Accepted Manuscript: Bioresource Technology Reports

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 28

Accepted Manuscript

Influence of one-step and two-step KOH activation on activated


carbon characteristics

Oluwatosin Oginni, Kaushlendra Singh, Gloria Oporto, Benjamin


Dawson-Andoh, Louis McDonald, Edward Sabolsky

PII: S2589-014X(19)30156-2
DOI: https://doi.org/10.1016/j.biteb.2019.100266
Article Number: 100266
Reference: BITEB 100266
To appear in: Bioresource Technology Reports
Received date: 18 April 2019
Revised date: 10 June 2019
Accepted date: 17 June 2019

Please cite this article as: O. Oginni, K. Singh, G. Oporto, et al., Influence of one-step
and two-step KOH activation on activated carbon characteristics, Bioresource Technology
Reports, https://doi.org/10.1016/j.biteb.2019.100266

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.
ACCEPTED MANUSCRIPT

Influence of One-step and Two-step KOH Activation on Activated Carbon Characteristics

1
Oluwatosin Oginni, 1Kaushlendra Singh, 1Gloria Oporto, 1Benjamin Dawson-Andoh,
2
Louis McDonald, 3Edward Sabolsky

T
IP
CR
1
School of Natural Resources, West Virginia University, Morgantown, WV 26506
2
Division of Plant and Soil Sciences, West Virginia University, Morgantown WV 26506

US
3
Department of Mechanical Engineering, West Virginia University, Morgantown WV 26506
AN
M
ED

Corresponding author: Oluwatosin Oginni


Email: ojoginni@mix.wvu.edu
PT

Phone: 334-444-8227
Contact Address: 322 Percival Hall, School of Natural Resources,
CE

West Virginia University, Morgantown, WV 26506


AC

1
ACCEPTED MANUSCRIPT

Abstract
This study investigated the influence of one-step and two-step KOH activation on the
characteristics of activated carbons. The biomass precursors [Kanlow Switchgrass (KS) and
Public Miscanthus (PM)] and their biochars produced at pyrolysis temperature of 500 ºC, were
impregnated with KOH and activated at a temperature 900 ºC for 30 min in an inert condition.
BET surface area, pore characteristics, surface functional groups and microstructures of the
activated carbons were examined. The biomass-derived activated carbons (KOH-KSBM and

T
KOH-PMBM) had the highest BET surface areas of 1272 and 1597 m2/g respectively. The

IP
activated carbons were used in adsorbing acetaminophen and caffeine and the biomass-derived

CR
activated carbons presented an overall better adsorption characteristic.

US
Keyword: Activated carbon, adsorption, biomass, biochar, BET surface area
AN
M

Highlight
ED

1. The biochar-derived activated carbons had visible pore structure


2. The biomass-derived activated carbons had the highest BET surface areas
PT

3. The Raman spectra of all the activated carbons exhibited the two signature G and D
bands
CE
AC

2
ACCEPTED MANUSCRIPT

1. Introduction
Activated carbon as defined by International Union of Pure and Applied Chemistry (IUPAC), is
a porous carbon material which has been subjected to reaction with gases, sometimes with the
addition of chemicals before, during or after carbonization in order to increase its adsorptive
properties (McNaught & Wilkinson, 1997). Activated carbon has a well-developed porous
structure and surface chemical functionalities for enhanced interaction with polar and non-polar

T
adsorbates (Bansal & Goyal, 2005). The presence of emerging contaminants in waste and

IP
drinking water and the improved ability to detect them at very low concentrations has heightened

CR
concerns regarding its long-term consequences to human health and has led to greater interest in
the application of activated carbon for water treatment (Donn et al., 2008; Oginni, 2018).

US
Álvarez-Torrellas et al. (2017) reported the use of activated carbons derived from peach stones
and rice husks in adsorbing non-biodegradable pharmaceuticals from hospital wastewater. Lam
AN
et al. (2019) also reported the use of activated carbon for removal of herbicides (2,4-
dichlorophenoxyacetic) from surface water in agricultural land.
M

Activated carbons are synthesized from various carbon rich precursors such as petroleum pitch,
lignite, coal, lignocellulosic biomass, agricultural, forest and municipal wastes by physical or
ED

chemical activation (Dittmar et al., 2018). The physical activation is a two-step procedure
involving the thermochemical conversion of the precursor to char via carbonization reaction in
PT

an inert atmosphere, followed by activation of the char using an oxidizing gas such as CO2,
steam or air at a high temperature (Mazlan et al., 2016). The chemical activation involves the
CE

impregnation of the precursor with chemical activating agent, followed by activating the
impregnated precursor at an elevated temperature. The chemical activation is a more versatile
AC

process in comparison to the physical activation because it allows the synthesis of activated
carbon with a tunable porosity (made up of micropores and mesopores) and surface chemistry
(Fuertes et al., 2018).
Among the chemical activating agents employed for chemical synthesis of activated carbon,
potassium hydroxide (KOH) has been reported to produce highly microporous activated carbons
and enhance the formation of -OH functional groups on the activated carbon surface (Hui &
Zaini, 2015). The effectiveness of KOH activation relative to other chemical activating agents
can be attributed to the ability of K to form intercalation compounds with carbon easily. In

3
ACCEPTED MANUSCRIPT

addition, the K2O formed during the process of KOH activation is reduced to K by carbon,
resulting in carbon gasification with a subsequent emission of CO2 leading to the formation of
pores (Viswanathan et al., 2009).
The chemical activation procedure can be divided into a one-step or two-step activation. The
one-step activation involves the impregnation of biomass precursor with the chemical activating
agent and subjecting the impregnated biomass precursor to activation/carbonization. Umran et al.
(2015) investigated the impregnation of Holm Oak with KOH and carbonizing the impregnated

T
biomass precursor at temperature range of 550 to 750 ºC for 1 h. The authors reported that the

IP
higher percentage of the total pore volume of the activated carbons were micropore. Liu et al.

CR
(2018) reported the preparation of activated carbon from KOH impregnation of coconut shell and
activating the impregnated precursor at a temperature of 700 ºC. The N2 adsorption isotherm of

US
the activated carbon was a type V with H4 hysteresis, which is an indication that the pores of the
activated carbon are slit-shaped. Also, the surface area and pore volume of the activated carbon
AN
were reported to be 860 m2/g and 0.427 cm3/g respectively. The micropore volume of the
activated carbon was 0.315 cm3/g (about 74% of the total pore volume), which is an indication
M

that the activated carbon is microporous in nature. Villota et al. (2019) carried out a study on the
microwave-assisted activation of KOH impregnated waste cocoa pod husk using an
ED

impregnation ratio (KOH/biomass) of 0.5 – 2.0. The maximum BET surface area and pore
volume reported were 494.51 m2/g and 0.258 cm3/g respectively. The activated carbons were
PT

identified to be microporous as a higher proportion of the surface area and the total pore volume
were micropores.
CE

The two-step chemical activation involves the initial carbonization of the biomass precursor to
produce biochar at a temperature range of 300 – 600 ºC. The thermal decomposition leads to
AC

release of volatile content in vapors, tars, and gases and creation of pores in the resulting biochar.
The deposition of some tar materials fills or partially block the created pores (Bansal & Goyal,
2005), resulting in biochar with low surface area. Oginni & Singh (2019) reported a BET surface
area of 0.38 m2/g for biochar produced from pyrolysis of Arundo donax at a pyrolysis
temperature of 500 ºC. The second step of the chemical activation involves impregnation of the
biochar precursor with chemical activating agent, followed by activating the impregnated biochar
precursor at a temperature range of 700 – 1200 ºC. The activation removes tar substances from
the pore spaces and create microporosity in the carbon structure (Bansal & Goyal, 2005).

4
ACCEPTED MANUSCRIPT

Marques et al. (2018) investigated the KOH impregnation of biochar derived from apple tree
small branches and activation of the impregnated biochar precursor at a temperature of 800 ºC
for 1 – 4 h. The activated carbons had high surface areas ranging between 2114 – 2472 m2/g,
micropore volume corresponding to 91 – 98% of the total pore volume and the pHpzc values
showed that the activated carbons were neutral or slightly acidic. Qiu et al. (2018) conducted a
study on KOH impregnation of corn straw derived biochar obtained at a pyrolysis temperature of
450 ºC and activated at a temperature of 800 ºC for 1 h. The BET surface areas of the activated

T
carbons were between 1475 – 2790 m2/g. The authors noted that the mesopore surface areas and

IP
volumes significantly increased with an increase in the impregnation ratio, which is an indication

CR
that the micropores collapsed and merged into larger mesopores when the impregnation ratio
increased. The surface chemistry of the activated carbons showed the presence of oxygen

US
functional groups which enhances the wettability of the activated carbon surface. Zhang et al.
(2017) prepared biochar based activated carbon derived from hemp stem using KOH as an
AN
activating agent with an impregnation ratio of 4.5/1 (KOH/biochar) and activated at a
temperature of 800 ºC for 1.5 h. The activated carbon presented a BET surface area of 2388 m 2/g
M

and 88.98 % of the total pore volume as micropore.


The use of KOH in the chemical synthesis of biomass-derived and biochar-derived activated
ED

carbons result in a variation of textural, microstructural and surface chemical characteristics.


This necessitate investigating a comparison of the single and two-steps chemical activation on
PT

the characteristics of the activated carbon. Elmouwahidi et al. (2017) investigated the
characteristics of activated carbons obtained from the one-step and two-step chemical synthesis
CE

of olive residue using KOH as activating agent. The biochar precursor was produced at a
temperature of 300 ºC and the activation of the biomass and biochar precursors was carried out at
AC

a temperature of 840 ºC for 2 h. The result showed that at a low impregnation ratio, the biomass-
derived activated carbon had a higher BET surface area of 1390 m2/g compared to the biochar-
derived activated carbon having a surface area of 1295 m2/g. Meanwhile, as the impregnation
ratio increased, the BET surface area of the biomass-derived activated carbon decreased to 1350
m2/g and the biochar-derived activated carbon had a higher BET surface area of 1626 m2/g. The
total pore and mesopore volumes of the biochar-derived activated carbons were higher than the
biomass-derived activated carbons. However, both the biochar-derived and biomass-derived
activated carbons presented a similar surface chemistry.

5
ACCEPTED MANUSCRIPT

Therefore, the objective of this study is to investigate the effect of one-step and two-step KOH
activation on the characteristics of activated carbons synthesized from herbaceous biomass.
Secondly, the performance of the biomass derived and biochar derived activated carbons in
adsorbing pharmaceutically active compounds was investigated.

2. Materials and Methods


2.1 Materials

T
Kanlow Switchgrass and Public Miscanthus biomass feedstocks were harvested from a reclaimed

IP
mine land in West Virginia. The biomass samples were chopped into small sizes using a kitchen

CR
meat knife and oven dried at 103 ºC for 24 h. The oven-dried samples were subsequently ground
to less than 1 mm particle size using a Retsch Grindomix (Model: GM 200). Additionally, the

US
biochar samples were prepared from the oven-dried biomass via pyrolysis. The detailed
description of the pyrolysis process is provided in a previous work of the authors (Oginni et al.,
AN
2017). Briefly, the pyrolysis was conducted in a fixed-bed batch reactor having a diameter of
22.86 cm and a height of 25.40 cm. The reactor was filled with oven-dried biomass and the
M

sample filled reactor was placed in a furnace (Model: BF51728C, Thermo scientific, NC). The
furnace was heated from room temperature to 500 ºC under a nitrogen gas flow of 2 L/min and
ED

kept a that temperature for 30 min. The pyrolysis vapor generated during the pyrolysis process
was swept into a series of ice-bath condensers. After the pyrolysis ended, the reactor was
PT

allowed to cool to room temperature under a constant flow of nitrogen gas. After cooling, the
biochars were collected from the reactor.
CE

The pharmaceutically active compounds used for the adsorption experiment, were
Acetaminophen purchased from Acros Organics, Morris, NJ and Caffeine purchased from
AC

Fischer Science Education, Nazareth, PA. The choice of acetaminophen and caffeine for the
adsorption test was because these compounds are among the prevalent emerging contaminants
found in wastewater treatment plants, rivers, streams, surface water and treated waters in the
United States (Ebele et al., 2017). The physicochemical characteristics of the compounds are
presented in the Supplementary Data.
2.2 Sample preparation
The biomass and biochar precursors were impregnated with potassium hydroxide KOH (100%
wt.) at an impregnation ratio of 1:1 (weight basis). The samples were soaked in 200 ml solution

6
ACCEPTED MANUSCRIPT

of KOH and continuously stirred inside a pressure reactor (Model: 4500, Parr Instrument
Company, Moline, IL, USA) at a temperature of 85 ºC for 24 h. After the impregnation was over,
the impregnated precursors were oven dried at a temperature of 103 ºC for 24 h. The oven-dried
impregnated precursors were thereafter activated using a thermogravimetric analyzer (Model:
TGA 701, LECO Corporation, St. Joseph, MI) at a temperature of 900 ºC for 1 h under nitrogen
flow to keep the process inert. In addition, biochars from both biomass feedstocks were prepared
at a temperature of 900 ºC using the same procedure that was used for the activation of the

T
impregnated precursors.

IP
The activated carbons were washed with dilute hydrochloric acid and deionized distilled water.

CR
The washed activated carbons were thereafter oven-dried at a temperature of 103 ºC for 24 h.
The labels and description for the activated carbons are presented in Table 1.

US
Table 1: Labels and description for activated carbons
s/n Label Full Description of Activated Carbon Label
1 KOH-KSBM Activated carbon from KOH impregnated Kanlow Switchgrass Biomass
AN
2 KOH-KSBC Activated carbon from KOH impregnated Kanlow Switchgrass Biochar
3 KOH-PMBM Activated carbon from KOH impregnated Public Miscanthus Biomass
M

4 KOH-PMBC Activated carbon from KOH impregnated Public Miscanthus Biochar


5 KS900 Kanlow Switchgrass Biochar at 900 ºC
ED

6 PM900 Public Miscanthus Biochar at 900 ºC


PT

2.3 Microstructural characterization


2.3.1 Morphology
CE

The solid morphology characterization was performed at the West Virginia University Shared
Research Facilities using a scanning electron microscope (Model: Hitachi-S4700, Hitachi High
AC

Technologies America, Schaumburg, IL). The sample was placed on a sample holder and then
transferred into the sample chamber in the equipment. The instrument software was used in
viewing and collecting the SEM images of the samples.
2.3.2 Surface area and pore characteristics
The surface area and pore characteristics of the activated carbons were determined by nitrogen
adsorption at 77 K using Accelerated Surface Area and Porosimetry System (Model: ASAP
2020, Norcross, GA, USA). Prior to the analysis, the samples were degassed at a temperature of
105 ºC for 24 h and cooled down to 30 ºC. The N2 adsorption isotherms were obtained over a

7
ACCEPTED MANUSCRIPT

relative pressure, p/p0, ranging from 0.01 to 0.99. The surface area, total pore volume, micropore
volume and average pore diameter were determined by the application of the Brunauer-Emmett-
Teller (BET) model and t-plot models. The pore size distribution was calculated from the
nitrogen adsorption isotherm with the free version of the SAIEUS program using the Non-Local
Density Function Theory (NLDFT) model (Kupgan et al., 2017).
2.3.3 XPS and Raman analysis
The determination of the surface functional groups was carried out using X-ray photoelectron

T
spectroscopy (Model: PHI 5000 VersaProbe XPS/UPS, ULVAC-PHI Inc., Kanagawa, Japan) in

IP
a spectral range of 0 – 1400 eV binding energy and energy resolution of 0.50 eV. For the XPS

CR
analysis, each sample was mounted on a steel specimen disk using an Ultra High Vacuum
approved spectral grade double-sided carbon tape and loaded into the introduction chamber.

US
Afterwards, the sample was transferred to the analysis chamber where the photoelectron spectra
were acquired. The element identification and the peak fitting was carried out using the PHI
AN
MultiPak software.
The Raman spectra was obtained using a Raman microscope (Model: Renishaw InVia) equipped
M

with an excitation laser of a wavelength of 532 nm. The laser beam was operated on a 1% laser
power input and at a 50x optical magnification. The sample scanning was within a spectrum
ED

range of 1000 – 2000 cm−1.


2.4 Determination of adsorbate concentrations in solution
PT

Standard solutions were prepared by dissolving 10 mg of each adsorbate in 50 ml distilled


deionized water to give a concentration of 200 ppm. From this solution, varying volumes were
CE

taken and adjusted with the distilled deionized water to give the desired concentration.
The selection of the appropriate wavelength for the adsorbate was carried out by scanning the
AC

standard solution in UV spectrophotometer (Model: Varian Cary 50) between 200 nm and 400
nm on the spectrum mode, using deionized water as a blank. The absorbance recorded for the
solution at a specific wavelength, is proportional to its concentration, which is based on Beer-
Lambert principle. The maximum wavelength (λmax) at which the solution showed the peak
absorbance were selected. The λmax of 242 nm and 274 nm were identified for acetaminophen
and caffeine respectively. A calibration curve was obtained in the concentration range of 0.10 –
50 ppm for both adsorbates. The plot of the concentration versus the absorbance (Supplementary

8
ACCEPTED MANUSCRIPT

Data) gave a straight-line curve and the linear regression equations had correlation coefficients
of 0.98 and 0.99 for acetaminophen and caffeine respectively.
2.4.1 Adsorption kinetics
To study the adsorption kinetics, 10 mg of activated carbon was added to 40 ml of adsorbate
solution in glass vial with an initial concentration of 40 ppm. The mixtures were agitated in a
multipoint agitation plate at a temperature of 25 °C. The time recording started when the
agitation began and several samples were collected between 15 min and 9 h. The collected

T
samples were filtered through a filter paper (Fisher Scientific, Qualitative P4) to separate the

IP
activated carbon and the adsorbate solution. The filtered adsorbate solution was analyzed in the

CR
UV-Vis spectrophotometer for concentration determination. From the initial concentration and
concentration values at time ‘t’, the amount of adsorbate uptake by the activated carbon was

US
calculated according to the following equation;

1
AN
Where qt is the amount (mg/g) of acetaminophen or caffeine adsorbed at time t, Co is the initial
concentration (ppm), Ct is the concentration at time t (ppm), V is the volume (ml) of the
M

adsorbate solution and W is the weight (mg) of the activated carbon used. Additionally, the qt
data for various durations (t) were fitted into the pseudo-second order kinetic model expressed
ED

as;
2
PT

Where qe is the maximum adsorption capacity for the pseudo-second order kinetic model, k2 is
CE

the equilibrium rate constant for the pseudo-second order kinetic model. Values of k2 and qe were
estimated from the intercept and slope, respectively, of the plot of t/qt versus t. The product k2qe2,
(designated as h in this study) represents the initial adsorption rate. The half-life time, t1/2, defines
AC

the time required for the activated carbon to uptake half of the adsorbate amount that will be
retained at equilibrium (Mestre et al., 2007). It is used as a measure of rate of adsorption and is
determined from the equation 3;
3

2.4.2 Adsorption equilibrium


10 mg of activated carbon was mixed with 40 ml solution of the adsorbate solution having initial
concentrations of 10 – 40 ppm and agitated for 5 h. After the agitation, the concentration of the

9
ACCEPTED MANUSCRIPT

adsorbate solution at equilibrium was determined. The agitation time of 5 h was selected for the
equilibrium assays based on the results of the adsorption kinetics which showed that there was
no difference in the uptake of the adsorbates between 5 and 9 h. All the experiments were carried
out in triplicate without pH adjustment.
The initial and final concentrations of adsorbate solution were used to calculate qe (t = 5 h) using
equation 1. The qe and equilibrium concentration (Ce) data were fitted in Langmuir and
Freundlich models. The linear form of the Langmuir isotherm model is provided in equation 4

T
(Mestre et al., 2007);

IP
4

CR
where qe = equilibrium quantity adsorbed (mg/g); Ce = equilibrium concentration of adsorbate
(ppm); qm = maximum adsorption capacity (mg/g); b = Langmuir constant. qm and b were

US
determined from the linear plot of Ce/qe versus qe. The Langmuir isotherm model is based on the
assumptions that, during an entire adsorption process, only monolayer adsorption occurs, the
AN
adsorbed adsorbates are attached to the adsorbent’s surface at definite localized sites with no
adsorbate-adsorbate interactions, and the adsorbent has an energetically homogenous surface
M

(Bansal & Goyal, 2005).


The linear form of the Freundlich model is shown in equation 5 (Mestre et al., 2007);
ED

where Kf and n are indicators of adsorption capacity and adsorption intensity or surface
PT

heterogeneity, respectively. The values of n and Kf were obtained from the slope and intercept of
CE

the linear plot of lnqe versus lnCe. The Freundlich isotherm assumes that the ratio of the
adsorbate adsorbed onto an adsorbent is not constant at different solution concentrations and the
amount adsorbed increases infinitely with increasing concentration (Foo & Hameed, 2010).
AC

2.5 Statistical analysis


The adsorption experiments were carried out in triplicate. A significance test (p < 0.05) was
conducted on the kinetics and equilibrium parameters using Proc GLM procedure in SAS
University Edition. Tukey’s test in SAS was used in comparing the means of the parameters. The
results are presented using the means and standard deviations of the triplicate runs.

10
ACCEPTED MANUSCRIPT

3. Results and Discussion


3.1 Microstructure
The microstructural characterization performed by SEM (Figure 1) shows the porous structures
of the activated carbons. The SEM images of the biochar-derived activated carbons (Figure 1A
& 1B) showed that these activated carbons have visible pores of varying sizes and still retained
the cell wall structures of the parent materials. The open pore structure makes it easy for
adsorbate to access maximum surface area of the activated carbon through enhanced diffusion

T
(Djilani et al., 2015). However, the biomass-derived activated carbons (KOH-KSBM and KOH-

IP
PMBM) shown in Figure 1C & 1D, did not show any distinctive pore structure, which is an

CR
indication that the original cell wall structures of the parent materials were not retained after
impregnation and activation. The non-distinctive porous structures exhibited by the biomass-

US
derived activated carbons can be attributed to the destruction of the lignin components of the
biomass precursors during the impregnation process (Oh et al., 2003). The lignin, which consists
AN
of many ester and ether bonds in crosslinked structure helps to sustain the morphology of the
biomass and upon impregnation with KOH, the lignin component was severely destroyed by
M

KOH (Oh et al., 2003). Also, during the activation process, KOH decompose to form inorganic
K, which reacts with carbon and accelerates the gasification of the impregnated biomass, thus
ED

destroying cell wall structure (Viswanathan et al., 2009). Oh et al. (2003) reported a similar
finding for the activated carbon obtained from the impregnation of Rice straw with KOH prior to
PT

activation. This effect was not evident in the biochar-derived activated carbons because the
lignin content had been thermally decomposed during the biochar production. The biochars
CE

(KS900 and PM900) shown in Figure 1E & 1F also have visible pores of varying sizes and
retained the original cell wall structure of the parent materials.
AC

11
ACCEPTED MANUSCRIPT

A C E
C

T
B D F

IP
CR
US
AN
Figure 1: SEM Images of (A) KOH-KSBC, (B) KOH-PMBC, (C) KOH-KSBM, (D) KOH-
PMBM, (E) KS900, (F) PM900
M

3.2 BET surface area and pore characteristics


Figure 2 shows the N2 adsorption-desorption isotherms for the biochars and the activated
ED

carbons. Based on the IUPAC classification of adsorption isotherms (Sing et al., 1985), the
adsorption isotherms for the biochars (Figure 2A) and the biochar-derived activated carbons
PT

(Figure 2B) showed a mixture of type I and type IV isotherms with a hysteresis loop, indicating
the presence of mesopores. The biomass-derived activated carbons (Figure 2C) had a type I
CE

isotherm, which is a typical representation of microporous materials. The difference in


adsorption-desorption isotherms of the biomass-derived and the biochar-derived activated
AC

carbons can be attributed to the initial carbonization of the biochar precursor and subsequent
KOH activation which helped to improve its textural properties (Elmouwahidi et al., 2017).
It is noteworthy that the adsorption hysteresis exhibited by the biochars and the biochar-derived
activated carbons (Figure 2A & 2B) were more prominent in comparison to the biomass-derived
activated carbons (Figure 2C). The adsorption hysteresis exhibited by the biochars and the
activated carbons had also been identified to have a wide variety of shapes, therefore, IUPAC
(Sing et al., 1985) classified this adsorption hysteresis into four classes (H1 – H4). Based on this

12
ACCEPTED MANUSCRIPT

classification, the adsorption hysteresis exhibited by the biochars and the activated carbons can
be classified as H4, which is associated with narrow slit-shaped pores.

T
IP
CR
US
AN
M
ED
PT
CE

Figure 2: N2 adsorption/desorption isotherms for (A) Biochars, (B) Biochar-derived


activated carbons, (C) Biomass-derived activated carbons
AC

Table 2 shows the BET surface, pore volume and average pore diameters of the biochars and
activated carbons. The BET surface areas ranged between 519 and 1597 m2/g. The KS900 and
PM900 biochars had the lowest BET surface areas of 519.49 and 783.74 m2/g, respectively. The
biochar-derived activated carbons (KOH-KSBC and KOH-PMBC) showed an improved BET
surface area of 599.19 m2/g and 957.51 m2/g respectively. The biomass-derived activated
carbons were found to have the highest BET surface areas of 1271.66 and 1596.52 m 2/g for
KOH-KSBM and KOH-PMBM, respectively. This implies that the biomass impregnation route

13
ACCEPTED MANUSCRIPT

led to activated carbons with high surface areas. This effect can be attributed to chemical
reaction between the chemical components of the biomass and KOH during the subsequent
carbonization. The impregnation of the biomass precursors with KOH resulted in the removal of
lignin and partial hemicellulose (Sun & Cheng, 2002). According to Mestre et al. (2014), the
KOH reaction mechanism involves the formation of K2CO3, H2O, CO2, CO and H2 at a relatively
low temperature (~ 400 ºC) and some of these compounds are commonly used as physical
activating agents. The K2CO3 decomposes at temperatures between 700 and 800 ºC, resulting in

T
the formation of elemental K, which remains intercalated in the carbon structure. The elemental

IP
K is removed during the washing of the activated carbon, hence unblocking the pore network.

CR
Table 2: Surface Area and Pore Characteristics of Biochars and Activated Carbons
Sample SBET Smicro Smicro/SBET Vtotal Vmicro Vmeso dp
(m2/g) (m2/g) (cm3/g) (cm3/g) 3

US
(%) (cm /g) (nm)
PM900 783.74 557.21 71.10 0.41 0.24 0.17 2.07
KS900 519.49 416.42 80.16 0.25 0.18 0.07 1.92
AN
KOH-PMBC 957.51 747.19 78.04 0.45 0.33 0.12 1.88
KOH-KSBC 599.19 504.80 84.25 0.28 0.22 0.06 1.86
M

KOH-PMBM 1596.52 534.29 33.47 0.89 0.23 0.66 2.23


KOH-KSBM 1271.66 647.66 50.93 0.63 0.28 0.35 1.97
ED

*dp: Average pore diameter (4V/A by BET); V: Pore volume; S: Surface area
PT

The micropore surface areas (Smicro) of the activated carbons ranged from 416 to 747 m2/g (Table
2). The micropore surface area describes the proportion of the total surface of the activated
CE

carbon that are microporous. The percentage proportion of the micropore surface area to the total
BET surface area was calculated for each sample. The biochar-derived activated carbons (KOH-
AC

PMBC and KOH-KSBC) showed an increase in their micropore surface areas in comparison to
the biochars (PM900 and KS900). However, there was a drastic reduction in the micropore
surface areas for the biomass-derived activated carbons This corroborate the explanation
provided for the hysteresis loops exhibited by the activated carbons whereby the biomass-derived
activated carbons showed narrow hysteresis loops in comparison to the hysteresis loops exhibited
by the biochars and the biochar-derived activated carbons.
The total pore volumes of the biochars and the activated carbons ranged between 0.25 and 0.89
cm3/g (Table 2). The total pore volumes of the biomass-derived activated carbons were higher

14
ACCEPTED MANUSCRIPT

than the total pore volumes of biochar-derived activated carbons and the biochars. The
impregnation and activation of the biochar precursors also led to a slight increase in their total
pore volumes in comparison to the total pore volumes of the biochars.
The total pore volume is divided into two, namely the micropore and the mesopore. Mesopore,
also regarded as transport pore, is very important property of the activated carbon. It influences
the adsorption kinetics in any liquid adsorption. From Table 2, the biomass-derived activated
carbons were seen to have the highest mesopore volume of 0.35 and 0.66 cm3/g for KOH-KSBM

T
and KOH-PMBM respectively. Mestre et al. (2009) reported that activated carbon possessing

IP
negligible mesopore presents the slowest adsorption kinetics while the fastest adsorption rate

CR
occurred with activated carbon having the most developed mesopore network.
The average pore diameters (dp) of the activated carbons were between 1.86 and 2.23 nm

US
respectively (Table 2). The average pore diameter defines the ability of the adsorbate molecules
to penetrate inside the activated carbon and interact with its inner surface. It is an indication of
AN
the mean size of both the pore openings with relatively narrow portions and the pore bodies with
relatively wide portions (Nimmo, 2005). The average pore diameters of the biomass-derived
M

activated carbons were higher than the ones for biochar-derived activated carbons and the
biochars. The difference can be explained to be that the biochar precursors had an ordered porous
ED

structure, such that the effect of the activating agent on improving the porous structure of the
resulting activated was minimal (Bazan-Wozniak et al., 2017). Juejun et al. (2013) stated that
PT

biochar prepared at high carbonization temperature possesses a more densely packed structure,
from which it is difficult to enlarge the small pores to larger pores.
CE

The pore size distribution determines the fractions of the total pore volume that are either
microporous or mesoporous (Pelekani & Snoeyink, 1999). The pore size distributions of the
AC

biochars and the activated carbons are shown in Figure 3. The biochars and the activated carbons
can be seen to have monomodal distributions between pore width range of 0 and 4 nm. The
peaks of the distributions are within 0 and 2 nm, which is in the microporous range. This implies
that majority of the pores that exist in the biochars and the activated carbons are microporous in
nature. The biochars and the biochar-derived activated carbons exhibited narrow pore size
distributions, meanwhile the biomass-derived activated carbons can be seen to have broader pore
size distribution across the pore width range. This is an indication that the impregnation of the
biochar and subsequent activation did not greatly influence the pore size distribution of the

15
ACCEPTED MANUSCRIPT

resulting activated carbons in comparison to the biochar. However, for the biomass-derived
activated carbons, the pore size distributions seem to be more improved in comparison to the
biochars.
0.8 PM900 KS900
KOH-PMBC KOH-PMBM
0.7 KOH-KSBC KOH-KSBM
Pore Size Distribution (cc/g/nm)

0.6

T
IP
0.5

CR
0.4 Microporosity Mesoporosity

0.3

0.2
US
AN
0.1
M

0
0 1 2 3 4 5
Pore Width (nm)
ED

Figure 3: Pore size distributions showing the microporosity and mesoporosity of the
biochars, biomass-derived and biochar-derived activated carbons
PT

3.3 Raman analysis


The Raman spectra for all the samples exhibited the two signature G and D bands
CE

(Supplementary Data), which is an indication of their heterogenous carbon microstructure


(Yumak et al., 2018). The D-band is attributed to in-plane vibration of sp2 bonded carbon with
AC

structural defects or disorderness while the G-band arises from the in-plane vibration of the sp2
bonded graphitic carbon structure (Zhao et al., 2013). The Raman signals for all the samples
showed an overlapping of the two bands, which is an indication of the presence of high
proportion of amorphous carbon structures (Guizani et al., 2017). The D-band positions for all
the samples were in the range of 1349 and 1365 cm-1 while the G-band positions ranged between
1592 and 1598 cm-1. These values are similar to the typical band positions reported in literatures
for disordered carbon materials such as biochars and activated carbons (Atta-Obeng et al., 2017;

16
ACCEPTED MANUSCRIPT

Zhao et al., 2013). The presence of both peaks in all the samples is an indication of C sp2 atoms
in benzene or condensed benzene rings in amorphous carbon (Atta-Obeng et al., 2017).
The ratio of disordered or strongly distorted structure of turbostratic carbon to ordered graphite
crystals (Zhao et al., 2013) in the biochars and the activated carbons was estimated by the
intensity ratio of D band to G band (ID/IG). The ID/IG for all the samples are presented in Table 3
alongside the band positions. The ID/IG ratios were between 0.87 and 1.51. Lower ID/IG ratio is an
indication of more aromatic ring structures and less carbon-containing defects that leads to the

T
formation of oxygen-containing functional groups on the surface of the carbon materials (Peng et

IP
al., 2016).

CR
Table 3: Raman Spectra D and G band positions and the corresponding intensity ratios
Sample D Peak (cm-1) G Peak (cm-1) ID/IG

US
PM900 1354.38 1593.67 0.88
KS900 1358.52 1596.00 0.87
AN
KOH-PMBC 1353.99 1597.55 1.51
KOH-KSBC 1365.34 1592.73 0.88
M

KOH-PMBM 1349.04 1592.71 1.10


KOH-KSBM 1381.33 1594.48 1.06
ED

3.4 Surface Functional Groups


PT

Activated carbons have a certain degree of surface chemical heterogeneity, which is due to the
presence of heteroatoms such as oxygen, nitrogen, hydrogen and phosphorus. The presence of
CE

these heteroatoms in relations to the carbon structure are found in various functional
arrangements such as carboxyl, carbonyl, hydroxyl, lactones, quinone and phosphates (Salame et
AC

al., 1999). Knowledge of surface chemistry of carbon material is of fundamental importance as


its behavior in various applications are greatly influenced by the presence of its surface
functional groups. For example, in the use of activated carbon for adsorption purpose, the
presence of certain functional groups can greatly enhance the adsorption of specifically
interacting molecules and these functional groups can also prevent the adsorption of non-
specifically interacting molecules, hindering the molecules from occupying the available
adsorption sites on the activated carbon surface (Salame et al., 1999).

17
ACCEPTED MANUSCRIPT

The X-ray photoelectron spectroscopy (XPS) was used in analyzing the surface functional
groups of the activated carbons. The deconvolved C1s and O1s peaks for the biochars and the
activated carbons are provided in the Supplementary Data. The C1s spectra was deconvolved
into the following peaks based on the chemical shifts showing the aliphatic/aromatic carbon
groups: Peak 1 (284.6 eV) – graphitic carbon C-C/C=C, Peak 2 (285.7 – 286.3 eV) – phenolic,
alcohol, ether or C=N groups, Peak 3 (288.4 – 289.2 eV) – carboxyl or ester groups (Zhou et al.,
2007). The O1s deconvolved peaks were assigned to the following known chemical shifts: Peak I

T
(531.0 – 531.9 eV) – carbonyl oxygen of quinines, Peak II (532.2 – 532.8 eV) – carbonyl oxygen

IP
atoms in esters, anhydrides and oxygen atoms in hydroxyl groups, Peak III (533.1 – 534.6 eV) –

CR
non-carbonyl oxygen single bond in esters and carboxylic acids (Zhou et al., 2007).
Table 4: XPS Carbon and Oxygen Surface Functional Groups of Activated Carbons

US
C1s (% wt.) O1s (% wt.) Atomic conc. (%)
Activated
Carbon Peak
Peak 1 Peak 2 Peak 3 Peak I Peak II C1s O1s Si2p P2p
II
AN
PM900 64.15 - 35.85 73.99 14.53 11.47 85.50 11.57 - 1.57
KS900 60.59 9.52 29.89 68.96 31.04 - 87.52 8.89 1.46 1.05
KOH-PMBC 61.86 8.99 29.15 33.47 61.94 4.59 77.90 16.71 5.39 -
M

KOH-KSBC 71.26 10.61 18.12 28.51 37.95 33.53 66.13 23.34 3.06 -
KOH-PMBM 43.89 25.94 30.17 1.09 77.94 20.97 82.08 13.74 4.18 -
ED

KOH-KSBM 42.89 26.55 30.55 7.51 73.75 18.74 69.62 28.36 2.03 -

Table 4 summarizes the results of the C1s and O1s peaks deconvolution and the atomic
PT

concentration of the detected surface functional groups. The graphitic C (Peak 1) of the activated
carbons were between 43 – 71% wt. with the biomass-derived activated carbons (KOH-PMBM
CE

and KOH-KSBM) having the lowest percentage. This functional group contributes to the
hydrophobicity of the activated carbon and hence adversely affects its wettability and adsorption
AC

characteristics. The hydroxyl and carboxyl groups are polar in nature, which makes the activated
carbon surface hydrophilic and enhances its wettability (Xi et al., 2019). The biomass-derived
activated carbons (KOH-PMBM and KOH-KSBM) had the highest fractions of hydroxyl group
(Peak 2). This can be attributed to the more pronounced effect of KOH on biomass-derived
activated carbon properties. During the KOH activation, the -OK groups formed on the activated
carbon surface are transformed to -OH groups on washing with water by ion exchange reaction,
making the activated carbon surface hydrophilic (Viswanathan et al., 2009). The fraction of the

18
ACCEPTED MANUSCRIPT

carboxyl group (Peak 3) was similar for all the activated carbons except KOH-KSBC which also
showed the highest graphitic carbon fraction.
The deconvolution of the O1s spectra also yielded three peaks and the biochars (PM900 and
KS900) presented the highest relative content of Peak I (carbonyl oxygen). Both the biochar-
derived activated carbons and the biomass-derived activated carbons showed a higher relative
content of Peak II (oxygen atoms in hydroxyl groups) and Peak III (oxygen atoms in carboxyl
group). The reduction in the relative content of Peak I for the activated carbons and an increase

T
in their relative contents of Peak II and III may be attributed to the effect of the activating agent

IP
and subsequent carbonization used for their synthesis.

CR
3.5 Adsorption Kinetics
The adsorption kinetics shows the rate of solute adsorption by measuring the concentration

US
change as a function of time, keeping constant the initial concentration and mass of adsorbent.
Adsorption kinetics helps in determining the time necessary for reaching equilibrium and to
AN
elucidate the mechanism of the adsorption process (Mestre et al., 2009). The pseudo-second
order model is based on the sorption capacity on solid phase and it has been shown to predict
M

adsorption behavior of the adsorbent when the initial concentration of solute is low (Ho, 2006).
The kinetic model parameters estimated for the adsorption of caffeine and acetaminophen using
ED

the activated carbons are presented in Table 5.


The equilibrium uptake (qe) for the activated carbons and biochars were significantly different (p
PT

< 0.05) for both caffeine and acetaminophen. The biomass-derived activated carbons (KOH-
PMBM and KOH-KSBM) and the biochar-derived activated carbon (KOH-PMBC) showed a
CE

higher equilibrium uptake for both adsorbates. This high equilibrium uptakes can be attributed to
their high surface areas and pore volumes. Mestre et al. (2007) reported a similar result for the
AC

adsorption of ibruprofen using two different activated carbons. The activated carbon which has a
higher surface area and mesoporous volume was reported to have a higher equilibrium uptake.
The higher volume of mesopores in the activated carbon was explained to favor quicker initial
adsorption uptake. In this present study, KOH-PMBM has the highest BET surface area of
1596.52 m2/g and mesoporous volume of 0.66 cm3/g, showed the highest equilibrium uptake of
166.67 mg/g and 153.09 mg/g for caffeine and acetaminophen, respectively.

19
ACCEPTED MANUSCRIPT

Table 5: Kinetic model parameters for the adsorption of caffeine and acetaminophen
Activated qe, k x 10-3 t1/2
h (mg/g/min)
carbon (mg/g) (g/mg/min) (min)
PM900 42.45e ± 1.18 0.77c, d ± 0.06 30.94b ± 2.31 1.38d ± 0.11
KS900 54.45d ± 0.17 0.41d ± 0.00 44.47b ± 0.18 1.22d ± 0.01
Caffeine

KOH-PMBC 152.29b ± 1.35 10.80a ± 0.71 0.61c ± 0.04 250.42a ± 12.54


KOH-KSBC 12.83f ± 0.71 1.14c, d ± 0.28 71.61a ± 19.34 0.19d ± 0.06
KOH-PMBM 166.67a ± 0.00 5.35b ± 0.20 1.12c ± 0.04 148.65b ± 5.52
148.52c ± 1.27 1.44c ± 0.01 4.67c ± 0.04 31.78c ± 0.33

T
KOH-KSBM
9.12e ± 0.20 1.08 b ± 0.17 102.57a ± 17.14 0.09d ± 0.02

IP
PM900
Acetaminophen

KS900 30.80d ± 0.33 0.75 b ± 0.14 44.28b, c ± 8.88 0.71d ± 0.12

CR
KOH-PMBC 142.19b ± 1.16 1.56 b ± 0.02 4.51c ± 0.05 31.55b ± 0.35
KOH-KSBC 7.58 e ± 0.87 1.99b ± 1.03 86.90b ± 49.05 0.10d ± 0.04
KOH-PMBM 153.09a ± 2.73 3.98a ± 0.06 1.67c ± 0.06 91.62a ± 4.75

US
KOH-KSBM 125.01c ± 1.56 1.10 b ± 0.01 7.29c ± 0.02 17.16c ± 0.25

*In each column, values with the same letter are not significantly different (p < 0.05)
AN
The pseudo-second order rate constant (k) is an indication of the adsorption efficiency, it
expresses the amount of adsorbent needed for adsorption of 1 mg of the adsorbate in 1 min
M

(Dong et al., 2011). This implies that a smaller value of k is an indication of higher adsorption
ED

efficiency. There was no significant difference (p < 0.05) between the k values of the biochars
(PM900 and KS900) for the adsorption of caffeine and they had the lowest k values compared to
PT

the biochar-derived and biomass-derived activated carbons. This implies that, to adsorb a given
quantity of caffeine, a higher amount of biochar-derived and biomass-derived activated carbons
CE

would be needed. For acetaminophen adsorption, the k value obtained for the biomass-derived
activated carbon (KOH-PMBM) was significantly higher (p <0.05) than the values recorded for
the biochars (PM900 and KS900) and the other activated carbons. This shows that KOH-PMBM
AC

was less efficient in the adsorption of acetaminophen in comparison to the other activated
carbons and biochars.
When considering the rate of the overall adsorption process, the adsorbents that have a lower k
value would be considered to have the best adsorption performance (Mestre et al., 2007).
However, when the initial adsorption rate (h) or the half-life time (t1/2) are considered, the
adsorbent having the higher initial adsorption rate and half-time value will be considered to have
the best adsorption characteristics. The initial adsorption rate (h) values recorded for the biochars

20
ACCEPTED MANUSCRIPT

and the activated carbons during the adsorption of caffeine was found to be higher than the
values recorded for the acetaminophen adsorption. Comparing the biochars to the activated
carbons, the biochars (PM900 and KS900) which were earlier reported to have higher rate
constant (k), were found to have the lowest initial adsorption rate values for both adsorbates.
The half-life time (t1/2) recorded for the biomass-derived activated carbons in adsorbing both
caffeine and acetaminophen were less than 10 mins, which implies that these activated carbons
will adsorb half of the target adsorbates within this time period. The biochar-derived activated

T
carbon (KOH-PMBC) also showed a similar half-life time in comparison to the biomass-derived

IP
activated carbons. However, the biochars showed half-life time that is about two to folds higher

CR
than the ones recorded for the biomass-derived activated carbons. The lower half-life time
reported for the activated carbons may be attributed to their higher mesopore volumes and their

US
hydrophilicity. The activated carbons that showed lower half-life time, had higher mesopore
volumes ranging between 0.12 – 0.66 cm3/g. The mesopore acts as a means of access for the
AN
adsorbate to the micropores. This therefore, enhance the rate of adsorption of the activated
carbons. Also, the activated carbons showing lower half-life time were reported to have a higher
M

proportion of hydroxyl and carboxyl groups on their surfaces, which helps to enhance their
wettability and in turns helps their adsorption behaviors.
ED

3.6 Adsorption Isotherms


The Langmuir and Freundlich model parameters alongside the coefficient of determination (R 2)
PT

of the linear plots are presented in Table 6. A comparison of the R2 values for the two models
showed that the Freundlich isotherm model best fit the adsorption of caffeine unto the biomass-
CE

derived activated carbons (KOH-PMBM and KOH-KSBM) and adsorption of acetaminophen


unto KOH-PMBM. Meanwhile, the Langmuir isotherm model provides the best fit for the
AC

adsorption of caffeine and acetaminophen unto the other activated carbons.


The Langmuir isotherm model assumes the surface of the activated carbon is energetically
homogenous and that a monolayer surface coverage is formed with no interactions between the
molecules adsorbed (Mestre et al., 2009). The Langmuir maximum monolayer adsorption
capacity (qm) of the activated carbons ranged between 11.13 – 198.69 mg/g and 26.36 – 187.52
mg/g for caffeine and acetaminophen, respectively. There was a significant difference (p < 0.05)
in the adsorption capacities of all the adsorbents for the two adsorbates with the biomass-derived
activated carbons showing the highest adsorption capacities for acetaminophen and the biochar-

21
ACCEPTED MANUSCRIPT

derived activated carbon (KOH-PMBC) having the highest adsorption capacity for caffeine
followed by the biomass-derived activated carbons. The high monolayer adsorption capacities
shown by these activated carbons were earlier attributed to both their high surface areas and
hydrophilicity.
The Langmuir constant (b) describes the affinity between the adsorbent and the adsorbate
molecules whereby a higher magnitude depicts a strong interaction between the adsorbent and
adsorbate molecules. There was no significant difference (p < 0.05) in the Langmuir constant

T
recorded for all the adsorbents in the adsorption of caffeine. This implies that all the adsorbents

IP
showed a similar affinity to the caffeine molecules during adsorption. For acetaminophen

CR
adsorption, there was no significant difference (p < 0.05) in the Langmuir constants for both the
biochar-derived and biomass-derived activated carbons. However, there was a significant

US
difference in the Langmuir constants for the biochars with KS900 having the highest value and
PM900 having the lowest value.
AN
The Freundlich isotherm model is an empirical model that describes multilayer adsorption with
non-uniform distribution of adsorption heat and affinities on a heterogenous surface having
M

different energy sites (Foo & Hameed, 2010). From Table 6, the surface heterogeneity (1/n)
values were 1.84 and 0.67 for KOH-PMBM and KOH-KSBM respectively for the adsorption of
ED

caffeine. For the acetaminophen adsorption, the surface heterogeneity value was 0.33 for KOH-
PMBM. When the surface heterogeneity (1/n) value is below unity, it is an indication of
PT

chemisorption process and when the value is above unity, it is indicative of cooperative
adsorption (Foo & Hameed, 2010). This means that the adsorption of caffeine on KOH-KSBM
CE

and acetaminophen on KOH-PMBM involved chemical reaction of the adsorbates with the
surface functional groups on the activated carbon surface. However, for the adsorption of
AC

caffeine on KOH-PMBM where 1/n is above unity, is an indication of cooperative adsorption.

22
ACCEPTED MANUSCRIPT

Table 6: Langmuir and Freundlich isotherm parameters for the adsorption of caffeine
Activated Langmuir Equation Freundlich Equation
2
Carbon qm (mg/g) b (l/mg) R 1/n KF R2
KS900 38.04d ± 0.87 4.71a ± 3.44 0.99 0.06 ± 0.01 31.28c ± 0.45 0.55
PM900 38.42d ± 0.52 4.41a ± 2.60 0.99 0.12 ± 0.00 26.79d ± 0.21 0.34
Caffeine

a a b
KOH-PMBC 198.69 ± 2.26 0.62 ± 0.02 0.93 0.53 ± 0.00 72.21 ± 0.25 0.75
11.13e ± 0.55 3.06a ± 2.84 8.19f ± 0.66

T
KOH-KSBC 0.89 0.11 ± 0.03 0.13
c a a

IP
KOH-PMBM 99.60 ± 9.31 0.67 ± 0.03 0.78 1.84 ± 0.06 181.23 ± 1.90 0.95
KOH-KSBM 173.42b ± 1.75 0.06a ± 0.00 0.98 0.67 ± 0.00 13.79e ± 0.02 0.99

CR
b, c
85.56 ±
KS900 0.76a ± 0.39 0.95 0.20c ± 0.05 45.35b ± 5.14 0.24
Acetaminophen

6.53
PM900 44.51d ± 0.30 0.24b ± 0.00 0.73 0.27b ± 0.00 16.38d ± 0.12 0.64

US
KOH-PMBC 76.55 c ± 1.37 0.58a, b ± 0.04 0.98 0.30b ± 0.01 31.08c ± 0.21 0.71
e a, b d d
KOH-KSBC 26.36 ± 0.29 0.51 ± 0.04 0.92 0.13 ± 0.01 16.12 ± 0.22 0.01
AN
KOH-PMBM 91.05b ± 4.49 0.64a, b ± 0.12 0.95 0.33b ± 0.01 34.51c ± 0.25 0.96
a a, b a a
KOH-KSBM 187.52 ± 2.02 0.53 ± 0.01 0.91 0.54 ± 0.00 62.52 ± 0.39 0.77
* In each column, values with the same letter are not significantly different (p < 0.05)
M

qm – maximum monolayer adsorption capacity; b - Langmuir constant


n – Freundlich exponent; KF – Freundlich constant; R2 – linear regression coefficient of determination
Ce – solution concentration at equilibrium (mg/l); qe - uptake at equilibrium (mg/g)
ED

The dimensionless constant separation factor (RL), describes the favorable or irreversible
PT

adsorption nature of the adsorbates unto the activated carbon. The factor is described in the
equation 6;
CE

where Co (ppm) is the initial concentration of the adsorbate and b is the Langmuir constant. The
AC

RL value classifies adsorption as irreversible (RL= 0), favorable (0 < RL< 1), linear (RL= 1) and
unfavorable (RL >1). Figure 4 shows the separation factors for both acetaminophen and caffeine
adsorbed by the activated carbons. From the plots, it can be observed that the factor reduces with
increase in the initial concentration of the adsorbates. All the R L values obtained in this study
were 0 < RL < 1, which implies that nature of the adsorption was favorable. Also, according to
Sharma & Forster (1993), the low RL value shows that the interaction between the adsorbate
molecules and the activated carbons was relatively strong.

23
ACCEPTED MANUSCRIPT

T
IP
CR
US
Figure 4: Plot of separation factor (RL) against initial concentration of acetaminophen and
caffeine using the biochars and the activated carbons
AN
4. Conclusion
The study showed that the surface areas of the biomass-derived activated carbons were higher
M

than the biochar-derived activated carbons. There was no improvement in the porous structure of
the biochar-derived activated carbons as shown in the pore size distribution. The effect of the
ED

biomass precursor impregnation was evident in the microstructure of the biomass-derived


activated carbons, such that the SEM images did not show any distinctive porous structure. Also,
PT

the biomass-derived activated carbons presented an overall better adsorption characteristic for
both caffeine and acetaminophen. This showed that the intermediate production of biochar as a
CE

precursor before impregnation and activation may not be necessary.


AC

Acknowledgement
The authors wish to acknowledge the financial supports from USDA National Institute of Food
and Agriculture, McIntire Stennis project (accession number 1007044) and Hatch Project
(accession number 1019107) towards this research work.

Conflict of Interest
The authors declare no conflict of interest including financial, personal or other relationships
with anybody or organizations

24
ACCEPTED MANUSCRIPT

References
Álvarez-Torrellas, S., Peres, J.A., Gil-Álvarez, V., Ovejero, G., García, J. 2017. Effective
adsorption of non-biodegradable pharmaceuticals from hospital wastewater with different
carbon materials. Chemical Engineering Journal, 320, 319-329.
Atta-Obeng, E., Dawson-Andoh, B., Seehra, M.S., Geddam, U., Poston, J., Leisen, J. 2017.
Physico-chemical characterization of carbons produced from technical lignin by sub-
critical hydrothermal carbonization. Biomass and Bioenergy, 107, 172-181.
Bansal, R.C., Goyal, M. 2005. Activated Carbon Adsoprtion CRC Press, Boca Raton, FL, USA.
Bazan-Wozniak, A., Nowicki, P., Pietrzak, R. 2017. The influence of activation procedure on the

T
physicochemical and sorption properties of activated carbons prepared from pistachio
nutshells for removal of NO2/H2S gases and dyes. Journal of Cleaner Production, 152,

IP
211-222.
Dittmar, S., Zietzschmann, F., Mai, M., Worch, E., Jekel, M., Ruhl, A.S. 2018. Simulating

CR
Effluent Organic Matter Competition in Micropollutant Adsorption onto Activated
Carbon Using a Surrogate Competitor. Environmental Science & Technology, 52(14),
7859-7866.

US
Djilani, C., Zaghdoudi, R., Djazi, F., Bouchekima, B., Lallam, A., Modarressi, A., Rogalski, M.
2015. Adsorption of dyes on activated carbon prepared from apricot stones and
commercial activated carbon. Journal of the Taiwan Institute of Chemical Engineers, 53,
AN
112-121.
Dong, Z.B., Liang, Y.R., Fan, F.Y., Ye, J.H., Zheng, X.Q., Lu, J.L. 2011. Adsorption Behavior
of the Catechins and Caffeine onto Polyvinylpolypyrrolidone. Journal of Agricultural
M

and Food Chemistry, 59(8), 4238-4247.


Donn, J., Mendoza, M., Pritchard, J. 2008. Pharmaceuticals found in drinking water, affecting
wildlife and maybe humans. in: Pharmawater, Associated Press.
ED

Ebele, A.J., Abou-Elwafa Abdallah, M., Harrad, S. 2017. Pharmaceuticals and personal care
products (PPCPs) in the freshwater aquatic environment. Emerging Contaminants, 3(1),
1-16.
PT

Elmouwahidi, A., Bailón-García, E., Pérez-Cadenas, A.F., Maldonado-Hódar, F.J., Carrasco-


Marín, F. 2017. Activated carbons from KOH and H3PO4-activation of olive residues
and its application as supercapacitor electrodes. Electrochimica Acta, 229, 219-228.
CE

Foo, K.Y., Hameed, B.H. 2010. Insights into the modeling of adsorption isotherm systems.
Chemical Engineering Journal, 156(1), 2-10.
Fuertes, A.B., Ferrero, G.A., Diez, N., Sevilla, M. 2018. A Green Route to High-Surface Area
AC

Carbons by Chemical Activation of Biomass-Based Products with Sodium Thiosulfate.


ACS Sustainable Chemistry & Engineering, 6(12), 16323-16331.
Guizani, C., Haddad, K., Limousy, L., Jeguirim, M. 2017. New insights on the structural
evolution of biomass char upon pyrolysis as revealed by the Raman spectroscopy and
elemental analysis. Carbon, 119, 519-521.
Ho, Y.S. 2006. Review of second-order models for adsorption systems. Journal of Hazardous
Materials, 136(3), 681-689.
Hui, T.S., Zaini, M.A.A. 2015. Potassium hydroxide activation of activated carbon: a
commentary. Carbon Letters, 16(4), 275-280.
Juejun, K., Supunnee, J., Chaiyot, T. 2013. Effect of carbonization temperature on properties of
char and activated carbon from coconut shell. Qingdao Daxue Shifanxueyuan
Xuebao/Journal of Teachers College Qingdao University, 20(4), 269-278.

25
ACCEPTED MANUSCRIPT

Kupgan, G., Liyana-Arachchi, T.P., Colina, C.M. 2017. NLDFT Pore Size Distribution in
Amorphous Microporous Materials. Langmuir, 33(42), 11138-11145.
Lam, S.S., Su, M.H., Nam, W.L., Thoo, D.S., Ng, C.M., Liew, R.K., Yuh Yek, P.N., Ma, N.L.,
Nguyen Vo, D.V. 2019. Microwave Pyrolysis with Steam Activation in Producing
Activated Carbon for Removal of Herbicides in Agricultural Surface Water. Industrial &
Engineering Chemistry Research, 58(2), 695-703.
Liu, L., Jin, S., Park, Y., Park, Y.C., Lee, C.-H. 2018. Sorption Equilibria and Kinetics of CO2,
N2, and H2O on KOH-Treated Activated Carbon. Industrial & Engineering Chemistry
Research, 57(50), 17218-17225.
Lladó, J., Lao-Luque, C., Ruiz, B., Fuente, E., Solé-Sardans, M., Dorado, A.D. 2015. Role of

T
activated carbon properties in atrazine and paracetamol adsorption equilibrium and

IP
kinetics. Process Safety and Environmental Protection, 95, 51-59.
Marques, S.C.R., Mestre, A.S., Machuqueiro, M., Gotvajn, A.Ž., Marinšek, M., Carvalho, A.P.

CR
2018. Apple tree branches derived activated carbons for the removal of β-blocker
atenolol. Chemical Engineering Journal, 345, 669-678.
Mazlan, M.A.F., Uemura, Y., Yusup, S., Elhassan, F., Uddin, A., Hiwada, A., Demiya, M. 2016.
Activated Carbon from Rubber Wood Sawdust by Carbon Dioxide Activation. Procedia

US
Engineering, 148, 530-537.
McNaught, A.D., Wilkinson, A. 1997. IUPAC. Compendium of Chemical Terminology.
Blackwell Scientific Publications, Oxford.
AN
Mestre, A.S., Freire, C., Pires, J., Carvalho, A.P., Pinto, M.L. 2014. High performance
microspherical activated carbons for methane storage and landfill gas or biogas upgrade.
Journal of Materials Chemistry A, 2(37), 15337-15344.
M

Mestre, A.S., Pires, J., Nogueira, J.M.F., Carvalho, A.P. 2007. Activated carbons for the
adsorption of ibuprofen. Carbon, 45(10), 1979-1988.
ED

Mestre, A.S., Pires, J., Nogueira, J.M.F., Parra, J.B., Carvalho, A.P., Ania, C.O. 2009. Waste-
derived activated carbons for removal of ibuprofen from solution: Role of surface
chemistry and pore structure. Bioresource Technology, 100(5), 1720-1726.
Nimmo, J.R. 2005. Porosity and Pore-Size Distribution. in: Encyclopedia of Soils in the
PT

Environment, (Ed.) D. Hillel, Elsevier. Oxford, pp. 295-303.


Oginni, O. 2018. Characteristics of Activated Carbons Produced from Herbaceous Biomass
Feedstock. in: Wood Science and Technology, Vol. Ph.D., West Virginia University.
CE

Morgantown, WV.
Oginni, O., Singh, K. 2019. Pyrolysis characteristics of Arundo donax harvested from a
reclaimed mine land. Industrial Crops and Products, 133, 44-53.
AC

Oginni, O., Singh, K., Zondlo, J.W. 2017. Pyrolysis of dedicated bioenergy crops grown on
reclaimed mine land in West Virginia. Journal of Analytical and Applied Pyrolysis, 123,
319-329.
Oh, G.H., Yun, C.H., Park, C.R. 2003. Role of KOH in the one-stage KOH activation of
cellulosic biomass. Carbon Science, 4(4), 180 – 184
Pelekani, C., Snoeyink, V.L. 1999. Competitive adsorption in natural water: Role of activated
carbon pore size. Water Research, 33(5), 1209-1219.
Peng, B., Chen, L., Que, C., Yang, K., Deng, F., Deng, X., Shi, G., Xu, G., Wu, M. 2016.
Adsorption of Antibiotics on Graphene and Biochar in Aqueous Solutions Induced by π-π
Interactions. Scientific Reports, 6, 31920.

26
ACCEPTED MANUSCRIPT

Qiu, Z., Wang, Y., Bi, X., Zhou, T., Zhou, J., Zhao, J., Miao, Z., Yi, W., Fu, P., Zhuo, S. 2018.
Biochar-based carbons with hierarchical micro-meso-macro porosity for high rate and
long cycle life supercapacitors. Journal of Power Sources, 376, 82-90.
Salame, I.I., Bagreev, A., Bandosz, T.J. 1999. Revisiting the Effect of Surface Chemistry on
Adsorption of Water on Activated Carbons. The Journal of Physical Chemistry B,
103(19), 3877-3884.
Sharma, D.C., Forster, C.F. 1993. Removal of hexavalent chromium using sphagnum moss peat.
Water Research 27(7), 1201-1208.
Sing, K.S.W., Everett, D.H., Haul, R.A.W., Moscou, L., Pierotti, R.A., Rouquerol, J.,
Siemieniewska, T. 1985. Reporting physisorption data for gas/solid systems with special

T
reference to the determination of surface area and porosity. Pure and Applied Chemistry,

IP
57(4), 603 - 619.
Sotelo, J.L., Ovejero, G., Rodríguez, A., Álvarez, S., Galán, J., García, J. 2014. Competitive

CR
adsorption studies of caffeine and diclofenac aqueous solutions by activated carbon.
Chemical Engineering Journal, 240, 443-453.
Sun, Y., Cheng, J. 2002. Hydrolysis of lignocellulosic materials for ethanol production: a review.
Bioresource Technology, 83(1), 1-11.

US
Umran, T.U., Ates, F., Erginel, N., Ozcan, O., Oduncu, E. 2015. Adsorption of Disperse Orange
30 dye onto activated carbon derived from Holm Oak (Quercus ilex) acorns: A 3k
factorial design and analysis. Journal of Environmental Management, 155, 89-96.
AN
Villota, S.M., Lei, H., Villota, E., Qian, M., Lavarias, J., Taylan, V., Agulto, I., Mateo, W.,
Valentin, M., Denson, M. 2019. Microwave-Assisted Activation of Waste Cocoa Pod
Husk by H3PO4 and KOH—Comparative Insight into Textural Properties and Pore
M

Development. ACS Omega, 4(4), 7088-7095.


Viswanathan, B., Indra Neel, P., Varadarajan, T.K. 2009. Methods of Activation and Specific
ED

Applications of Carbon Materials. National Centre for Catalysis Research, Indian


Institute of Technology Madras, Chennai, India.
Xi, X., Jiang, S., Zhang, W., Wang, K., Shao, H., Wu, Z. 2019. An experimental study on the
effect of ionic liquids on the structure and wetting characteristics of coal. Fuel, 244, 176-
PT

183.
Yumak, T., Bragg, D., Sabolsky, E.M. 2018. Effect of synthesis methods on the surface and
electrochemical characteristics of metal oxide/activated carbon composites for
CE

supercapacitor applications. Applied Surface Science, 469, 983 – 993


Zhang, J., Gao, J., Chen, Y., Hao, X., Jin, X. 2017. Characterization, preparation, and reaction
mechanism of hemp stem based activated carbon. Results in Physics, 7, 1628-1633.
AC

Zhao, L., Cao, X., Mašek, O., Zimmerman, A. 2013. Heterogeneity of biochar properties as a
function of feedstock sources and production temperatures. Journal of Hazardous
Materials, 256–257, 1-9.
Zhou, J.H., Sui, Z.J., Zhu, J., Li, P., Chen, D., Dai, Y.C., Yuan, W.K. 2007. Characterization of
surface oxygen complexes on carbon nanofibers by TPD, XPS and FT-IR. Carbon, 45(4),
785-796.

27

You might also like