You are on page 1of 12

Science of the Total Environment 826 (2022) 154133

Contents lists available at ScienceDirect

Science of the Total Environment


journal homepage: www.elsevier.com/locate/scitotenv

Insights into CO2 adsorption on KOH-activated biochars derived from the


mixed sewage sludge and pine sawdust

Kai Li a, Dongqing Zhang b, , Xiaojun Niu a,b,d,e,
⁎⁎, Huafang Guo c,⁎⁎⁎, Yuanyuan Yu b, Zhihua Tang c,
Zhang Lin a, Mingli Fu a
a
School of Environment and Energy, South China University of Technology, Guangzhou 510006, PR China
b
Guangdong Provincial Key Laboratory of Petrochemical Pollution Processes and Control, School of Environmental Science and Engineering, Guangdong University of
Petrochemical Technology, Maoming 525000, PR China
c
The Key Laboratory of Renewable Energy, Guangzhou Institute of Energy Conversion, Chinese Academy of Sciences, Guangzhou 510640, PR China
d
Guangdong Provincial Key Laboratory of Atmospheric Environment and Pollution Control, South China University of Technology, Guangzhou Higher Education Mega Centre,
Guangzhou 510006, PR China
e
The Key Lab of Pollution Control and Ecosystem Restoration in Industry Clusters, Ministry of Education, South China University of Technology, Guangzhou Higher Education Mega
Centre, Guangzhou 510006, PR China

H I G H L I G H T S G R A P H I C A L A B S T R A C T

• KOH-modified biochars displayed tunable


porous features including surface area and
microporosity.
• KOH-modified biochars exhibit superior
CO2 adsorption capacity.
• The sorption mechanisms of CO2 were
verified by in situ DRIFT spectroscopy.
• It behaves perfect cyclic regeneration per-
formance as high as 97.1% retention.

A R T I C L E I N F O A B S T R A C T

Article history: The environment issues associated with global warming and climate change caused by continuous increase in green-
Received 15 December 2021 house gas emissions have attracted worldwide concerns. As renewable resources with good adsorption property, bio-
Received in revised form 6 February 2022 char is an efficient and environmental friendly adsorbsent for CO2 capture. In this study, the CO2 adsorption behavior
Accepted 21 February 2022
of biochars derived from feedstock mixtures of 70% pine sawdust and 30% sewage sludge by KOH modification was
Available online 25 February 2022
investigated. The textual properties and functional groups of the pristine biochars have been significantly enhanced
Editor: Daniel CW Tsang after KOH activation. With highly developed microporosity, the specific surface area (SSA) of the KOH-modified bio-
chars increased by 3.9–14.5 times. Furthermore, higher CO2 adsorption capacities of 136.7–182.0 mg/g were ob-
Keywords: served for the modified biochars, compared to pristine ones (35.5–42.9 mg/g). The development of micropores by
Biochars composite KOH activation significantly increased the CO2 adsorption capacity. Meanwhile, the presence of hetero atoms (O
High surface area and K) also positively influenced CO2 adsorption capacity of biochar. Noticeably, both physical and chemical adsorp-
Tunable porosity tion played a crucial role in CO2 capture, which was verified by different characterization methods including high res-
Chemical sorption olution scanning electron microscope, X-ray photoelectron spectroscopy and in situ diffuse reflectance infrared Fourier
CO2 capture
transform (DRIFT) spectroscopy. The Findings of this study demonstrate the -significance of chemical sorption by iden-
tifying the transformation of CO2 by biochar composites and in situ characterization of weakly adsorbed and newly

⁎ Correspondence to: D. Zhang, College of Environmental Science and Engineering, Guangdong University of Petrochemical Technology, Maoming 525000, PR China.
⁎⁎ Correspondence to: X. Niu, School of Environment and Energy, South China University of Technology, Guangzhou 510006, PR China.
⁎⁎⁎ Correspondence to: H. Guo, The Key Laboratory of Renewable Energy, Guangzhou Institute of Energy Conversion, Chinese Academy of Sciences, Guangzhou 510640, PR China.
E-mail addresses: dqzhang3377@outlook.com (D. Zhang), scutenv@outlook.com (X. Niu), guohf@ms.giec.ac.cn (H. Guo).

http://dx.doi.org/10.1016/j.scitotenv.2022.154133
0048-9697/© 2022 Elsevier B.V. All rights reserved.
K. Li et al. Science of the Total Environment 826 (2022) 154133

formed mineral species during the CO2 sorption process. Moreover, BC700K showed 97% recyclability during 10 con-
secutive adsorption-desorption cycles at 25 °C, 1 bar. The results obtained in the present study may inspire new re-
search interest and provide a comprehensive insight into the research subject to biochars derived from feedstock
mixtures for CO2 capture.

1. Introduction nitrogen-rich agricultural waste, and the micropore structure played an


important role in CO2 capture performance at the lower adsorption
Carbon dioxide (CO2), one of the major anthropogenic greenhouse gas temperature.
(GHG) derived mainly from fossil fuel combustion, has been intensively dis- More recently, the modification of biochar using different physic-
cussed worldwide, due to its role as the principal sponsor toward global chemical processes as well as synthesis of biochar composites to enhance
warming (Peter, 2018; Wei et al., 2018). From 1970 to 2018, worldwide the adsorptive capacities of CO2 has been extensively explored (Ahmed
CO2 emission has grown from roughly 16 to 37.5 Gigatonnes (Gt), and et al., 2019; Igalavithana et al., 2020). However, studies have shown that
will reach 60 Gt by 2050 (Jahandar et al., 2019; Kamkeng et al., 2021). Ex- most activated modified carbon materials do not have abundant acidic
cessive emission of CO2 into atmosphere may cause tremendous adverse groups. Therefore, the preparation of carbon materials with high specific
impacts on global ecosystem, such as glacial melting, heavy droughts, ex- surface area and rich oxygen/nitrogen functional groups is very important
treme heat waves, and sea level rising (Wang et al., 2010). Therefore, it is for the adsorption and retention of CO2. Biochar activation through impreg-
imperative to explore efficient and sustainable strategies for mitigating nating with proper chemicals, such as KOH, is one of the most effective
the rising level of CO2 as a means to address the challenges of global climate modification technologies for promoting the selectivity of biochar and the
change (Esposito et al., 2019). adsorption capacity for CO2, since the alkalinity favors the dissolution of
A series of technologies have been explored for achieving effective acidic CO2 to modify the biochar surface properties (Dissanayake et al.,
carbon emission control, such as solvent absorption, adsorption with solid 2020a), and the formed potassium species after KOH activation diffused
sorbents, membrane separation, and cryogenic separation (Ahmed et al., into the internal structure of the biochar, which increased the width of
2020; Dissanayake et al., 2020a; Yaumi et al., 2017). Among them, CO2 pores and granted biochars for more excellent pore structure (Mao et al.,
adsorbents in solid form have been proven to be the most feasible and 2014). The enhanced CO2 adherence to the surface of modified biochar
economical alternative for remediating GHG emission in a short period of can be achieved by forming a chemical bond, predominantly a covalent
time, owing to its cost-effectiveness, low energy consumption, ease regener- bond (Wei et al., 2017; Xu et al., 2016). For example, Wei et al. (2017) pro-
ation, and sustainability merits (Lahijani et al., 2018). As such, various posed a low-cost method for facile synthesis of nitrogen-doped porous
adsorbents, such as porous carbon materials, molecular sieves, metal- carbon by modification of longan shells as carbon source and carbamide
organic frame-works (MOFs), covalent organic frameworks (COFs), and as additional nitrogen source through KOH activation, and reported that
hydrogen-bonded organic frameworks (HOFs), have been developed for the modified biochar with pyrolysis at 800 °C exhibited great CO2 uptake
capturing CO2 over last few years (Igalavithana et al., 2020; Jalilov et al., of 246 mg/g with KOH:biomass in 1:1 ratio. Similar result was reported
2017; Lin et al., 2019; Zeng et al., 2016). However, their wide application in a study conducted by Ahmed et al. (2019), in which slash pine impreg-
on a large scale using the existing adsorbents is usually associated with nated with ZnCl2 of 1:1 obtained high CO2 capture ability of 137.3 mg/g.
certain limitations, such as high cost, difficulties in regeneration, and Moreover, Yuan et al. (2019) proposed the one-step synthesis using bio-
adsorption competition, making them fail to match the targeted CO2 mass tar as feedstock, porous CaO as a template, and KOH as an activation
capture capacity (Shao et al., 2018). agent. The authors revealed that the highest CO2 adsorption capacity of
Biochar, a class of novel porous carbonaceous materials through the 133.3 mg/g was attained with biomass tar:CaO:KOH in 1:2:3 ratio.
thermochemical conversion of organic materials under oxygen-deficient Activated sludge, which contains various contaminants, such as organic
or oxygen-free conditions, has attracted increasing attention from the scien- matters, heavy metals, micro pollutants, and pathogens, may pose serious
tific community. Owing to its large specific surface area (Huang et al., risk to ecosystem and human health. It has been estimated that in China ap-
2018), high porosity (Dissanayake et al., 2020a), abundant functional proximately 60 million tons of sewage sludge is produced from the waste-
groups (Leng et al., 2020), and low regeneration energy (Dissanayake water treatment plants (WWTPs) each year (Gao et al., 2019). The
et al., 2020b), biochar has been recognized as emerging promising and utilization of sludge-derived biochar as economic feedstocks can therefore
eco-friendly sorbents for various environmental applications, such as soil benefit both carbon abatement and sustainable waste management. Saw-
quality improvement, mitigation of GHS emission and removal of emerging dust, which is a kind of plant-based biomass and contains high content of
contaminants (Igalavithana et al., 2020). Various biochars, derived from volatile matters (70–80%) and low ash content (<40%) (Yahya et al.,
abundant feedstock, such as walnut shell (Lahijani et al., 2018), sugarcane 2015), is conducive to producing biochar with developed pore structure
bagasse and hickory wood (Creamer et al., 2014), sewage sludge (Xu et al., and rich functional groups on the surface (Hashem et al., 2020). Many pre-
2016), cotton stalk (Xiong et al., 2013), and soybean straw (Zhang et al., vious studies have indicated that the co-pyrolysis of mixed sewage sludge
2016), have been demonstrated great potential for CO2 adsorption. Mean- and biomass is a promising method, due to the increase in the quality of py-
while, biochar characteristics are critically influenced by the types of se- rolysis products, cost-effectiveness, low energy consumption moisture re-
lected feedstock (Dutta et al., 2021). For instance, Xu et al. (2016) duction, and more developed porous structures generated (Wang et al.,
evaluated the sorption capacity for CO2 using biochars produced from pig 2016). For example, Fang et al. (2018) observed that the activation energy
manure, sewage sludge, and wheat straw at 500 °C under slow pyrolysis, for the co-pyrolysis of sewage sludge and biomass was significantly lower,
and reported that all three types of biochars exhibited high adsorption ca- indicating that an optimized blending ratio of two materials can effectively
pacities (18.2–34.4 mg/g) for CO2 capture. The aurhors concluded that reduce energy consumptions. Similarly, Alvarez et al. (2015) reported that
CO2 sorption not only resulted from physical sorption related with the sur- during co-pyrolysis, the synergistic effect between sewage sludge and
face area and the total pore volume, Mineral components such as Mg, Ca, different types of biomass promoted the productivity and quality of
Fe, K, etc. in biochar induced the chemical sorption of CO2. Xiong et al. products. Additionally, Wang et al. (2019) prepared biochars via co-
(2013) investigated the performance of CO2 uptake using the biochar ob- pyrolysis of sewage sludge and cotton stalks containing lignin, and
tained from cotton stalk and revealed that the highest adsorption capacity observed that biochars with a high mixing ratio of 9:1 (cotton stalks/
of 58 mg/g was observed at 500 °C, which was predominately caused by sweage sludge, w/w) exhibited high degrees of aromaticity and more
micropore volume. Zhang et al. (2016) reported that the maximum CO2 ad- developed porous structures, which was conducive to CO2 capture. Further-
sorption of 45 mg/g was observed at 500 °C using soybean straw with more, according to Kończak et al. (2020), after adding willow to sewage

2
K. Li et al. Science of the Total Environment 826 (2022) 154133

sludge during pyrolysis, fewer tarry substances (which may cause pore oven at 105 °C overnight. The KOH-activated samples were stored in
blocking) can be produced and reducing mobility and toxicity of metals sealed containers for the subsequent biochar characterization and CO2
in biochar. With this regard, the application of highly-efficient biochar adsorption experiments.
modified by the mixed wood chips and sewage sludge could not only reme-
diate environmental contamination, but also reduce the impacts associated 2.2. Characterization of biochar
with climate change and global warming through limiting the release of
CO2. In addition, 2-step pyrolysis was utilised to produce bochar, which All the biochar samples were placed in the oven and dried overnight at
was also employed in a couple of previous studies due to its unique advan- 80 °C to remove the moisture prior to the usage. The specific surface area
tages. For example, Ahmed et al. (2019) found that pre-pyrolysis was a fac- (SSA) was determined by the Brunauer-Emmett-Teller (BET) method by
ile process that reduces the chemicals required to produce porous carbon measuring the adsorption amount of the sample to high purity liquid nitro-
material during activated process, which still maintained excellent physico- gen. The test environment was liquid nitrogen at low temperature (77 K).
chemical properties for CO2 adsorption capacity. Similarly, mesoporous The pore size distribution (PSD) information was obtained using the den-
nitrogen-doped carbon derived from coconut shell was synthesized through sity functional theory (DFT) method with N2 adsorption isotherm. The
carbonization and subsequent KOH activation for CO2 adsorption by Yaumi total pore volume (Vt) and average pore size (Dap) were calculated at p/
et al. (2018), who found that beside from the process being viable, 2-step p0 ~ 0.99. The micropore volume (Vmic) and specific surface area (Smic)
pyrolysis has the advantage of incorporating nitrogen functionalities into were determined following the t-plot method. The textural properties of
the carbon matrix, which is appealing and favorable for CO2 adsorption. the biochars were investigated by N2 adsorption- desorption measurement
However, to date, most of previous studies have mainly focused on the ad- at −196 °C using an Automatic Micromeritics Analyser 2020 apparatus.
sorption capacity of modified biochar from a single material, and the The changes in surface morphologies of the biochars were determined
knowledge on the potential for using modified biochar derived from a mix- using scanning electron microscopy (SEM, Hitachi S-4800, Japan),
ture of materials, such as activated sludge and woodchips, as a CO2 captur- equipped with the energy dispersive spectrometer (EDS). EDS energy spec-
ing medium, remains largely unknown. trum analysis was performed on the corresponding areas of some samples.
Within this context, the overall objective of this study is to elucidate the The contents of elements on the surface of biochars, such as C, H, O, N, and
CO2 adsorption behavior of biochars derived from feedstock mixtures of S, were detected by CHNS/O Analyzer (Elementar Analysensysteme GmbH,
sewage sludge and pine sawdust, which were modified by KOH. More Germany). The structural features of the biochars were determined using a
specifically, this study is aimed to (i) assess the enhanced physicochemical Raman spectrometer (LabRAM, HR800-LSS5, France). Potassium of bio-
characterization of biochars modified with KOH; (ii) investigate CO2 chars were measured by a flame atomic absorption spectrophotometer
adsorption capacity of different modified biochars; (iii) determine the (FAAS) (M6, Termo Elemental, USA) after microwave digestion. The spec-
potential chemical adsorption mechanisms involved in CO2 adsorption tral range was 1000–2000 cm−1, and the wavelength of light source was
process; and (iv) examine the regenerability and CO2/CH4/N2 selectivity 532 nm. A multifunctional X-ray diffractometer (XPS, ULVAC-PHI
of modified biochars. These results will provide a comprehensive insight VersaProbe II) was used to quantitatively characterize the microcrystalline
into the biochar modification with enhanced adsorption capacity for CO2 structures of the samples with Cu Kα radiation. The chemical bonding
capture. states of elements on the biochar surface were characterized using X-ray
photoelectron spectroscopy (XPS) performed on a Thermo Scientific
2. Materials and methods ESCALAB250xi (USA), with a monochromatic Al Kα source (Mono AlKα)
(hv = 1486.6 eV).
2.1. Preparation of biochar
2.3. CO2 adsorption mechanisms
Raw materials including pine wood (PW) and sewage sludge (SS)
were collected from South China Agricultural University and urban do- CO2 adsorption kinetics were performed using Thermogravimetric
mestic sewage treatment plant in Guangzhou, China. The elemental Analysis (TGA, Mettler Toledo, Switzerland) at 25 °C and 1 bar,
composition of feedstock was showed in Table S1. Pine wood (PW) according to Xu et al. (2020). Prior to CO2 adsorption, biochar samples
consisted of more C and O content (94%) compared to those in sewage were placed on the sample holder, then heated to 125 °C with N 2
sludge (35%). All the raw feedstocks were ground into powders with (50 mL/min) for 120 min. After the temperature of inner equipment
size of 0.5–1 mm and oven-dried overnight at 80 °C to remove the mois- was lowered to 25 °C, the N2 was switched to CO2 (50 mL/min). CO2 ad-
ture. The pyrolysis materials were mixed by the combination ratio of sorption isotherms were obtained using a volumetric sorption analyzer
70% pine wood and 30% sewage sludge. Two-step pyrolysis was (Quadrasorb, Quantachrome, USA) at room temperature (25 °C)
involved during the biochar production. The dried samples were first and up to atmospheric pressure (1 bar). Degasing was conducted at
pyrolyzed under nitrogen (N2 ) flow at a flow rate of 100 mL/min. 150 °C for 12 h under vacuum condition before the CO 2 adsorption
Then, the pyrolysis temperature was kept at 300 °C for 4 h and the measurements.
heating rate remained 10 °C/min. The adsorption isothem of N2 and CH4 was determined using an auto-
Prior to second pyrolysis, 10 g dried mixtures of activated biochar, mated gas sorption analyzer (Quadrasorb, Quantachrome, USA) at 25 °C,
10 g KOH, and an appropriate volume of deionized water were mixed 1 bar. According to Yuan et al. (2019), the Ideal Adsorbed Solution Theory
and continuously stirred for 4 h to attain a well-distributed mixture. (IAST) model was used to calculate the selectivity of CO2/N2 and CO2/CH4
Then the wet mixture was dried at 105 °C for 48 h until no vapor was for BC700K and BC800K. The recyclabilities of BC700K and BC800K were
emitted out. In rapid sequence, dried activated carbon was pyrolyzed also evaluated using an automated gas sorption analyzer. The same pre-
for 4 h under continuous flowing N2 (100 mL/min), where the increased treatment and stabilization conditions were employed for the investigation
heating temperature was 10 °C/min. According to the activation of cyclic stability. After pretreatment, 10 adsorption- desorption cycles
conditions and pyrolysis temperature of the feedstock, resultant biochar were carried out.
samples produced using 70% pine wood and 30% sewage sludge with The changes in surface functional groups on biochar surface were deter-
600 °C, 700 °C, and 800 °C were denoted as BC600, BC700, and BC800 mined using in-situ diffuse reflectance infrared Fourier transform (DRIFT)
respectively, while the modified biochars by KOH at 600 °C, 700 °C, spectroscopy (Nicolet Nexus 6700, Japan), equipped with a liquid nitrogen
and 800 °C were denoted as BC600K, BC700K, and BC800K, respec- cooled MCT detector, KBr windows and a heating chamber. The spectra
tively. All biochars were rinsed by 1 mol/L HCl, and then finely ground were collected in the range of 3000–500 cm−1. Before each adsorption
and passed through a 0.25 mm sieve. Finally, the sample was washed trial, the biochar samples were pretreated at 125 °C in N2 atmosphere for
with deionized water until the pH became neutral and then dried in an 120 min. Then N2 was switched to CO2 until the temperature of the inner

3
K. Li et al. Science of the Total Environment 826 (2022) 154133

equipment was lowered to 25 °C, and the adsorption trial lasted 15 min for The reason why the biochars pyrolyzed at 800 °C did not exhibit the opti-
modified biochars and 180 min for pristine biochar. mal textural surface properties was that biochar porous structure may be
blocked by the tar when pyrolyzed at high temperature, resulting in the
3. Results and discussion massive destruction of its micropore structure and the decrease in its sur-
face area (Jin et al., 2016).
3.1. Characteristics of biochar In addition, the pore size distribution of adsorbents was also an impor-
tant factor for CO2 adsorption behavior. According to the pore size distribu-
3.1.1. Specific surface area and pore size distribution tion analysis (Fig. 1), small pores with an average size of 0.56 nm was
The CO2 adsorption capacity of biochar is mainly attributed to the phys- observed for pristine biochars. By contrast, the average pore size of
icochemical properties of the biochar, such as the surface area, pore size, 0.5–2 nm for the KOH-activated biochars was observed, especially pores
and functional groups (Dissanayake et al., 2020a). According to the BET size of 0.59–0.72 nm. Meanwhile, the presence of a narrow pore size of
analysis (Table 1), the pore structure of the modified biochar was obviously 2.19 nm for modified biochars indicated the formation of highly tuneable
enhanced after KOH modification. With highly developed microporosity, nature of the biochars produced using KOH activation, verifying that
the specific surface area (SSA) of the KOH-modified biochars increased by these modified biochars were primarily composed of micropores and a con-
3.9–14.5 times. The SSA of BC600, BC700, and BC800 for pristine biochars siderable amount of mesopores. These results might be attributed to the
were 182, 223, and 150 m2/g, respectively, while the values for KOH- formed potassium species after KOH activation, diffusing into the internal
modified biochars, i.e., BC600K, BC700K and BC800K, were 703, 2623, structure of the biochar, which increased the width of pores and granted
and 2047 m2/g, respectively. Meanwhile, the KOH-modified biochars also biochars for more excellent pore structure (Mao et al., 2014). Remarkably,
exhibited larger total pore volume (0.29–0.90 cm3/g), compared to those the best micropore structure was observed in nature with pores size of
of pristine biochars (0.05–0.09 cm3/g). In addition, the similar tendency 0.5–2 nm for BC700K, implying that the BC700K might be the most desired
was observed for micropore area and micropore volume. Besides, higher adsorbent for CO2 adsorption among all types of biochars.
percentage of micropore volume was observed on KOH-modified biochars Fig. S1 shows the N2 adsorption behavior at −196 °C of six biochars. At
(72–83%) compared to pristine biochars (60–78%), implying that modified the low relative pressure, for all biochars, the N2 adsorption curves exhib-
biochars may exhibit higher CO2 adsorption capacity, since high micropore ited as steep as vertical, which may be due to the drastic interaction be-
volume (especially under 0.6 or 0.8 nm pore diameter) played a critical role tween the adsorbate and the pore wall of biochar. At a high relative
in the CO2 capture performance. These findings suggested that more pressure, the adsorbate displayed capillary condensation, and the curve be-
developed textural and structural surface properties were obtained for came flat and steady. Moreover, N2 adsorption of 92.8–110.1 cm3/g were
KOH-activated biochars, compared to their corresponding pristine ones. observed for the pristine biochars, while the values of 255.5–782.0 cm3/g
The reasons behind these activating effects exerted by KOH were mainly were found for the KOH-modified biochars at 1 bar, implying the develop-
due to the facts: (1) a plenty of carbons were exhausted, leading to a lattice ment of pore structure and enlarged surface area for modified biochars. Par-
defect during the redox reaction; (2) the carbon lattices was further pro- ticularly, the highest N2 adsorption capacity was observed for BC700K, due
moted, owing to the embedded K; (3) the gases, such as H2, CO and CO2, to its highest SSA (2623 m3/g) and total pore volume (0.90 cm3/g).
were produced during pyrolysis processes and escaped from the interior
of carbons, leading to the formation of abundant micropores (Lozano- 3.1.2. Elemental analysis
Castelló et al., 2007; Raymundo-Piñero et al., 2005; Singh et al., 2017). Table 2 depicts the elemental composition on the surface of biochars. As
Similarly, Singh et al. (2017) produced KOH-activated biochars derived pyrolysis temperature increased, the percentage of C in both pristine and
from Arundo donax at 600 °C and reported that higher SSA of 1122 m2/g modified biochars increased significantly, accounting for 60.34%–79.23%
and total pore volume of 0.592 cm2/g were obtained by the activated car- of the total elements, presumably due to the fact that the carbon of raw bio-
bons, compared to the pristine biochars (i.e., SSA of 16 m2/g and micropore mass was not volatilized and enriched massively during pyrolysis process.
volume of 0.02 cm3/g). Furthermore, the KOH activating process increased the oxygen content
Moreover, it was observed that higher SSA (i.e., 223 m2/g for BC700 (14.28%–25.89%) compare to pristine biochars (12.90%–17.53%), due to
and 2623 m2/g for BC700K) and greater pore volume (i.e., 0.09 cm3/g the incorporation of oxygen atoms derived from the KOH agent. By con-
for BC700 and 0.90 cm3/g for BC700K) were observed for the biochars py- trast, the decrease in carbon content (60.34%–78.72%) was observed for
rolyzed at 700 °C, compared to other biochars, suggesting that 700 °C may modified biochars in contrast to the pristine ones (69.95%–79.23%), pri-
be an appropriate pyrolysis temperature for activating the mixed material marily because KOH could contently dissolve ash and other compounds,
of 70% wood chips and 30% sewage sludge to achieve the most developed such as lignin and cellulose (Singh et al., 2017). In addition, the atomic ra-
textural and structural properties. Similarly, He et al. (2021) also indicated tios of O/C (0.14–0.32) and (N + O)/C (0.14–0.34) was increased signifi-
that the pyrolysis temperature of 750 °C was critical for forming accessible cantly after KOH activation, compared to the pristine biochars (0.12–0.18
surface and pore structure of biochar derived from wood waste feedstock. for O/C and 0.13–0.19 for (N + O) /C), implying the less aromatic and
hydrophobic structures observed in modified biochars compared to
pristine ones (Wang et al., 2015). Furthermore, the analysis of K content
of biochars suggested that more K content existed in modified biochar
Table 1 (3.91–5.99 mg/g) than pristine biochars (1.43–2.67 mg/g), which was con-
Textural properties of the biochars. BC600, BC700, and BC800 represent pristine ducive to CO2 adsorption (Fig. S2).
biochars pyrolyzed at 600 °C, 700 °C, and 800 °C, respectively, while BC600K,
BC700K, and BC800K represent KOH-activated biochars pyrolyzed at 600 °C, 700 3.1.3. SEM-EDS analysis
°C, and 800 °C, respectively.
The changes in surface morphology of biochars after KOH modifica-
Biochar Surface Total Average Micro Micropore Vmicro/Vtotal tion are presented in Fig. 2. The pristine biochars presented loosely and
area pore pore pore volume (%)
irregularly distributed pores (Fig. 2a–c). By contrast, after KOH activa-
(m2/g) volume diameter area (cm3/g)
(cm3/g) (nm) (m2/g) tion, more pores and fewer impurities were found on the surface of
KOH-modified biochar, being consistent with textural and structural
BC600 182 0.09 1.13 154 0.06 67
BC700 223 0.09 0.97 210 0.07 78 properties of the activated biochars. Furthermore, more polygonal and
BC800 150 0.05 1.17 141 0.03 60 regular shape was observed for KOH-modified biochar, and most of
BC600K 703 0.29 0.93 682 0.24 83 them were nano-sized, owing to the successful KOH modification
BC700K 2623 0.9 0.85 2087 0.74 82 (Fig. 2d–f). The moderate content of KOH could decompose the carbon
BC800K 2047 0.88 0.88 1663 0.63 72
walls connecting continuously oriented porous structure of the carbon

4
K. Li et al. Science of the Total Environment 826 (2022) 154133

a 0.6 0.16
BC600 0.25 BC700 BC800
0.14
0.5
0.12
0.20

dV(D) (cm3/g/nm)
0.4
dV(D) (cm3/g/nm)

dV(D) (cm3/g/nm)
0.10

0.15
0.08
0.3

0.06
0.10
0.2
0.04

0.1 0.05
0.02

0.00
0.0 0.00
-0.02
0 2 4 6 8 10 0 2 4 6 8 10 0 2 4 6 8 10
Pore width (nm) Pore width (nm) Pore width (nm)
1.2
0.30
BC600K 1.2 BC700K BC800K
0.25
1.0
1.0

dV(D) (cm3/g/nm)
0.8

dV(D) (cm3/g/nm)
0.20
dV(D) (cm3/g/nm)

0.8

0.15 0.6
0.6

0.10 0.4
0.4

0.05 0.2 0.2

0.00 0.0 0.0

0 2 4 6 8 10 0 2 4 6 8 10 0 2 4 6 8 10
Pore width (nm) Pore width (nm) Pore width (nm)

Fig. 1. Pore size distribution of a) BC600; b) BC700; c) BC800; d) BC600K; e) BC700K; and f) BC800K. BC600, BC700, and BC800 represent pristine biochars pyrolyzed at 600
°C, 700 °C, and 800 °C, respectively, while BC600K, BC700K, and BC800K represent KOH-activated biochars pyrolyzed at 600 °C, 700 °C, and 800 °C, respectively.

through oxidation at a high temperature, while a lot of potassium resi-


Table 2
Elemental composition of biochars. BC600, BC700, and BC800 represent pristine dues that were generated during the activation process completely
biochars pyrolyzed at 600 °C, 700 °C, and 800 °C, respectively, while BC600K, changed the surface morphology of the KOH-activated biochars (Singh
BC700K, and BC800K represent KOH-activated biochars pyrolyzed at 600 °C, 700 et al., 2017). These phenomena also suggested that the controlled acti-
°C, and 800 °C, respectively. vation process could grant activated biochars with regular morphology,
Biochar N (%) C (%) H (%) O (%) H/C O/C (N + O)/C continuously faciliating with porous structure and excellent surface
morphology. In addition, the EDS images (Fig. S3) reveals that a plenty
BC600 0.55 69.95 1.79 17.10 0.31 0.18 0.19
BC700 0.73 70.22 2.57 17.53 0.44 0.19 0.20
of carbon element was observed on the porous carbons surface of bio-
BC800 0.43 79.23 1.39 12.90 0.21 0.12 0.13 chars accompanying with a certain amount of oxygen for six biochars
BC600K 1.08 60.34 2.41 25.89 0.48 0.32 0.34 (Table 2). Meanwhile, more oxygen content and less carbon contents
BC700K 0.93 68.77 1.34 19.63 0.23 0.21 0.23 were observed in the activated biochars compared with corresponding
BC800K 0.43 78.72 1.24 14.28 0.19 0.14 0.14
pristine ones, being consistent with elemental analysis (Table 2).

Fig. 2. SEM images of a) BC600; b) BC700; c) BC800; d) BC600K; e) BC700K; and f) BC800K. BC600, BC700, and BC800 represent pristine biochars pyrolyzed at 600 °C, 700
°C, and 800 °C, respectively, while BC600K, BC700K, and BC800K represent KOH-activated biochars pyrolyzed at 600 °C, 700 °C, and 800 °C, respectively.

5
K. Li et al. Science of the Total Environment 826 (2022) 154133

Fig. 3. Curve fitting of Raman spectra a) BC600; b) BC700; c) BC800; d) BC600K; e) BC700K; and f) BC800K. BC600, BC700, and BC800 represent pristine biochars pyrolyzed
at 600 °C, 700 °C, and 800 °C, respectively, while BC600K, BC700K, and BC800K represent KOH-activated biochars pyrolyzed at 600 °C, 700 °C, and 800 °C, respectively.

3.1.4. Raman spectra analysis In addition, the intensity ratio of D band to G band (AD/AG) was calcu-
Fig. 3 illustrates the Raman spectra of six biochars in the region of lated to evaluate the degree of structural ordering of carbon materials
1000–2000 cm−1. Raman spectroscopic analysis reveals that two (Cong et al., 2015). High AD/AG ratio indicates high proportions of sp3
predominant peaks were observed at 1345–1350 cm−1 (D band) and carbon, owning to the presence of more structural defects, while low AD/
1600–1605 cm−1 (G band) for all biochars, which were attributed to the AG ratios reflects high degree of ordered carbon (Shao et al., 2017). It is
disordered graphitic structure and graphitic structure, respectively worth noting that the ratios of AD/AG for the biochars after KOH-
(Alabadi et al., 2014).Furthermore, compared to the pristine biochar, modification (i.e., 1.02, 1.27, 1.00 for BC600K, BC700K, BC800K, respec-
second-order weak bands were observed after KOH activation, such as GL tively) were decreased, compared to the corresponding pristine ones
band associated with carbonyl groups (BC600K and BC800K) and S band (i.e., 1.08, 1.37, 1.21 for BC600, BC700, BC800, respectively), implying
assigned to CH bonds of aromatic rings (BC600K and BC700K) (Li et al., the increasing tendency of the long-range order for the modified biochars.
2006). This could be further supported by SEM images, in which the modified

a b c
Intensity (a.u.)

Intensity (a.u.)

Intensity (a.u.)

C=C C=C
42.0% 39.8%
C=C
41.2%
C-OH C-OH
C-OH
33.5% 39.5%
31.5%

C-O C-O
C-O 27.3%
24.5% 24.7%

290 288 286 284 282 290 288 286 284 282 290 288 286 284 282
Binding Energy (eV) Binding Energy (eV) Binding Energy (eV)

d e f

C-OH
59.5%
Intensity (a.u.)
Intensity (a.u.)

Intensity (a.u.)

C-OH
45.8%

C=C
38.3% C=C
C-OH 35.1%
39.5%
C=C
28.8%
C-O C-O
22.2% C-O
19.1% 11.7%

290 288 286 284 282 290 288 286 284 282 290 288 286 284 282
Binding Energy (eV) Binding Energy (eV) Binding Energy (eV)

Fig. 4. C1s XPS spectra of biochars a) BC600; b) BC700; c) BC800; d) BC600K; e) BC700K; and f) BC800K. BC600, BC700, and BC800 represent pristine biochars pyrolyzed at
600 °C, 700 °C, and 800 °C, respectively, while BC600K, BC700K, and BC800K represent KOH-activated biochars pyrolyzed at 600 °C, 700 °C, and 800 °C, respectively.

6
K. Li et al. Science of the Total Environment 826 (2022) 154133

biochars exhibited more regular morphology, continuously oriented porous et al., 2020; Yang et al., 2018). Although all the biochars possessed the
structure and enhanced surface morphology. Similarly, Dissanayake et al. same types of functional groups, a decrease in CC bond and an increase in
(2020b) investigated the structural features of pristine and KOH-activated C-OH content were observed for the modified biochars, reinforcing the
biochars produced from the mixture of 40% food waste and 60% wood, changes observed in carbon and oxygen elements after KOH modification.
and reported that the lower AD/AG ratio (1.58) was observed for the In addition, XRD were complementarily performed to investigate the
KOH-activated biochar, compared to that of pristine biochar (1.84). crystallite morphology of biochars. As shown Fig. S5, several sharp peaks
were observed in XRD pattern of the pristine biochars, verifying the
3.1.5. XPS and XRD analysis presence of crystallized minerals (Yuan et al., 2011), which were mainly
The surface chemical compositions and element bonding states of bio- originated from sewage sludge in the raw feedstock mixtures. Similarly,
chars was analyzed using XPS (Fig. S4). Two prominent peaks, i.e. C1s Gao et al. (2019) produced biochars derived from sewage sludge at
and O1s peaks, were manifested in the survey spectra Meanwhile, it is evi- 500 °C and found the presence of crystallized mineral peaks in XRD
denced that a decrease in C1s peak intensity was observed, presumably due analysis. By contrast, no significant peaks were observed for the modified
to the massive lignin and cellulose dissolved by KOH. By contrast, O1s peak biochars, suggesting that the impurities of crystalline carbons or potassium
intensity exhibited an increasing tendency after KOH activation, mainly residues were not present (Chen et al., 2016), since KOH could corrode
owing to the introduction of oxygen atoms onto the surface of the carbon crystalline mineral and decompose the carbon walls, resulting in the
during KOH activation (Shahkarami et al., 2015). This finding also indi- oriented porous structure after KOH modification. This finding is also con-
cated that KOH activation exerted significant influence on the enhance- sistent with SEM analysis, implying that the KOH-modified biochars were
ment of the porous structure and surface chemical elements of biochars. granted for more regular porous structure compared to the pristine ones.
Moreover, the analysis of high-resolution C1s deconvoluted XPS spectra
was also employed for better understanding the presence of surface oxygen- 3.2. CO2 adsorption behavior and mechanism
containing groups (Fig. 4). The three major peaks at around 284.6, 285.0,
286.3 eV were observed for all the biochars, associated with COH, CC, 3.2.1. CO2 adsorption kinetics
and CO bond, respectively (Dissanayake et al., 2020b; Igalavithana A significant difference in the kinetic behavior between the pristine and
modified biochars was observed (Fig.5). It took more than 180 min for the
pristine biochar to reach the equilibrium, while the adsorption of CO2 on
a modified biochar composites exhibited faster progress, although the modi-
180
fied biochars (i.e., BC500K, BC700K, and BC800K) exhibited similar kinetic
CO2 adsorption capacity(mg/g)

160
behaviors compared with pristine ones. For the first 5 min, CO2 adsorption
140 rate for modified biochar was high with the adsorption amount increased
almost vertically as a function of time. Thereafter the rate was slowed
120
down and finally reached a plateau (Fig. 5). This dramatic increase in
100 CO2 adsorption rates for KOH-modified biochars indicated that the possible
BC600 transition of adsorption mechanism shifted from physical sorption to chem-
80
BC700 ical adsorption and subsequent carbonation process (Xu et al., 2020). Fur-
60 BC800 thermore, although it took longer time for BC700K and BC800K to reach
BC600K CO2 adsorption equilibrium compared to BC600K, it still remained less
40 BC700K
than 15 min. The fact that the observed slowdown of CO2 adsorption rate
BC800K
20 for BC700K compared to BC600K was probably attributed to its relatively
larger micropores, which was not favorable for CO2 diffusion to some ex-
0
tent (Song et al., 2016).
0 10 20 30 40 50 60
Time(min) 200
BC600
b 60 BC700
CO2 adsorption capacity(mg/g)

BC800
CO2 adsorption capacity(mg/g)

150
50 BC600K
BC700K
40
BC800K
100
30

20 50

10 BC600K
BC700K
BC800K 0
0

0 50 100 150 200 250 0.0 0.2 0.4 0.6 0.8 1.0

Time(min) Pressure(bar)

Fig. 5. Gravimetric CO2 adsorption of biochars at 25 °C and 1 bar. BC600, BC700, Fig. 6. CO2 adsorption isotherms at 25 °C. BC600, BC700, and BC800 represent
and BC800 represent pristine biochars pyrolyzed at 600 °C, 700 °C, and 800 °C, pristine biochars pyrolyzed at 600 °C, 700 °C, and 800 °C, respectively, while
respectively, while BC600K, BC700K, and BC800K represent KOH-activated BC600K, BC700K, and BC800K represent KOH-activated biochars pyrolyzed at
biochars pyrolyzed at 600 °C, 700 °C, and 800 °C, respectively. 600 °C, 700 °C, and 800 °C, respectively.

7
K. Li et al. Science of the Total Environment 826 (2022) 154133

3.2.2. CO2 adsorption isotherms proportionally with the cumulative micropore volume for pore sizes less
The analysis for CO2 adsorption isotherms is illustrated in Fig. 6. The than 0.7–1.0 nm (Casco et al., 2014; Choi et al., 2019). Moreover, the
CO2 adsorption capacity for BC600, BC700, and BC800 were 35.5, 42.9, highest pore size of <0.8 nm observed on BC700K (Fig. 1) also resulted in
and 42.7 mg/g, respectively, while the values for BC600K, BC700K and the highest CO2 adsorption capacity. This finding has been attributed the
BC800K were 136.7, 182.0, 141.7 mg/g, respectively. It is worth noting fact that CO2 capture was a process of filling pore at 25 °C and 0–1 bar, in
that KOH activation increased the CO2 adsorption capacity of biochars by which the narrow pore size played a predominant role in CO2 adsorption.
nearly 5-folds, compared to the corresponding pristine biochars, since the And narrow micropores <0.8 nm was proven to be the most effective for
activation process significantly changes the chemical composition and tex- CO2 adsorption at this low relative pressure (Igalavithana et al., 2020).
tural properties of biochars (Igalavithana et al., 2020). The excellent CO2 Therefore, KOH activation can be considered as an effective modification
capture ability for the modified biochar underscores the merits of KOH ac- method for CO2 capture, and BC700K would be the optimal alternative
tivation of biomass for better CO2 adsorption. Furthermore, the surface among six biochars.
chemistry of the KOH-modified biochar subject to pore structure, including Furthermore, the oxygen contents in the modified biochars exhibited in-
a large SSA and micropore volume, contributed to the enhanced CO2 ad- creasing tendency (14.28%–25.89%) in contrast to pristine ones (12.90%–
sorption performance, as they not only provided more available CO2 ad- 17.53%) (Table 2), due to the incorporation of oxygen atoms with KOH
sorption sites, but also promoted the interaction between the CO2 during activating process. It has been evidenced that the presence of
molecules (Table 2) (Singh et al., 2017). Similarly, Creamer et al. (2014) in- oxygen-containing acidic functional groups (e.g., hydroxyl groups, carboxyl
vestigated the correlation between CO2 capture capacity and basic biochar groups, and carbonyl groups) could also enhance CO2 adsorption on carbo-
properties, and found that the surface area was a significant factor deter- naceous surfaces by facilitating hydrogen bonding between the CO2 mole-
mining the CO2 adsorption capacity for biochars (Pearson's r = 0.82). cules and the carbon surface (Liu et al., 2018; Xing et al., 2014). For
Moreover, during the KOH activation process, different potassium species, instance, Liu et al. (2018) reported that the highest CO2 adsorption of
e.g. K2O and K2CO3, were formed and diffused into the internal structure of 170.3 mg/g was obtained by the carbon material, which was oxidated at
the biochar matrix, which increased the width of the existing mesopores 50 °C to achieve high percentage of oxygen, despite the absence of highest
and enhanced the generation of new mesopores (Mao et al., 2014; Otowa micro pore volume, further emphasizing the crucial role of oxygen content
et al., 1997). played in CO2 adsorption. According to XPS analysis (Fig. 4), the modified
It was evidenced that owing to the highest surface area (2623 m3/g) and biochars exhibited more oxygen-containing functional groups compared to
largest pores volume (0.90 cm3/g), as well as the enhanced porous mor- the pristine ones, reinforcing the fact that higher CO2 adsorption capacity
phology, BC700K exhibited the highest CO2 capacity (182.0 mg/g), com- observed in the modified biochars.
pared to other modified biochars (136.7–141.7 mg/g), reinforcing that In addition, the adsorption of CO2 could be induced by mineralogical re-
the surface area and pores volume were crucial for physical adsorption of actions, owing to the presence of minerals on the surface of biochar, such as
CO2. Interestingly, it was observed that only within a specific range of potassium. More potassium content in modified biochars after KOH activa-
CO2 pressure, large SSA could facilitate with the enhancement of CO2 ad- tion may further enhance the CO2 adsorption capacity, since the presence of
sorption. Among the cumulative micropore volumes with various pore alkali metals and alkaline earth metals (e.g., Na, K, Ca, Mg, and Li) could
sizes, the narrow micropore volume with pore sizes less than 0.8 nm exhib- promote the formation of basic sites with strong affinity to CO2, which ex-
ited the best correlation with CO2 uptake at 1 bar. Previous studies indi- hibited an acidic nature (Creamer et al., 2016; Xu et al., 2016). Similarly,
cated that the CO2 adsorption by porous carbons at 1 bar increased Xu et al. (2016) also emphasized the significant role of the presence of

Fig. 7. In situ DRIFTS (600–3000 cm−1) of a) BC600; b) BC700; c) BC800; d) BC600K; e) BC700K; and f) BC800K at 25 °C during CO2 sorption. BC600, BC700, and BC800
represent pristine biochars pyrolyzed at 600 °C, 700 °C, and 800 °C, respectively, while BC600K, BC700K, and BC800K represent KOH-activated biochars pyrolyzed at 600 °C,
700 °C, and 800 °C, respectively.

8
K. Li et al. Science of the Total Environment 826 (2022) 154133

alkali or alkali earth metals played in the CO2 adsorption by the modified 1457 cm−1was detected for both BC700 and BC800, associated with CH2
biochars derived from sewage sludge, wheat straw, and pig manure. In in aliphatic hydrocarbon chain (Lin et al., 2017). Owing to developed
their works, Fe(OH)2CO3 was formed in the biochar derived from sewage pore structure of modified biochars pyrolyzed at appropriate temperature,
sludge through the transformation of FeOOH after CO2 sorption, whereas physical adsorption of modified biochars could be enhanced compared to
K2Ca(CO3)2 and CaMg(CO3)2 were the transformation products in the bio- pristine ones. Furthermore, more functional groups appeared after KOH ac-
char oriented from pig manure after CO2 sorption. Furthermore, the reac- tivation and relatively high peak intensity subject to carbonates indicated
tion between adsorbed CO2 and calcium carbonate (CaCO3) resulted in that chemical adsorption was also strengthened in contrast to pristine bio-
the formation of Ca(HCO3)2 in the case of the biochar produced by wheat chars. As a whole, in situ DRIFTS revealed that both physical and chemical
straw. This phenomenon suggested that the chemical adsorption could adsorption was actively involved in CO2 remediation by biochar.
also be predominant in CO2 adsorption process. Meanwhile, since other
elements including phosphorus and sulfur existed in the tested biochars 3.3. Adsorption-desorption cycles and IAST selectivity analysis
with low levels (Table 2), it is speculated that such minor elements played
a trivial role in CO2 capture. In addition, biochar with hydrophobic The regeneration also plays a vital role in selecting the optimal biochars
and non-polar characteristics may also facilitate the CO2 adsorption for CO2 capture, in addition to the adsorption capacities, since it is closely
(Dissanayake et al., 2020a). According to elemental analysis discussed related to the economic benefits of the involved industry. To further inves-
above, the KOH-modified biochar surface exhibited fewer hydrophobic tigate the regeneration, BC700K and BC800K were chosen owing to their
structures with higher polarity compared to the pristine ones, which high CO2 adsorption capacities. As shown in Fig. 8, both BC700K and
might result in undesirable affinity to CO2. However, KOH-modified bio- BC800K exhibited excellent CO2 adsorption capacities even after 10 cycles
chars still exhibited higher CO2 adsorption capacities, implying that except of adsorption-desorption. Particularly, high CO2 retention of 98% was ob-
for hydrophobic and polarity structures, the physico-chemical properties of served for BC800K after the 10 cycles of adsorption-desorption, with the cy-
the biochar, e.g. surface area, pore size, pore volume, the presence of sur- clic capacity decrease from 140.7 to 138.1 mg/g. In comparison, the cyclic
face functional groups and alkali metals (K), were predominant factors for capacity of BC700K decreased from 187.2 to 181.3 mg/g, with 97% reten-
enhancing CO2 adsorption. tion. The reason for the slight decline in CO2 adsorptive capacity might
In situ DRIFTS analysis was performed to investigate transformation of have been attributed to the fact that during the desorption process, some
CO2 species on the surface of biochar during CO2 adsorption. Fig. 7 displays
the in situ DRIFTS spectra (500–3000 cm−1) of CO2 adsorbed on biochars
during the whole CO2 adsorption equilibrium process at 25 °C. The absorp- a 160
tion region between 1018 and 1360 cm−1 was detected for all biochars,
which was assigned to CO stretching vibrations (Li et al., 2017). Further- 140
CO2 adsorption capacity(mg/g)

more, a very intense absorption peak at 2360–2380 cm−1 was observed


120
for all biochars, associated with the asymmetric stretching (v3) of
physically-sorbed CO2, and the intensity increased as a function of adsorp-
100
tion time, reinforcing the fact that physical adsorption played a vital role in
CO2 capture. Similarly, Xu et al. (2020) compared the behavior of CO2 80
capture between pristine and ball-milled biochars, and reported that
with in situ DRIFT spectroscopy, the intensive peak in the region of 60
2330–2380 cm−1 was observed for the biochar modified with Fe
oxyhydroxide, assigned to asymmetric stretching (v3) of physically- 40
sorbed CO2. In addition, all the biochars except BC600K exhibited the
weak peak intensity at 1551–1744 cm−1, which was ascribed to the forma- 20
tion of carbonates (Lara-García et al., 2019). The transformation of CO2 by
biochar composites and in situ characterization of weakly adsorbed and 0

newly formed mineral species during the CO2 sorption process suggested 1 2 3 4 5 6 7 8 9 10
that chemical adsorption was actively involved in CO2 capture process. In- Cycles
terestingly, these weak peaks occurred on the surface of all biochars even
b 200
after the biochars were rinsed with HCl solution, implying that chemical ad-
sorption was actively involved in CO2 adsorption on biochars surface. The
CO2 adsorption capacity(mg/g)

weak peak related to carbonates on the biochar surface also suggested


that the effect of CO2 chemical adsorption was shadowed by physical ad- 150
sorption to some extent. Noteworthy, although the functional groups sub-
ject to carbonates were not observed on BC600K surface, higher CO2
adsorption capacity was still obtained for BC600K, compared to pristine
biochars, further indicating the importance of physisorption during CO2 100
capture. Meanwhile, the vibrations in the carbonate ions (i.e., CO2− 3 )
were observed at 1416 cm−1 on the surface of BC600K, highlighting the
important role of chemical adsorption played in CO2 capture (Shaltout
et al., 2011). The importance of chemical adsorption for CO2 capture was 50
also emphasized by Xu et al. (2016), who found carbonates were generated
on the biochars derived from pig manure, e.g., K2Ca(CO3)2 and CaMg(CO3)
2, which were the transformation products during CO2 adsorption process.
0
As for BC700K and BC800K, the absorption peaks ranging from 667 cm−1
to 675 cm−1 were observed, associated with CO2 bending absorption 1 2 3 4 5 6 7 8 9 10
(Hosakun et al., 2017). In particular, an increasing intensity of peak at Cycles
1444 cm−1 was observed for BC700K as a function of time, assigned to
anti-symmetric in-plane bending of CH3 (Silva et al., 2014), which was Fig. 8. Cyclic adsorption-desorption tests. a) BC800K; b) BC700K. BC700K, BC800K
presumably associated with CO2 adsorption. Additionally, a peak at represent KOH-activated biochars pyrolyzed at 700 °C, and 800 °C, respectively.

9
K. Li et al. Science of the Total Environment 826 (2022) 154133

a
160 BC700K b
BC700K
BC800K 70
BC800K
140
60

IAST selectivity

IAST selectivity
120

50
100

40
80

60 30

40 20
0.0 0.2 0.4 0.6 0.8 1.0 1.2 0.0 0.2 0.4 0.6 0.8 1.0 1.2
Pressure (bar) Pressure (bar)

c d 550
BC700K
30 BC700K
BC800K 500
BC800K
450
25
IAST selectivity

400

IAST selectivity
350
20
300

250
15
200

150
10
100

0.0 0.2 0.4 0.6 0.8 1.0 1.2 0.0 0.2 0.4 0.6 0.8 1.0 1.2
Pressure (bar) Pressure (bar)

Fig. 9. IAST selectivities of BC700K and BC800K at 25 °C for a) CO2/CH4:30/70; b) CO2/CH4:50/50; c) CO2/CH4:70/30; d) CO2/N2:15/85. BC700K, BC800K represent KOH-
activated biochars pyrolyzed at 700 °C, and 800 °C, respectively.

chemically bound CO2 molecules on the surface of the biochars were not 4. Conclusion
completely released (Yaumi et al., 2018). Yaumi et al. (2018) reported
a stable performance with 8.8% decrease in CO2 adsorptive capacity In this work, a novel KOH–activate nanoporous biochars derived from
by activated carbon RHPM1 after 12 cycles and indicated that the the mixed sewage sludge and pine sawdust were successfully synthesized,
decrease was probably due to blockages of pores by some chemically which exhibited extremely high specific surface area and microporosity,
bonded CO2 molecules during the desorption process. Furthermore, in as well as abundant functional groups for CO2 adsorption. Particularly,
situ DRIFTS analysis highlighted the important role of chemical adsorp- the modified biochars exhibited higher surface area (703–2623 m2/g)
tion involved in CO2 adsorption, whereas the carbonates generated on compared to pristine ones (150–223 m2/g). The development of micropo-
the biochars could not be completely desorbed by N2 despite 6 cycles rous structures on the surface of the modified biochars via KOH activation
of adsorption-desorption, Despite slightly lower CO2 retention com- significantly accelerated the CO2 adsorption rates. Moreover, the adsorp-
pared to BC800K, BC700K possessed supreme CO2 adsorption capacity tive capacity of CO2 by KOH-activated biochar increased by 3.3–4.2 folds,
with high regeneration potential. and the maximum CO2 adsorption capacity was up to 182.0 mg/g.
Furthermore, IAST selectivity for CO2/N2 and CO2/CH4 in binary gas Meanwhile, the high percentage of hetero atoms (i.e., O and K) and func-
mixtures was determined for BC700K and BC800K (Fig. 9). According to tional groups on surface of modified biochars also contributed to high
CO2/CH4 molar ratio in biogas and CO2/N2 molar ratio in flue gas streams, CO2 adsorption. Furthermore, in situ DRIFTS revealed that both physical
the selectivity subject to CO2/CH4 (v/v; 30/70, 50/50, and 70/30) and and chemical adsorption was actively involved in the CO2 adsorption. In
CO2/N2 (v/v; 15/85) were calculated. Both BC700K and BC800K shared addition, BC700K exhibited ultrahigh CO2/CH4 and CO2/N2 IAST selectiv-
similar selectivity trends for CO2/N2 and CO2/CH4, presumably due to ities, and the excellent cyclic regeneration property with 97% retention was
their similar physicochemical properties discussed above. According to observed after 6 consecutive adsorption-desorption cycles at 25 °C, 1 bar.
Fig. 9a–c, the IAST selectivity between CO2 and CH4 for BC700K This study provided theoretical basis for the development of new mixed
(11.6–63.5) and BC800K (9.7–52.7) was high at 1.0 bar, due to more avail- biochar for CO2 capture system, bridging the current knowledge gap in
able sites that could strongly adsorb CO2 over CH4. The similar tendency the roles of physical and chemical adsorption, and would help develop a
was also observed for the selectivity between CO2 and N2, in which the better CO2 absorbent for the global warming.
highest IAST selectivity was 122.2 and 115.8 for BC700K and BC800K at
1.0 bar, respectively. However, with the increasing pressure, the number CRediT authorship contribution statement
of available sites for CO2 adsorption might be occupied, resulting in the de-
crease in the selectivity (Choi et al., 2019). Remarkably, the selectivity of Kai Li: Methodology, Data curation, Writing-original draft, Writing-
BC800K was always inferior to BC700K. Combined with the persistent reus- review & editing, Visualization. Dongqing Zhang: Writing-review &
ability, BC700K was considered as the most promising adsorbent for CO2 editing. Xiaojun Niu and Huafang Guo: Resources and Supervision.
capture among all the biochars. Zhang Lin, zhihua Tang, Yuanyuan Yu and Mingli Fu: Date curation.

10
K. Li et al. Science of the Total Environment 826 (2022) 154133

Declaration of competing interest Jahandar, L.M., Khiavi, S., Sayari, A., 2019. Stability of amine-functionalized CO2 adsorbents:
a multifaceted puzzle. Chem. Soc. Rev. 48 (12), 3320–3405.
Jalilov, A.S., Li, Y., Tian, J., Tour, J.M., 2017. Ultra-high surface area activated porous asphalt
The authors declare that they have no known competing financial inter- for CO2 capture through competitive adsorption at high pressures. Adv. Energy Mater. 7
ests or personal relationships that could have appeared to influence the (1), 1600693.
Jin, J., Li, Y., Zhang, J., Wu, S., Cao, Y., Liang, P., Zhang, J., Wong, M.H., Wang, M., Shan, S.,
work reported in this paper. Christie, P., 2016. Influence of pyrolysis temperature on properties and environmental
safety of heavy metals in biochars derived from municipal sewage sludge. J. Hazard.
Acknowledgement Mater. 320, 417–426.
Kamkeng, A.D.N., Wang, M., Hu, J., Du, W., Qian, F., 2021. Transformation technologies for
CO2 utilisation: current status, challenges and future prospects. Chem. Eng. J. 409,
This work was supported by the National Natural Science Foundation of 128138.
China (No. 42177369), National Key Research and Development Program Kończak, M., Pan, B., Ok, Y.S., Oleszczuk, P., 2020. Carbon dioxide as a carrier gas and mixed
feedstock pyrolysis decreased toxicity of sewage sludge biochar. Sci. Total Environ. 723,
of China (No. 2019YFA0210400), the Science and Technology Planning
137796.
Project of Maoming, China (No. 2019S002), and the Fundamental Research Lahijani, P., Mohammadi, M., Mohamed, A.R., 2018. Metal incorporated biochar as a poten-
Funds for the Central Universities, SCUT (No. 2020ZYGXZR105). tial adsorbent for high capacity CO2 capture at ambient condition. J. CO₂ Util. 26,
281–293.
Lara-García, H.A., Gao, W., Gómez-Cortés, A., Diaz, G., Pfeiffer, H., Wang, Q., 2019. High and
Appendix A. Supplementary data efficient CO2 capture in molten nitrate-modified mg-Al-palmitate layered double oxides
at high pressures and elucidation of carbonation mechanisms by in situ DRIFT spectros-
Supplementary data to this article can be found online at https://doi. copy analysis. Ind. Eng. Chem. Res. 58 (14), 5501–5509.
Leng, L., Xu, S., Liu, R., Yu, T., Zhuo, X., Leng, S., Xiong, Q., Huang, H., 2020. Nitrogen con-
org/10.1016/j.scitotenv.2022.154133. taining functional groups of biochar: an overview. Bioresour. Technol. 298, 122286.
Li, X., Hayashi, J.I., Li, C., 2006. FT-raman spectroscopic study of the evolution of char struc-
References ture during the pyrolysis of a victorian brown coal. Fuel 85 (12–13), 1700–1707.
Li, J., Chen, F., Yang, L., Jiang, L., Dan, Y., 2017. FTIR analysis on aging characteristics of
ABS/PC blend under UV-irradiation in air. Spectrochim. Acta A Mol. Biomol. Spectrosc.
Ahmed, M.B., Hasan Johir, M.A., Zhou, J.L., Ngo, H.H., Nghiem, L.D., Richardson, C., Moni, 184, 361–367.
M.A., Bryant, M.R., 2019. Activated carbon preparation from biomass feedstock: clean Lin, H., Wang, S., Zhang, L., Ru, B., Zhou, J., Luo, Z., 2017. Structural evolution of chars from
production and carbon dioxide adsorption. J. Clean. Prod. 225, 405–413. biomass components pyrolysis in a xenon lamp radiation reactor. Chin. J. Chem. Eng. 25
Ahmed, R., Liu, G., Yousaf, B., Abbas, Q., Ullah, H., Ali, M.U., 2020. Recent advances in (2), 232–237.
carbon-based renewable adsorbent for selective carbon dioxide capture and separation- Lin, R.B., He, Y., Li, P., Wang, H., Zhou, W., Chen, B., 2019. Multifunctional porous hydrogen-
a review. J. Clean. Prod. 242, 118409. bonded organic framework materials. Chem. Soc. Rev. 48 (5), 1362–1389.
Alabadi, A., Yang, X., Dong, Z., Li, Z., Tan, B., 2014. Nitrogen-doped activated carbons derived Liu, B., Li, H., Ma, X., Chen, R., Wang, S., Li, L., 2018. The synergistic effect of oxygen-
from a co-polymer for high supercapacitor performance. J. Mater. Chem. A 2 (30), containing functional groups on CO2 adsorption by the glucose–potassium citrate-
11697–11705. derived activated carbon. RSC Adv. 8 (68), 38965–38973.
Alvarez, J., Amutio, M., Lopez, G., Bilbao, J., Olazar, M., 2015. Fast co-pyrolysis of sewage Lozano-Castelló, D., Calo, J.M., Cazorla-Amorós, D., Linares-Solano, A., 2007. Carbon activa-
sludge and lingo cellulosic biomass in a conical spouted bed reactor. Fuel 159, 810–818. tion with KOH as explored by temperature programmed techniques, and the effects of hy-
Casco, M.E., Martínez-Escandell, M., Silvestre-Albero, J., Rodríguez-Reinoso, F., 2014. Effect drogen. Carbon 45 (13), 2529–2536.
of the porous structure in carbon materials for CO2 capture at atmospheric and high- Mao, H., Zhou, D., Hashisho, Z., Wang, S., Chen, H., Wang, H.H., 2014. Preparation of
pressure. Carbon 67, 230–235. pinewood- and wheat straw-based activated carbon via a microwave-assisted potassium
Chen, T., Liu, R., Scott, N.R., 2016. Characterization of energy carriers obtained from the py- hydroxide treatment and an analysis of the effects of the microwave activation condi-
rolysis of white ash, switchgrass and corn Stover-biochar, syngas and bio-oil. Fuel tions. Bioresources 10 (1), 809–821.
Process. Technol. 142, 124–134. Otowa, T., Nojima, Y., Miyazaki, T., 1997. Development of KOH activated high surface area
Choi, S.W., Tang, J., Pol, V.G., Lee, K.B., 2019. Pollen-derived porous carbon by KOH activa- carbon and its application to drinking water purification. Carbon (New York) 35 (9),
tion: effect of physicochemical structure on CO2 adsorption. J. CO₂ Util. 29, 146–155. 1315–1319.
Cong, H., Zhang, M., Chen, Y., Chen, K., Hao, Y., Zhao, Y., Feng, L., 2015. Highly selective Peter, S.C., 2018. Reduction of CO2 to chemicals and fuels: a solution to global warming and
CO2 capture by nitrogen enriched porous carbons. Carbon 92, 297–304. energy crisis. ACS Energy Lett. 3 (7), 1557–1561.
Creamer, A.E., Gao, B., Zhang, M., 2014. Carbon dioxide capture using biochar produced from Raymundo-Piñero, E., Azaïs, P., Cacciaguerra, T., Cazorla-Amorós, D., Linares-Solano, A.,
sugarcane bagasse and hickory wood. Chem. Eng. J. 249, 174–179. Béguin, F., 2005. KOH and NaOH activation mechanisms of multiwalled carbon nano-
Creamer, A.E., Gao, B., Wang, S., 2016. Carbon dioxide capture using various metal tubes with different structural organisation. Carbon 43 (4), 786–795.
oxyhydroxide–biochar composites. Chem. Eng. J. 283, 826–832. Shahkarami, S., Azargohar, R., Dalai, A.K., Soltan, J., 2015. Breakthrough CO2 adsorption in
Dissanayake, P.D., You, S., Igalavithana, A.D., Xia, Y., Bhatnagar, A., Gupta, S., Kua, H.W., bio-based activated carbons. J. Environ. Sci. (China) 34, 68–76.
Kim, S., Kwon, J., Tsang, D.C.W., Ok, Y.S., 2020a. Biochar-based adsorbents for carbon Shaltout, A.A., Allam, M.A., Moharram, M.A., 2011. FTIR spectroscopic, thermal and XRD
dioxide capture: a critical review. Renew. Sust. Energ. Rev. 119, 109582. characterization of hydroxyapatite from new natural sources. Spectrochim. Acta A Mol.
Dissanayake, P.D., Choi, S.W., Igalavithana, A.D., Yang, X., Tsang, D.C.W., Wang, C., Kua, Biomol. Spectrosc. 83 (1), 56–60.
H.W., Lee, K.B., Ok, Y.S., 2020b. Sustainable gasification biochar as a high efficiency ad- Shao, J., Ma, F., Wu, G., Dai, C., Geng, W., Song, S., Wan, J., 2017. In-situ MgO (CaCO3)
sorbent for CO2 capture: a facile method to designer biochar fabrication. Renew. Sust. templating coupled with KOH activation strategy for high yield preparation of
Energ. Rev. 124, 109785. various porous carbons as supercapacitor electrode materials. Chem. Eng. J. 321,
Dutta, S., He, M., Xiong, X., Tsang, D.C., 2021. Sustainable management and recycling of food 301–313.
waste anaerobic digestate: a review. Bioresour. Technol. 341, 125915. Shao, J., Zhang, J., Zhang, X., Feng, Y., Zhang, H., Zhang, S., Chen, H., 2018. Enhance SO2
Esposito, E., Dellamuzia, L., Moretti, U., Fuoco, A., Giorno, L., Jansen, J.C., 2019. Simulta- adsorption performance of biochar modified by CO2 activation and amine impregnation.
neous production of biomethane and food grade CO2 from biogas: an industrial case Fuel 224, 138–146.
study. Energy Environ. Sci. 12 (1), 281–289. Silva, S.D., Feliciano, R.P., Boas, L.V., Bronze, M.R., 2014. Application of FTIR-ATR to
Fang, S., Yu, Z., Ma, X., Lin, Y., Chen, L., 2018. Analysis of catalytic pyrolysis of municipal moscatel dessert wines for prediction of total phenolic and flavonoid contents and antiox-
solid waste and paper sludge using TG-FTIR, py-GC/MS and DAEM ( distributed activa- idant capacity. Food Chem. 150, 489–493.
tion energy model ). Energy 143, 517–532. Singh, G., Kim, I.Y., Lakhi, K.S., Srivastava, P., Naidu, R., Vinu, A., 2017. Single step synthesis
Gao, L., Deng, J., Huang, G., Li, K., Cai, K., Liu, Y., Huang, F., 2019. Relative distribution of of activated bio-carbons with a high surface area and their excellent CO2 adsorption ca-
Cd2+ adsorption mechanisms on biochars derived from rice straw and sewage sludge. pacity. Carbon 116, 448–455.
Bioresour. Technol. 272, 114–122. Song, G., Zhu, X., Chen, R., Liao, Q., Ding, Y., Chen, L., 2016. An investigation of CO2 adsorp-
Hashem, A., Badawy, S.M., Farag, S., Mohamed, L.A., Fletcher, A.J., Taha, G.M., 2020. Non- tion kinetics on porous magnesium oxide. Chem. Eng. J. 283, 175–183.
linear adsorption characteristics of modified pine wood sawdust optimised for adsorption Wang, Q., Luo, J., Zhong, Z., Borgna, A., 2010. CO2 capture by solid adsorbents and their ap-
of Cd(II) from aqueous systems. J. Environ. Chem. Eng. J. 8 (4), 103966. plications: current status and new trends. Energy Environ. Sci. 4 (1), 42–55.
He, M., Xu, Z., Sun, Y., Chan, P.S., Lui, I., Tsang, D.C., 2021. Critical impacts of pyrolysis con- Wang, Z., Liu, G., Zheng, H., Li, F., Ngo, H.H., Guo, W., Liu, C., Chen, L., Xing, B., 2015. Inves-
ditions and activation methods on application-oriented production of wood waste- tigating the mechanisms of biochar’s removal of lead from solution. Bioresour. Technol.
derived biochar. Bioresour. Technol. 341, 125811. 177, 308–317.
Hosakun, Y., Halász, K., Horváth, M., Csóka, L., Djoković, V., 2017. ATR-FTIR study of the in- Wang, X., Deng, S., Tan, H., Adeosun, A., Vujanovic, M., Yang, F., Duic, N., 2016. Synergetic
teraction of CO2 with bacterial cellulose-based membranes. Chem. Eng. J. 324, 83–92. effect of sewage sludge and biomass co-pyrolysis: a combined study in thermogravimetric
Huang, D., Deng, R., Wan, J., Zeng, G., Xue, W., Wen, X., Zhou, C., Hu, L., Liu, X., Xu, P., Guo, analyzer and a fixed bed reactor. Energy Convers. Manag. 118, 399–405.
X., Ren, X., 2018. Remediation of lead-contaminated sediment by biochar-supported Wang, Z., Xie, L., Liu, K., Wang, J., Zhu, H., Song, Q., Shu, X., 2019. Co-pyrolysis of sewage
nano-chlorapatite: accompanied with the change of available phosphorus and organic sludge and cotton stalks. Waste Manag. 89, 430–438.
matters. J. Hazard. Mater. 348, 109–116. Wei, H., Chen, H., Fu, N., Chen, J., Lan, G., Qian, W., Liu, Y., Lin, H., Han, S., 2017. Ex-
Igalavithana, A.D., Choi, S.W., Dissanayake, P.D., Shang, J., Wang, C., Yang, X., Kim, S., cellent electrochemical properties and large CO2 capture of nitrogen-doped acti-
Tsang, D.C.W., Lee, K.B., Ok, Y.S., 2020. Gasification biochar from biowaste (food vated porous carbon synthesised from waste longan shells. Electrochim. Acta 231,
waste and wood waste) for effective CO2 adsorption. J. Hazard. Mater. 391, 121147. 403–411.

11
K. Li et al. Science of the Total Environment 826 (2022) 154133

Wei, Q., Xu, J., Yang, S., Liao, L., Jin, G., Li, Y., Hameed, F., 2018. Subsurface watering re- Yaumi, A.L., Bakar, M.Z.A., Hameed, B.H., 2017. Recent advances in functionalized composite
sulted in reduced soil N2O and CO2 emissions and their global warming potentials solid materials for carbon dioxide capture. Energy 124, 461–480.
than surface watering. Atmos. Environ. 173, 248–255. Yaumi, A.L., Bakar, M.Z.A., Hameed, B.H., 2018. Melamine-nitrogenated mesoporous acti-
Xing, W., Liu, C., Zhou, Z., Zhou, J., Wang, G., Zhuo, S., Xue, Q., Song, L., Yan, Z., 2014. vated carbon derived from rice husk for carbon dioxide adsorption in fixed-bed. Energy
Oxygen-containing functional group-facilitated CO2 capture by carbide-derived carbons. 155, 46–55.
Nanoscale Res. Lett. 9 (1), 189. Yuan, J., Xu, R., Zhang, H., 2011. The forms of alkalis in the biochar produced from crop res-
Xiong, Z., Shihong, Z., Haiping, Y., Tao, S., Yingquan, C., Hanping, C., 2013. Influence of idues at different temperatures. Bioresour. Technol. 102 (3), 3488–3497.
NH3/CO2 modification on the characteristic of biochar and the CO2 capture. Bioenerg. Yuan, H., Chen, J., Li, D., Chen, H., Chen, Y., 2019. 5 ultramicropore-rich renewable porous
Res. 6 (4), 1147–1153. carbon from biomass tar with excellent adsorption capacity and selectivity for CO2 cap-
Xu, X., Kan, Y., Zhao, L., Cao, X., 2016. Chemical transformation of CO2 during its capture by ture. Chem. Eng. J. 373, 171–178.
waste biomass derived biochars. Environ. Pollut. 213, 533–540. Zeng, Y., Zou, R., Zhao, Y., 2016. Covalent organic frameworks for CO2 capture. Adv. Mater.
Xu, X., Xu, Z., Gao, B., Zhao, L., Zheng, Y., Huang, J., Tsang, D.C.W., Ok, Y.S., Cao, X., 2020. 28 (15), 2855–2873.
New insights into CO2 sorption on biochar/Fe oxyhydroxide composites: kinetics, mech- Zhang, X., Wu, J., Yang, H., Shao, J., Wang, X., Chen, Y., Zhang, S., Chen, H., 2016. Prepara-
anisms, and in situ characterization. Chem. Eng. J. 384, 123289. tion of nitrogen-doped microporous modified biochar by high temperature CO2–NH3
Yahya, M.A., Al-Qodah, Z., Ngah, C.W.Z., 2015. Agricultural bio-waste materials as potential treatment for CO2 adsorption: effects of temperature. RSC Adv. 6 (100), 98157–98166.
sustainable precursors used for activated carbon production: a review. Renew. Sust.
Energ. Rev. 46, 218–235.
Yang, X., Igalavithana, A.D., Oh, S., Nam, H., Zhang, M., Wang, C., Kwon, E.E., Tsang, D.C.W.,
Ok, Y.S., 2018. Characterization of bioenergy biochar and its utilization for metal/
metalloid immobilization in contaminated soil. Sci. Total Environ. 640–641, 704–713.

12

You might also like