effects of

feedstock type, production method, and pyrolysis temperature on biochar and hydrochar

properties, Chem. Eng. J. 240 (2014) 574-578.

16
[28] M.S. Shafeeyan, W.M.A.W. Daud, A. Houshmand, A. Arami-Niya, Ammonia modification

of activated carbon to enhance carbon dioxide adsorption: Effect of pre-oxidation, Appl. Surf.

Sci. 257 (2011) 3936-3942.

[29] Y.C. Guo, J.F. Zhao, J.X. Xu, W. Wang, F.S. Tian, G.Y. Yang, M.P. Song, Aerobic

oxidation of benzyl alcohol catalyzed by Cu-Mn mixed oxides and 2,2,6,6-tetramethyl-piperidyl-

1-oxyl, Journal of Natural Gas Chemistry 16 (2007) 210-212.

[30] J. Lehmann, S. Joseph, Biochar for environmental management: Science and technology,

Earthscan/James & James, 2009.

[31] M. Uchimiya, S. Chang, K.T. Klasson, Screening biochars for heavy metal retention in soil:

Role of oxygen functional groups, J. Hazard. Mater. 190 (2011) 432-441.

[32] M.D. Inyang, B. Gao, W.C. Ding, P. Pullammanappallil, A.R. Zimmerman, X.D. Cao,

Enhanced lead sorption by biochar derived from anaerobically digested sugarcane bagasse,

Separ. Sci. Technol. 46 (2011) 1950-1956.

[33] Y. Guo, D.A. Rockstraw, Physicochemical properties of carbons prepared from pecan shell

by phosphoric acid activation, Bioresource Technol. 98 (2007) 1513-1521.

[34] A. Mukherjee, A.R. Zimmerman, W. Harris, Surface chemistry variations among a series of

laboratory-produced biochars, Geoderma 163 (2011) 247-255.

[35] G.N. Kasozi, A.R. Zimmerman, P. Nkedi-Kizza, B. Gao, Catechol and humic acid sorption

onto a range of laboratory-produced black carbons (biochars), Environ. Sci. Technol. 44 (2010)

6189-6195.

[36] Z. Zhang, J. Zhou, W. Xing, Q. Xue, Z. Yan, S. Zhuo, S.Z. Qiao, Critical role of small

micropores in high CO2 uptake, Phys. Chem. Chem. Phys. 15 (2013) 2523-2529.

17
[37] C.M. Zhang, W. Song, G.H. Sun, L.J. Xie, J.L. Wang, K.X. Li, C.G. Sun, H. Liu, C.E.

Snape, T. Drage, CO2 capture with activated carbon grafted by nitrogenous functional groups,

Energ. Fuel. 27 (2013) 4818-4823.

[38] Y. Yao, B. Gao, M. Inyang, A.R. Zimmerman, X.D. Cao, P. Pullammanappallil, L.Y. Yang,

Removal of phosphate from aqueous solution by biochar derived from anaerobically digested

sugar beet tailings, J. Hazard. Mater. 190 (2011) 501-507.

[39] Y.S. Ho, G. McKay, Pseudo-second order model for sorption processes, Process. Biochem.

34 (1999) 451-465.

[40] N.A. Rashidi, S. Yusup, B.H. Hameed, Kinetic studies on carbon dioxide capture using

lignocellulosic based activated carbon, Energy 61 (2013) 440-446.

[41] Anonymous, Manual of symbols and terminology for physicochemical quantities and units -

Appendix-3 - Electrochemical nomenclature, Denki Kagaku 48 (1980) 46-55.

[42] R.T. Yang, Adsorbents: Fundamentals and applications, Wiley-Interscience, 2003.

[43] J. Pires, M.B. de Carvalho, A.P. Carvalho, J.M. Guil, J.A. Perdigon-Melon, Heats of

adsorption of N-hexane by thermal gravimetry with differential scanning calorimetry (TG-DSC):

A tool for textural characterization of pillared clays, Clay. Clay. Miner. 48 (2000) 385-391.

[44] Y.C. Chang, D.H. Chen, Recovery of gold(III) ions by a chitosan-coated magnetic nano-

adsorbent, Gold. Bull. 39 (2006) 98-102.

18
Figure captions

Figure 1. FTIR Spectra of the biochar samples prepared at different temperature: (a) Sugarcane

bagasses, and (b) Hickory wood.

Figure 2. Adsorption kinetics of CO2 onto (a) BG biochars at 25 oC, (b) BG biochars at 75 oC,

(c) HW biochars at 25 oC, and (d) HW biochar at 75 oC.

Figure 3. Desorption kinetics of CO2 on BG600 as temperature is raised incrementally.

Figure 4. Effect of temperature on CO2 adsorption onto BG600

19
Table 1. Basic physical properties of biochar samples.

C% O% H% N% O/C H/C O/H Surface Area


HW300 69.13 24.36 4.85 0.39 0.35 0.07 5.00 0.10
HW450 83.62 11.46 3.24 0.17 0.14 0.04 3.50 12.90
HW600 81.81 14.03 2.17 0.73 0.17 0.03 5.67 401.00
BG300 69.50 24.36 4.20 0.90 0.35 0.06 5.83 5.20
BG450 78.60 15.45 3.52 0.92 0.20 0.04 5.00 13.60
BG600 76.45 18.32 2.93 0.79 0.24 0.04 6.00 388.30

Table 2. Best-fit parameter values from model simulations of CO2 adsorption kinetics

Heat of
T (oC) k2 (g· mg-1·s-1) Ce (mg g-1) k2Ce (s-1) adsorption R2
(kJ mol-1)
HW300 25 0.0000356 34.48 0.001227 9.14 0.999
HW450 25 0.00002517 44.96 0.001132 32.19 0.995
HW600 25 0.00002614 61.00 0.001595 18.06 0.997
HW300 75 0.0001283 11.15 0.001431 9.15 0.997
HW450 75 0.00007619 23.40 0.001783 32.19 0.997
HW600 75 0.00008087 40.63 0.003286 18.06 0.999
BG300 25 0.00008302 38.72 0.003215 13.60 0.996
BG450 25 0.00004385 53.83 0.002360 7.47 0.998
BG600 25 0.00001531 73.55 0.001126 26.62 1.000
BG300 75 0.0002462 19.82 0.004880 13.60 0.988
BG450 75 0.00009848 30.35 0.002989 7.47 0.997
BG600 75 0.00003603 43.67 0.001573 26.62 0.999

20
Table 3. Pearson and Spearman tests for the correlation between Ce and basic biochar properties.

Pearson Test Spearman Test


Results
coefficient (r) p-value coefficient (ρ) p-value
C% 0.47 0.17 0.49 0.36 Not Detected
H% -0.80 0.03 -0.89 0.03 Not Detected
N% 0.51 0.15 0.26 0.66 Not Detected
O% -0.37 0.23 -0.62 0.30 Not Detected
O/C -0.49 0.16 -0.52 0.42 Not Detected
H/C -0.77 0.04 -0.70 0.24 Not Detected
O/H 0.48 0.17 0.54 0.30 Not Detected
Surface Area 0.84 0.02 0.94 0.02 Correlated

21
Figure 1

a
BG300
-COOCO-
BG450
-NH2
BG600
Intensity

-NH2
-COOH

1000 1500 2000 2500 3000 3500


Wavelength (cm-1)

b HW300
-COOCO-
HW450

-NH2 HW600
-NH2
Intensity

1000 1500 2000 2500 3000 3500


Wavelength (cm-1)

22
Figure 2
70 70
BG300 BG300
BG450
a BG450
b
60 60
BG600 BG600
Amout of CO2 Adsorbed ( mg g-1)

Amout of CO2 Adsorbed ( mg g-1)


Model Model
50 50

40 40

30 30

20 20

10 10

0 0
0 20 40 60 80 0 20 40 60 80
Time (minutes) Time (minutes)

70 70
HC300
HC300 c d
HC450 HC450
60 60
HC600 HC600
Amout of CO2 Adsorbed ( mg g-1)
Amout of CO2 Adsorbed ( mg g-1)

Model Model
50 50

40 40

30 30

20 20

10 10

0 0
0 20 40 60 80 0 20 40 60 80
Time (minutes) Time (minutes)

23
Figure 3
90 140
CO2 Adsorbed
80 Temperature
120

70
100
Amount of CO2 Adsorbed (mg g-1)

60

Temperature (oC)
50 80

40 60

30
40
20

20
10

0 0
0 50 100 150
Time (minutes)

24
Figure 4
80

70

60

50
Ce (mg g-1)

y = -0.5065x + 76.705
R² = 0.9329
40

30

20

10

0
0 20 40 60 80 100 120 140

Adsorption Temperature (oC)

25
 Biochar effectively captured CO2

 CO2 weakly bound to the surface of biochar through physisorption

 Surface area was a significant determinant of CO2 adsorption onto biochar

26

You might also like