Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

sustainability

Article
Prediction for the Adsorption of Low-Concentration Toluene by
Activated Carbon
Ying Sheng *, Qingqing Dong, Qiang Ren and Mingyang Wang

Tianjin Key Lab of Indoor Air Environmental Quality Control, School of Environmental Science and Engineering,
Tianjin University, Tianjin 300350, China
* Correspondence: ysheng@tju.edu.cn

Abstract: Activated carbon filters are widely used to remove gaseous pollutants in order to guarantee
a healthy living environment. The standard method for evaluating the adsorption performance of
filters is conducted at ~100 ppm. Although this accelerates the test and avoids the high requirements
of the test device, it is still far from the contaminant concentration in the indoor environment, and
adsorbents in practical application may show different capabilities. Therefore, this study compared
several methods for predicting the adsorption performance of activated carbon and recommended a
procedure based on the Wheeler–Jonas model to estimate the breakthrough curve at low concentra-
tions using experimental data at high concentrations. The results showed that the Langmuir model
and Wood–Lodewyckx correlation were the most suitable for obtaining the equilibrium adsorption
capacity and mass transfer coefficient, which are critical parameters in the Wheeler–Jonas model.
The predicted service life was derived from the breakthrough curve. A modification method based
on a relationship with inlet gas concentration was proposed to reduce the prediction deviation of
the service life. After modification, the maximum deviation was within two hours and the relative
deviation was no more than 7%.

Keywords: adsorption; toluene; activated carbon; prediction method; model

1. Introduction
Citation: Sheng, Y.; Dong, Q.; Ren, Q.;
Volatile organic compounds (VOCs) are primary pollutants contributing to the deteri-
Wang, M. Prediction for the
oration of indoor air quality [1]. According to the chemical structure, VOCs can be divided
Adsorption of Low-Concentration
into alkanes, aromatic hydrocarbons, alkenes, halogen hydrocarbons, esters, aldehydes, etc.
Toluene by Activated Carbon.
They are mainly released from decoration materials, various furniture in buildings, and
Sustainability 2023, 15, 1555. https://
doi.org/10.3390/su15021555
human activities, and are transmitted from the outdoors. Long-time exposure to formalde-
hyde and highly hazardous VOCs, such as benzene and toluene, harms human health,
Academic Editor: Ning Yuan inducing various diseases such as Sick Building Syndrome (SBS), respiratory diseases,
Received: 22 November 2022
cardiovascular diseases, and even leukemia and cancer [2]. It is reported that indoor VOC
Revised: 2 January 2023 pollution is severe in China. Zhou et al., tested 436 air samples in residential buildings
Accepted: 10 January 2023 from 6 cities, indicating that the over-standard rates of formaldehyde and TVOC were 70%
Published: 13 January 2023 and 50%, respectively. The concentrations of benzene, toluene, and xylene were also over
the limits by varying degrees [3]. Therefore, removing VOCs is essential for improving
indoor air quality and protecting the occupants’ health.
Adsorption is considered one of the most promising VOC removal strategies due to
Copyright: © 2023 by the authors. its low cost, easy operation, and lack of by-products [4,5]. Among numerous adsorbents,
Licensee MDPI, Basel, Switzerland. granular activated carbon (GAC) is the most popular because of its large specific surface
This article is an open access article area, well-developed pore structure, strong adsorption capacity, stable chemical properties,
distributed under the terms and
and reusability [6,7]. The concentration of indoor VOCs is only at the ppb~ppm level.
conditions of the Creative Commons
To acquire the adsorption performance of activated carbon at low concentrations, some
Attribution (CC BY) license (https://
researchers attempted to test the gas-phase filters at the ppb~ppm level [8–10]. However,
creativecommons.org/licenses/by/
the test duration needed to be shorter, yet was sometimes as long as hundreds of hours. It
4.0/).

Sustainability 2023, 15, 1555. https://doi.org/10.3390/su15021555 https://www.mdpi.com/journal/sustainability


Sustainability 2023, 15, 1555 2 of 16

also had high demands for experimental instruments. Therefore, it is infeasible to test the
gas-phase filters with low-concentration pollutants considering the test duration and eco-
nomic cost. To limit the measurement effort, ISO 10121-2 was set to assess the performance
of gas-phase air cleaning media, in which toluene was recommended as a representative
substance for VOCs at concentrations of 9 or 90 ppm [11]. In another standard of ASHRAE
145.1, the concentration of toluene in the accelerated test was 100 ppm [12]. However, the
concentration in the standard methods is still significantly higher than that in the actual
buildings, and adsorbents may show different capabilities in practical application, so the
standard test methods did not seem suitable for evaluating the long-term performance of
the air cleaning media under realistic conditions.
In contrast, there is a quick and effective method to build a model that can predict
low-concentration adsorption of VOCs by fitting the measurement data at higher concen-
trations. The differential equations based on the adsorption kinetics mechanism describe
the breakthrough adsorption capacity of the gas-phase filters. Vizhemehr et al. [13] studied
three typical models based on the absorption process in activated carbon, namely the homo-
geneous surface diffusion model (HSDM), the pore diffusion model (PDM), and the pore
surface diffusion model (PSDM). The finite difference method was used to discretize the
space of the adsorption bed, and the adsorption process was regarded as a series of nodes.
The concentration of pollutants was obtained by time integration using the Matlab software.
Many parameters involved in the models were estimated from empirical or semi-empirical
formulas, which led to the prediction from the models being inconsistent with the experi-
mental data at higher concentrations of 15~150 ppm but disagreeing with those at a lower
concentration. From the perspective of service life, the deviation of predicted and measured
values at 15 ppm was over 5 h. Subsequently, Shaverdi et al. [14] used the differential
equation in place of the mass balance equation to simplify the calculation process of the
iterative algorithm in the Matlab software. They made the predicted results closer to the
adsorption data of methyl ethyl ketone, at 15~100 ppm, and n-hexane, at 18~100 ppm.
In the study of Ligoski et al. [15], the breakthrough curves of toluene at concentrations
of 0.09 or 0.9 ppm were predicted using the experimental data at 9~90 ppm. However,
the predicted curves were only in good agreement with the S-shaped breakthrough curve.
The experiment at 0.09 ppm still needed to be completed, and the accuracy of the service
life still needed to be determined. Yao et al. [16] summarized the dynamic models for
predicting the breakthrough curves, including the Yoon–Nelson model, Adams–Bohart
model, Wolborska model, Thomas model, Boltzmann distribution function, etc. It was
concluded that the Yoon–Nelson model was widely used in various adsorption systems
because the input parameters were easily obtained. The breakthrough curve predicted
by Shiue et al. [17], based on the D–R equation and modified Yoon–Nelson model, could
estimate the service life of chemical filters at different VOC concentrations (10~70 ppm)
and different filtration speeds (0.076, 0.114, and 0.152 m/s). Nevertheless, since the critical
parameters for the modeling were determined under the low face velocities deviated from
the realistic operating conditions, the feasibility of this method needed to be verified by the
experimental data when the face velocity was above 1 m/s.
As previously mentioned, a variety of prediction models were proposed. Due to the
complicated calculation process [13,14] for differential equations, it is more efficient to
predict chemical filters’ adsorption performance and service life using a breakthrough
model [15,17]. With the problems of the limited range of applications [15] and insufficient
experimental verification [17], it is necessary to verify the prediction method further to
accurately evaluate the adsorption performance and the service life of the gas-phase air
cleaning medium in a realistic operation environment. In this paper, a test on activated car-
bon for toluene adsorption at low concentrations was conducted. Then, we compared four
prediction methods based on the Wheeler–Jonas model to predict the toluene adsorption in
the activated carbon at the ppm level. To accomplish this, we used the experimental data
from relatively high concentrations and further optimized the prediction method to obtain
the accurate service life of the adsorbent.
data from relatively high concentrations and further optimized the prediction method to
data from
obtain the relatively high concentrations
accurate service and further optimized the prediction method to
life of the adsorbent.
obtain the accurate service life of the adsorbent.
Sustainability 2023, 15, 1555 2. Experimental Section 3 of 16
2. Experimental Section
2.1. Characteristics of Activated Carbon
2.1. Characteristics of Activated Carbon
Activated Section
2. Experimental carbon made from coconut shells was chosen for the experiments and
named Activated
granular carbon
activated made from(GAC). coconutTheshells was chosen for the experiments and
2.1. Characteristics of Activatedcarbon
Carbon structural parameters through BET charac
named granular
terization arecarbon activated carbon
summarized in Table (GAC).TheThe
1.shells structural
adsorption andparameters through BET charac
Activated
terization are made from
summarized in coconut
Table 1. The was chosen for
adsorption thedesorption
and experimentscurves
desorption and named
curves
of activated
of activated
carbon inactivated
granular nitrogen, as well
carbon as the
(GAC). Thepore distribution,
structural are shown
parameters throughin Figures
BET 1 and 2. The BET
characterization
carbon
specific in nitrogen,
surface inarea as well
of 1.
GAC as the pore distribution, are shown in Figures 1by
and 2. The BET
are summarized Table The was 1143.9 and
adsorption m /g.
2 The GAC
desorption was dominated
curves mesoporosity
of activated carbon in
specific
nitrogen, surface
with a proportion area
as well as the of GAC
of 65.2%, was 1143.9
and its average
pore distribution, m 2/g. The GAC was dominated by mesoporosity
poreinsize
are shown was 11.79
Figures andnm. TheBET-specific
2. The total pore volume
with
surface
of GAC a proportion
area GAC of
wasof0.1020 65.2%,
was
cm 31143.9
/g. Toand
m its average
2 /g. The GAC
eliminate pore size was
was dominated
the influence of the1.79
by nm. Thegranular
totalwith
mesoporosity
size of the pore volume
aactivated
of GAC
proportion was
carbon onofthe 0.1020
65.2%, cm 3/g. To eliminate the influence of the size of the granular activated
and its average
adsorption, pore size carbon
the activated was 1.79sample
nm. Thewastotaluniformly
pore volume of GACto 20~30
ground
carbon
was on
0.1020 cm 3 /g.
the adsorption,
To eliminate thetheactivated
influence carbon
of the sample
size of the was uniformly
granular ground
activated carbonto 20~30
meshes (0.6~0.9 mm).
on the adsorption,
meshes (0.6~0.9 mm).the activated carbon sample was uniformly ground to 20~30 meshes
(0.6~0.9 mm).
Table 1. Structural parameters of activated carbon sample.
Table 1. Structural parameters of activated carbon sample.
Table 1. Structural parameters of activated carbon sample.
Average Total
Pore Microporous Mesoporous
BET-Specific Average Total Pore Microporous Mesoporous
Sample BET-Specific
BET-Specific 2 Pore Size TotalVolume VolumeMesoporous
Volume
Sample Surface Area (m /g) Average
Pore
Pore
Size Volume
Volume
Microporous
Volume Volume
SampleSurface
Surface
AreaArea
(m /g)
2 Pore Size
(nm) (cm /g)
3 Volume (cm /g)
3 Volume (cm3/g)
2
(m /g) (nm)(nm) (cm (cm
3 /g) /g)
3 3 (cm /g) (cm /g) (cm3/g)
(cm /g) 3 3
GAC 1143.9 1.79 0.1020 0.0356 0.0665
GAC
GAC 1143.9
1143.9 1.791.79 0.1020
0.1020 0.03560.0356 0.0665 0.0665

Figure1.1.The
Figure Theadsorption
adsorption and
and desorption
desorption curves
curves of activated
of activated carboncarbon in nitrogen.
in nitrogen.
Figure 1. The adsorption and desorption curves of activated carbon in nitrogen.

Figure2.2.The
Figure Thepore
poredistribution of the
distribution activated
of the carbon.
activated carbon.
Figure 2. The pore distribution of the activated carbon.
Sustainability 2023, 15, x FOR PEER REVIEW 4 of 17

The breakthrough curves of toluene with different inlet concentrations were meas-
Sustainability 2023, 15, 1555 ured in the experiments. The prediction methods were compared based on the4 ofmeasure-
16
ment data of GAC, and then the accuracy and feasibility of the methods were analyzed.

2.2. Experimental Facilitycurves of toluene with different inlet concentrations were measured
The breakthrough
in the experiments. The prediction methods were compared based on the measurement
The experimental system is shown in Figure 3 with the functions of challenge gas
data of GAC, and then the accuracy and feasibility of the methods were analyzed.
generation and adsorption in the activated carbon. The pressurized air successively
passed through aFacility
2.2. Experimental silica gel drying column and a high-efficiency particulate air (HEPA)
filter toThebecome dry and
experimental cleaniscarrier
system showngas. The toluene
in Figure 3 with the with a high of
functions concentration
challenge gasof 100
ppmgeneration and adsorption in the activated carbon. The pressurized air successively passed The
in the cylinder was mixed with carrier gas to reach the desired concentration.
flow rate of
through cleangel
a silica gasdrying
was controlled
column and by amass flow controller
high-efficiency 1, with
particulate aira (HEPA)
range offilter
0~10toL/min,
andbecome dry and
the flow rateclean carrier was
of toluene gas. The toluene by
controlled with a high
mass concentration
flow controller of 2, 100
with ppm in the of 0~2
a range
L/min. The accuracy of each controller was 1% of the full range. The flow rate ofofthe ad-
cylinder was mixed with carrier gas to reach the desired concentration. The flow rate
clean gas was controlled by mass flow controller 1, with a range of 0~10 L/min, and the
sorption column was 2.5 L/min, controlled by a rotameter, and the excess gas was by-
flow rate of toluene was controlled by mass flow controller 2, with a range of 0~2 L/min.
passed and discharged into the atmosphere after treatment. Before the experiment, the
The accuracy of each controller was 1% of the full range. The flow rate of the adsorption
concentration
column was 2.5 of L/min,
toluenecontrolled
was tested by atorotameter,
ensure a and constant concentration
the excess gas was bypassedupstream and of the
adsorption
dischargedcolumn
into the by using a blank
atmosphere column. Before
after treatment. The fluctuation was within
the experiment, 5% of theof
the concentration desired
concentration.
toluene was testedThe toactivated
ensure a carbon
constantsample was fixed
concentration by metal
upstream of the mesh in the
adsorption adsorption
column
column.
by usingThe adsorption
a blank column.column had anwas
The fluctuation inner diameter
within of 10
5% of the mm and
desired a filling height
concentration. The of 3
activated carbon sample was fixed by metal mesh in the adsorption
mm. The ppbRAE 3000 device was used to continuously monitor the concentration of tol- column. The adsorption
column
uene had an inner
downstream diameter
of the of 10 mm
adsorption and a and
column, filling height
the dataof 3 mm.
were The ppbRAE
recorded 3000
at an interval of
device was used to continuously monitor the concentration of toluene
1 min. The ppbRAE 3000 is a device to measure the concentration of organic gaseous pol- downstream of the
adsorption column, and the data were recorded at an interval of 1 min. The ppbRAE 3000 is
lutants, with a wide detection range from 1 ppb to 10,000 ppm. It uses a third-generation
a device to measure the concentration of organic gaseous pollutants, with a wide detection
photoionization
range from 1 ppb detector
to 10,000(PID)
ppm.to It improve detection accuracy
uses a third-generation and reach
photoionization a resolution
detector (PID) of 1
ppb.
to improve detection accuracy and reach a resolution of 1 ppb.

Figure
Figure Schematics of
3.3.Schematics of the experimental
experimentalsystem.
system.

Chen et al. [18] tested the newly renovated houses and found that the TVOC reached
9.22 mg/m3 (2.41 ppm). During the decoration period, the highest TVOC concentration in
Sustainability 2023, 15, 1555 5 of 16

Chen et al. [18] tested the newly renovated houses and found that the TVOC reached
9.22 mg/m3 (2.41 ppm). During the decoration period, the highest TVOC concentration in
the office building reached 11 mg/m3 (2.87 ppm) [19]. Then, for indoor TVOC pollution,
we selected 6 desired concentrations of 0.5, 1.0, 2.0, 3.0, 4.0 and 5.0 ppm.

3. Prediction Method
With the simple formulas and easily available parameters, the Wheeler–Jonas model
was popular for predicting the breakthrough curve. The Wheeler–Jonas model was initially
derived from the mass balance of gases entering and passing through the adsorption bed,
and was then modified [20]. The model is expressed as follows:

qe mGAC qe ρb Cb
tb = + ln( ) (1)
Ci Q Kv Ci Ci − Cb

where tb is the breakthrough time, h; Kv is the mass transfer coefficient, 1/h; qe is the
equilibrium adsorption capacity at the concentration of Ci , mg/g; Cb is the breakthrough
concentration of adsorbate, ppm; Ci is the initial concentration of adsorbate, ppm; ρb is the
bulk density of activated carbon, g/m3 ; Q is the carrier gas flow rate through the adsorption
column, m3 /min.
The advantage of the Wheeler–Jonas model was considering the characteristics of
fixed bed and operation conditions. It was easily noted that it was essential to acquire the
equilibrium adsorption capacity and the mass transfer coefficient for the prediction. Then,
the appropriate methods for obtaining the two critical parameters were presented.

3.1. Isotherms
3.1.1. Langmuir Model
The Langmuir model was the theoretical model for monolayer adsorption, and it was
widely used to describe the VOCs adsorption in the adsorbents [21]. It is expressed as:

qe Kl Ci
= (2)
qm 1 + Kl Ci

where Ci is the inlet concentration, mg/m3 ; qe is the equilibrium adsorption capacity at the
concentration of Ci , mg/g; qm is the maximum adsorption capacity in theory, mg/g; and Kl
is the equilibrium adsorption constant, m3 /mg, which is related to the nature of adsorbate.
The equilibrium adsorption capacity qe can also be calculated by Equation (3) using
numerical integration:
R tf
Q(Ci t f − 0 Cb dt)
qe = (3)
mGAC

where Q is the flow rate of carrier gas through the adsorption column, m3 /min; Cb is the
breakthrough concentration, mg/m3 ; t f is the time at 95% breakthrough fraction, min;
mGAC is the quality of absorbent, g.
Sustainability 2023, 15, 1555 6 of 16

3.1.2. Freundlich Model


The Freundlich model is an empirical model based on experimental results without
theoretical derivation [22]. The expression of the Freundlich model is:

qe = K f Ci1/n (4)

where K f is the equilibrium adsorption constant describing how strong the absorbate is
attached to the surface of the adsorbent. 1/n represents the difficulty in the adsorption
process. The value of 1/n is between 0 and 1 in most cases. A smaller value of 1/n indicates
that the adsorption occurs more easily. Once the value reaches more than 2, it indicates that
the it is challenging for adsorption to occur.

3.1.3. Dubinin–Radushkevich (D–R) Model


The D–R model is a semi-empirical equation based on adsorption potential theory and
micropore filling theory [23]. The expression of the D–R model is:
(  )
1 2 p0 2

We = W0 exp −( ) RT ln( ) (5)
E p

where We is the pore volume occupied by the adsorbate, cm3 /g; W0 is the limiting pore
volume of adsorption, cm3 /g; E is the adsorption free energy of the adsorbate, J/mol;
RT ln( p0 /p) is the adsorption potential; R is the gas constant, 8.314 J/mol·K; T is absolute
temperature, K; p0 /p is the relative pressure of the adsorbate; p0 is the saturation pressure
of the adsorbate at the adsorption temperature, Pa; and p is the pressure of the adsorbate
vapor, Pa.
Since the pore volume occupied by the adsorbate is challenging to obtain, the D–R model
isotherm can be rewritten by the expression of adsorption capacity as Equation (5) [24]:
(  )
1 2

1
qe = qm exp − 2 RT ln(1 + ) (6)
2E Ci

The calculation of equilibrium adsorption capacity is generally performed with the


Langmuir, Freundlich, and D–R models. To determine the applicability of the three
isotherms for different concentration ranges, the application of isotherms for toluene
adsorption in the previous studies is shown in Figure 4. The results showed that three
isotherms could fit toluene adsorption without a strict adsorbate concentration requirement.
Particularly, the Langmuir model was widely used at the ppm level [25,26], which was
applicable at concentrations close to the indoor environment. In the study conducted by
Seo et al. [26], the value of R2 of the Langmuir model was 0.9956, which could fit the
adsorption of toluene by activated carbon at 0.4–1.6 ppm. Mobasser et al. [27] conducted
adsorption tests on activated carbon, zeolite, and organosilica with a toluene concentration
of 0.02–1 ppm and found that the Freundlich model fitted well. The D–R model was less
frequently used than the Langmuir and Freundlich models because of its complex form
and the greater number of parameters involved. The Langmuir and Freundlich models
were used to describe the adsorption of toluene at the ppm level, whereas it should still be
verified which isotherm was the most suitable.
Sustainability 2023, 15, x FOR PEER REVIEW 7 of 17
Sustainability 2023, 15, 1555 7 of 16

Figure
Figure4. Application of the
4. Application adsorption
of the isotherms
adsorption for toluene
isotherms adsorption
for toluene in the
adsorption inprevious studies
the previous stud-
[15,17,25–32].
ies [15,17,25–32].

3.2.3.2. Breakthrough
Breakthrough Models
Models
TheThe criticalparameter
critical parameterof of the mass
masstransfer
transfercoefficient cancan
coefficient be obtained by breakthrough
be obtained by break-
models
through describing
models the dynamic
describing breakthrough
the dynamic curvescurves
breakthrough of adsorption. With With
of adsorption. a simple form
a sim-
pleand
formeasily
andavailable parameters,
easily available the Yoon–Nelson
parameters, model was
the Yoon–Nelson widely
model wasused. Theused.
widely application
The
of this model
application of thiswas not restricted
model to the absorbent’s
was not restricted physicalphysical
to the absorbent’s and chemical properties
and chemical prop-[33].
The expression of the Yoon–Nelson model is:
erties [33]. The expression of the Yoon–Nelson model is:
1 CCb b
t btb==τ τ+ + 1 ln(
ln( )) (7)(7)
KKv v CCi i−−CCbb
whereKKv v isis the
where the mass
mass transfer
transfer coefficient, 1/h;t btb is the breakthrough time, h; ττ isisthe
coefficient, 1/h;
breakthrough time required for 50% breakthrough,
the breakthrough time required for 50% breakthrough, h; Cbh;is C
the is
breakthrough concentration
the breakthrough con-
b
of adsorbate, ppm; and Ci is the initial concentration of adsorbate, ppm.
centration
Theofmassadsorbate,
transferppm; and C
coefficient i
wasiscalculated
the initial by
concentration of adsorbate,
the Wheeler–Jonas model ppm.
expressed in
The mass
Equation (1).transfer coefficient
The expression waswas calculated
similar by the Wheeler–Jonas
to the Yoon–Nelson model, and model
couldexpressed
be regarded
in as
Equation (1). expression
a complex The expression
of thewas similar to the
Yoon–Nelson Yoon–Nelson model, and could be re-
model.
garded as a complex expression of the Yoon–Nelson model.

3.3. Prediction Procedure


Sustainability 2023, 15, 1555 8 of 16

3.3. Prediction Procedure


Step 1: The equilibrium adsorption capacity was predicted at low concentrations
(0.5 and 1.0 ppm or lower) by fitting adsorption isotherms and experimental data. The
actual adsorption capacity was calculated by integrating the breakthrough curve using
Equation (3). We compared different fitting results by isotherms and chose the one with
the highest accuracy in predicting adsorption capacity.
Step 2: The mass transfer coefficient at low concentrations was obtained in several
ways. The critical parameter was usually obtained by the breakthrough model describing
the dynamic adsorption properties in Section 3.2. We were also able to obtain it from
another empirical formula. Wood and Lodewyckx [34] proposed an equation to calculate
the mass transfer coefficient Kv :

−1.5 qe n
Kv = αβ0.33 v0.75
L dp ( ) (8)
M
where α is the overall correlation coefficient; β is the affinity coefficient; v L is the linear flow
rate through the adsorption column, cm/s; d p is the diameter of activated carbon; M is the
molecular weight; and n is the adjustable exponent.
Compared with the model mentioned in Section 3.2, the Wood–Lodewyckx equation
took into account not only the parameters of the adsorption bed (d p ) and adsorption
conditions (v L ), but also the parameters of the adsorbate properties (β, M).
Then, we summarized four methods for obtaining the mass transfer coefficient:
- Method 1: Kv was calculated by fitting experimental data and the Yoon–Nelson model
at high concentrations (2.0–5.0 ppm). The linear function relationship of Kv and Ci
was established to predict Kv at low concentrations (0.5 and 1.0 ppm or lower).
- Method 2: The procedure was similar to method 1, but contrastingly, the influence of
the quadratic function relationship of Kv and Ci on the prediction was studied.
- Method 3: Considering the impact of adsorbent characteristics and adsorption con-
ditions, we used the Wheeler–Jonas model to obtain the mass transfer coefficient at
high concentrations (2.0–5.0 ppm). The linear function relationships of Kv and Ci were
used to obtain the predicted Kv at low concentrations.
- Method 4: The transfer coefficient was calculated by the Wheeler–Jonas model at high
concentrations (2.0–5.0 ppm). The Wood–Lodewyckx correlation, which took into
account the properties of the adsorbate, was used to obtain the predicted values at
low concentrations.
Step 3: After obtaining the two key parameters and inputting the other parameters
into Equation (1), we were able to obtain the corresponding breakthrough concentration at
different moments as well as the prediction curves at low concentrations.

4. Results
4.1. Validation of Adsorption Isotherms
The breakthrough curves of toluene in the GAC with inlet concentrations of 0.5, 1.0,
2.0, 3.0, 4.0, and 5.0 ppm are shown in Figure 5. The breakthrough curves show S-shapes.
In theory, the mass transfer of toluene into the activated carbon follows three steps [35]:
(1) Film diffusion: toluene molecules pass through a layer of gas film on the absorbent
surface before being adsorbed. Contaminant flux through the film is directly proportional
to the linear concentration gradient across the film. (2) Pore diffusion: this occurs within
the interior of activated carbon; toluene molecules diffuse within the pore space from
the surface of activated carbon. The pore diffusion rate is influenced by the gas-phase
concentration of toluene [36]. For small pores, the number of molecule–wall collisions
exceeds that of intermolecular collisions, and the Knudsen diffusion dominates. For large
pores, more intermolecular collisions occur, and the viscous flow dominates [37]. (3) Surface
diffusion: the toluene molecules are adsorbed on the active sites on the inner surface of
the pores of the activated carbon, and they continue to diffuse along the adsorbent surface
Sustainability 2023, 15, 1555 9 of 16

within the pores. Mass transfer resistances at the three stages are different, resulting in
different diffusion rates. The stage with the slowest diffusion rate significantly affects the
Sustainability 2023, 15, x FOR PEER REVIEW 9 of 17
mass transfer process. Kim et al. [38] indicated that the rate of gas adsorption by activated
carbon was determined by pore diffusion. Therefore, as the inlet concentration of toluene
was increased, the diffusion rate of toluene in the activated carbon was increased, and the
thatphysical adsorption was
of intermolecular accelerated
collisions, and intheorder to shorten
Knudsen the time
diffusion necessary For
dominates. to reach
largethe
pores,
moreadsorption equilibrium.
intermolecular collisions occur, and the viscous flow dominates [37]. (3) Surface dif-
Thetoluene
fusion: the equilibrium adsorption
molecules are capacities
adsorbedofontoluene at 0.5,sites
the active 1.0, 2.0, 3.0, 4.0,
on the innerand 5.0 ppm
surface of the
on the GAC were 69.00, 102.08, 134.26, 150.03, 159.38, and 165.58 mg/g, respectively. These
pores of the activated carbon, and they continue to diffuse along the adsorbent surface
values were calculated by the numerical integration method, as in Equation (3), and are
within
showntheinpores.
Table Mass
2. Thetransfer
isothermsresistances at the three
for the Langmuir, stages and
Freundlich, are different,
D–R models resulting
were in
different
obtained by fitting the experimental data for toluene adsorption on the GAC using the the
diffusion rates. The stage with the slowest diffusion rate significantly affects
mass transfer process.
least-square Kim et 6).
method (Figure al. Table
[38] indicated that the
3 summarizes theresults
rate ofofgas
theadsorption bycan
regression. It activated
be
carbon
seenwas determined
that both by pore
the Langmuir anddiffusion.
FreundlichTherefore, as the
models fit with theinlet concentration
experimental data ofof toluene
GAC.
wasThe R2 of the the
increased, Langmuir model
diffusion was
rate ofas high asin
toluene ~0.99. It can be speculated
the activated carbon was thatincreased,
the adsorption
and the
of toluene
physical molecules
adsorption on the
was surface of in
accelerated activated carbon
order to was monolayer
shorten adsorption.toThe
the time necessary R2 the
reach
of the D–R
adsorption model was 0.9223 for GAC, which indicated that micropore filling might not
equilibrium.
occur during the adsorption process.

Figure
Figure 5. Breakthrough
5. Breakthrough curvesof
curves oftoluene
toluene adsorption
adsorptionon
onGAC
GACat at
various lowlow
various concentrations.
concentrations.

Table Equilibrium adsorption


The2.equilibrium adsorptioncapacities (mg/g).
capacities of toluene at 0.5, 1.0, 2.0, 3.0, 4.0, and 5.0 ppm
on theSample
GAC were 0.5 69.00,
ppm
102.08, 134.26,
1.0 ppm
150.03,
2.0 ppm
159.38, and 165.58 4.0
3.0 ppm
mg/g,
ppm
respectively.
5.0 ppm
These
values were calculated by the numerical integration method, as in Equation (3), and are
shown GACin Table 2. 69.00 102.08
The isotherms for the134.26
Langmuir, 150.03Freundlich,159.38 165.58
and D–R models were
obtained by fitting the experimental data for toluene adsorption on the GAC using the
least-square methodresults
Table 3. Regression (Figure 6).three
using Table 3 summarizes the results of the regression. It can be
isotherms.
seen that both the Langmuir and Freundlich models fit with the experimental data of GAC.
2
The R2Sample
of the Langmuir Model
model was as high as ~0.99. Equation
It can be speculated thatR the adsorp-
tion of toluene molecules on the surface ofqactivated
Langmuir carbon was monolayer
e = 0.0051Ci + 0.018
adsorption.
0.9899
The R2 ofGAC
the D–R model was
Freundlich0.9223 for GAC, which indicated
qe = 61.64Ci0.3489 that micropore filling
0.9591 might
not occur during the adsorption
D–R
process.
q = 151.33 exp[−4.435 ln(1 + 1/C )] 0.9223
e i

Table 2. Equilibrium adsorption capacities (mg/g).

Sample 0.5 ppm 1.0 ppm 2.0 ppm 3.0 ppm 4.0 ppm 5.0 ppm
GAC 69.00 102.08 134.26 150.03 159.38 165.58

Table 3. Regression results using three isotherms.

Sample Model Equation R2


Sustainability 2023, 15, x FOR PEER REVIEW 10 of 17

Sustainability 2023,
Sustainability 15,15,
2023, x FOR
1555 PEER REVIEW 10 10
of of
1716

Figure 6. Adsorption isotherms fitted by the Langmuir, Freundlich, and D–R models.

4.2. Validation of Breakthrough Models and Predicted Models


Method 1: The prediction curve was established based on the Wheeler–Jonas equa-
tion, expressed
Figure
Figure 6. 6. as Equation
Adsorption
Adsorption isotherms (1),
isotherms
fittedinby
fittedwhich
by the
the the valueFreundlich,
Langmuir,
Langmuir, of equilibrium
Freundlich, andD–R
and models.capacity qe
adsorption
D–Rmodels.
could be obtained from the Langmuir equation in Section 4.1. Next, it was essential to
4.2. Validation of of BreakthroughModels Modelsand andPredicted
PredictedModels
Models
calculate the mass transfer coefficient K v . This method used the Yoon–Nelson model to
4.2. Validation Breakthrough
Method 1: The prediction curve was established based on the Wheeler–Jonas equation,
Method 1: The prediction curve was established based on the Wheeler–Jonas equa-
expressed
predict
tion, K as
expressed
Equation
v at aslow
(1), in which the value of equilibrium adsorption capacity qe could
concentrations.
Equation (1), in which the value of equilibrium adsorption capacity qe
be obtained from the Langmuir equation in Section 4.1. Next, it was essential to calculate
could The
the mass
calculation
be obtained of
transfer coefficient
K was basedequation
from thev Langmuir
Kv . This method
on the least-square
used in the
Section technique.
4.1.
Yoon–Nelson itThe
Next,model wastoYoon–Nelson
essential
predict Kto v at
equation the mass transfer coefficient K v . This method used the Yoon–NelsonThe
yielded
low concentrations.
calculate linear curves using the Levenberg–Marquardt algorithm. model break-
to
through Thetime t b was
calculation of the
Kv wasabscissa,
basedthe logarithmic
on the least-square concentration
technique. The was Yoon–Nelson
the ordinate, equa-and
predict
tion K
yielded at low concentrations.
linear curves using the Levenberg–Marquardt
the slope of the curve was K v . The relationship between K v and C i was determined
v algorithm. The breakthrough
timeThetb was
by linear calculation
function based of K von
the abscissa, thewas
logarithmic
the based
resultson
concentration
at the
2.0,least-square
was the ordinate,
3.0, 4.0, andtechnique.
5.0 ppm. The Theand the slope of
Yoon–Nelson
linear function
the curve was Kv . The relationship between Kv and Ci was determined by linear function
equation
could
basedbe
yielded
onexpressed
linear
the resultsasat K
curves
2.0, = 0 .0 0 1 6 C i ppm.
using the
v 3.0, 4.0, and 5.0
+ 0 .0The
Levenberg–Marquardt
0 4 5linear
(Figure algorithm.
7). The
function could
The
excellent fitbreak-
be expressedfound as
through time t was the abscissa, the logarithmic concentration
fit found K
was the ordinate, and
Kv =the
that 0.0016C
R was
2
i+
b 0.0045 The
0.9107. (Figure
linear7). relationship
The excellentbetween that
v
the
and R C
2 was
i
0.9107.
was also The linear
obtained
the
at slope of the
relationship
concentrations curve
between was
of 9,K40,
v andK
and . The
vCi90was relationship
ppmalsoinobtained between
the study at of K v and
concentrations
Ligotski C [15].
of
et al. was determined
i9, 40,Itand 90 ppm
seemed in
that
by the study
linear of Ligotski
function basedet al.
on [15].
the It seemed
results at that3.0,
2.0, the 4.0,
relationship
and 5.0 for KvThe
ppm. waslinear
suitable, since it
function
the relationship for K v was suitable, since it was obtained based on the experimental
was be
could
data
obtained
expressed
at the
based as onKthe
concentration =experimental
v of the ppm level
+and
0 .0 0 1 6 C i data 0 at the
.0the 5concentration
0 4excellent
(Figure of the ppm level and the
fit. 7). The excellent fit found
excellent fit.
that the R2 was 0.9107. The linear relationship between K v and C i was also obtained
at concentrations of 9, 40, and 90 ppm in the study of Ligotski et al. [15]. It seemed that
the relationship for K v was suitable, since it was obtained based on the experimental
data at the concentration of the ppm level and the excellent fit.

Linearrelationship
Figure7.7.Linear
Figure relationship between mass transfer coefficient KK
between mass v and
v
inlet
and concentration
inlet of toluene
concentration of tol-
Ci in Method 1.
uene C i in Method 1.
Then, the structural parameters of activated carbon, such as mGAC and ρb , the inlet
concentration of toluene Ci , the equilibrium capacity qe , and the mass transfer coefficient
Figure 7. Linear
Kv were inputrelationship between mass transfer
into the Wheeler–Jonas coefficient
equation K v (1))
(Equation andto
inlet concentration
generate of tol-
the predicted
uene C in Method
breakthrough
i
curve1. (Figure 8).
According to Figure 8, the deviation between the prediction and the measured data
was significant. Especially when the toluene concentration was at 0.5 and 1.0 ppm, the
predicted breakthrough time significantly exceeded the tested value. This equation, ex-
pressed as K v = 0 .0 0 1 6 C i + 0 .0 0 4 5 in Step 2, was indicated to be unsuitable for
Sustainability 2023, 15, 1555 predicting the mass transfer coefficient. 11 of 16

Figure 8. Comparison of predicted and experimental breakthrough curves at the toluene concentra-
Figure 8. Comparison
tion of 0.5~5.0 ppm inofMethod
predicted
1. and experimental breakthrough curves at the toluene concentra-
tion of 0.5~5.0 ppm in Method 1.
According to Figure 8, the deviation between the prediction and the measured data was
Method Especially
significant. 2: Observing
whenthethe prediction results inwas
toluene concentration Figure
at 0.58, it 1.0
and could
ppm,bethe
speculated
predicted that
the linear expression could not accurately reveal the relationship between K v and
breakthrough time significantly exceeded the tested value. This equation, expressed as C .
i
Kv = 0.0016Ci + 0.0045 in Step 2, was indicated to be unsuitable for predicting the mass
Then, determination
transfer coefficient. of the quadratic function relationship was attempted using Method
2. The Method
equation2: was K v =the
Observing − 0.0006 C i 2results
prediction + 0.0055 C i −8, 0.0016
in Figure it could beand the R2that
speculated wasthe0.9997
linear expression could not accurately reveal the relationship
(Figure 9). Then, K v could be calculated by this equation at 0.5 and v1.0 ppm. between K and C i Then,
. The other
determination of the quadratic function relationship was attempted using Method 2. The
steps were the
wassame
Kv = as those in
−0.0006C 2 Method 1. It was found that 2 the predicted curves were still
equation i + 0.0055Ci − 0.0016 and the R was 0.9997 (Figure 9). Then,
farKfrom the measured data in Figure 10, especially when
v could be calculated by this equation at 0.5 and 1.0 ppm. The other the toluene concentration
steps were the same as was
Sustainability 2023, 15, x FOR PEER REVIEW
0.5those
and 1.0 ppm. It1.could
in Method It wasbe speculated
found that the Yoon–Nelson
that the predicted curves were stillmodel was
far from thenot 12 of 17
appropriate
measured
fordata in Figure 10, especially when the toluene concentration was 0.5 and 1.0 ppm. It could
the prediction.
be speculated that the Yoon–Nelson model was not appropriate for the prediction.

Figure 9. Quadratic relationship between mass transfer coefficient Kv and inlet concentration of
Figure C
toluene 9. in
i Method 2.relationship between mass transfer coefficient K v and inlet concentration o
Quadratic
toluene Ci in Method 2.
Figure 9. Quadratic relationship between mass transfer coefficient Kv and inlet concentration of
Sustainability 2023, 15, 1555 toluene Ci in Method 2. 12 of 16

Figure
Figure 10.10. Comparison
Comparison of of predicted
predicted and
and experimental
experimental breakthrough
breakthrough curves
curves with
with toluene
toluene concentra-
concentra-
tions
tions of of 0.5~5.0
0.5~5.0 ppmppm
in in method
method 2. 2.

Method
Method 3: 3:
TheThe Wheeler–Jonas
Wheeler–Jonas model
model was
was used
used to to calculate
calculate thethe mass
mass transfer
transfer coeffi-
coeffi-
cient K at 0.5 and 1.0 ppm. By fitting the experimental data and Equation (1), the equation
cient K vv at 0.5 and 1.0 ppm. By fitting the experimental data and Equation (1), the equa-
between Kv and Ci was expressed as Kv = −0.0004Ci + 0.0072, and the values are shown
tion 4. K v and C i was expressed as K v = − 0 .0 0 0 4 C i + 0 .0 0 7 2 , and
between
in Table
the values are shown in Table 4.
Table 4. Results of the predicted mass transfer coefficient Kv in Method 3.
Table 4. Results of the predicted mass transfer coefficient Kv in Method 3.
Toluene Concentration (ppm) 0.5 1.0 2.0 3.0 4.0 5.0
Toluene Concentration
Kv (1/min) (ppm) 0.5
0.0070 1.0
0.0068 2.0
0.0064 3.0
0.0060 4.0
0.0056 5.0
0.0052
K v (1/min) 0.0070 0.0068 0.0064 0.0060 0.0056 0.0052
Sustainability 2023, 15, x FOR PEER REVIEW 13 of 17
Then, the calculated qe in Step 1 and Kv in Table 4 were input into the Wheeler–Jonas
model, and the predicted curves are shown in Figure 11. Compared with the previous two
Then, the calculated q in Step 1 and K in Table 4 were input into the Wheeler–
methods, this method wase able to reflect the vadsorption process more accurately. Due to
Jonas model, and
characteristics theadsorption
of the
the Wheeler–Jonas predicted
model curves
bed
considering are
( m GAC , ρ b ), parameters,
shown
detailed in Figure 11.condition
adsorption Compared
such with
, Q the previ-of
( Ccharacteristics
as the i
), and
ous
thetwo methods,bed
adsorption this method was able to reflect the adsorption
(Ci , Q), process more accurately.
equilibrium adsorption(m GAC , ρb ),itadsorption
capacity, condition
is more widely applicable. Theand equilibrium
breakthrough adsorption
curves
Due to
of activated
the Wheeler–Jonas
capacity, itcarbon
is more widely
and
model
applicable. The
the breakthrough
considering detailed
breakthrough
time curves
under different
parameters, such
carbonasand
of activatedconditions
experimental
the
the
canbreakthrough
be obtained by time under the
changing different
different experimental
parametersconditions can be obtained by changing
of the equation.
the different parameters of the equation.

Figure 11. Comparison of predicted and experimental curves with toluene concentrations of 0.5~5.0 ppm
Figure 11. Comparison of predicted and experimental curves with toluene concentrations of 0.5~5.0
ppmin in
Method 3. 3.
Method

Method 4: Step 1 was to calculate qe by the Langmuir model, and Step 2 was to
calculate K v by Equation (8). K v could be calculated by fitting experimental data and
the Wheeler–Jonas model, and the equilibrium adsorption capacities were obtained by the
Langmuir model at 2.0, 3.0, 4.0, and 5.0 ppm. The relationship was built between K v and
Sustainability 2023, 15, 1555 13 of 16

Figure 11. Comparison of predicted and experimental curves with toluene concentrations of 0.5~5.0
Method4:3. Step 1 was to calculate qe by the Langmuir model, and Step 2 was to
ppm inMethod
calculate Kv by Equation (8). Kv could be calculated by fitting experimental data and the
Wheeler–Jonas model, and the equilibrium adsorption capacities were obtained by the
Method 4: Step 1 was to calculate q by the Langmuir model, and Step 2 was to
Langmuir model at 2.0, 3.0, 4.0, and 5.0 ppm.eThe relationship was built between Kv and
qe , and theKpredicted
calculate v
by Equation wereK
values (8). could(Table
obtained
v
be calculated
5). Step 3 by
wasfitting experimental
to input data and
Kv and qe into
Equation (1) and obtain the predicted curves (Figure 12).
the Wheeler–Jonas model, and the equilibrium adsorption capacities were obtained by the
Langmuir model at 2.0, 3.0, 4.0, and 5.0 ppm. The relationship was built between K v and
Table 5. Results of the predicted mass transfer coefficient in Method 4.
qe , and the predicted values were obtained (Table 5). Step 3 was to input K v and qe
Toluene Concentration (ppm) 0.5 1.0 2.0 3.0 4.0 5.0
into Equation (1) and obtain the predicted curves (Figure 12).
Kv (1/min) 0.0055 0.0057 0.0060 0.0065 0.0080 0.0109

Figure 12. Comparison of predicted and experimental curves with toluene concentrations of 0.5~5.0 ppm
Figure 12. Comparison of predicted and experimental curves with toluene concentrations of 0.5~5.0
in Method 4.
ppm in Method 4.
The predicted curves in Method 4 also achieved good results. From the user’s point
Table 5. Results
of view, peopleof the
are predicted
more massabout
concerned transfer
thecoefficient
service lifeinofMethod 4. The service life was
the filter.
derived from the breakthrough curves. The predicted and actual service lives are listed
in Toluene
Table 6 to Concentration
more effectively(ppm) 0.5
compare the accuracy 1.0
of the two 2.0methods.3.0Regarding
4.0 the 5.0
aspect of service life, the predicted value in method 4 was closer to the measured data.
When the toluene concentration was 0.5 ppm, the service life was improved by 572 min.
At 1.0 ppm, the predicted service life was improved by 187 min. However, on the whole,
the deviation between the measured and the predicted values was large at the beginning
and the end of adsorption, which were the sources of the deviation in the prediction. In
the early adsorption stage, there were many unoccupied adsorption sites on the surface of
the activated carbon, and the toluene molecules had a higher probability of contacting and
being adsorbed on the surface. Therefore, the actual breakthrough time was later than the
predicted value. At the end of adsorption, numerous toluene molecules were adsorbed,
and there were few empty adsorption sites on the surface of the activated carbon. Therefore,
the toluene molecules would directly penetrate the adsorption column, so the activated
carbon would quickly reach saturation and adsorption would no longer occur. As a result,
the actual service life was significantly shorter than the predicted value.
Sustainability 2023, 15, 1555 14 of 16

Table 6. Measured and predicted service life.

Toluene Measured Service Predicted Time in Predicted Time in


Concentration (ppm) Life (min) Method 3 (min) Method 4 (min)
5.0 904 931 901
4.0 950 1076 1067
3.0 1094 1302 1307
2.0 1452 1690 1678
1.0 1869 2495 2308
0.5 2487 3326 2754

4.3. Modification of Prediction Method


Although the predicted values in Method 4 showed good agreement with the mea-
sured data, a deviation still existed in predicted service life. To improve the prediction
accuracy, a relationship was built between the predicted deviation and toluene concentra-
tion (Figure 13), and the expression was ∆t = −13.197Ci2 − 3.1937Ci + 339.6 (R2 = 0.8093).
The predicted value in Method 4 minus the deviation ∆t computed by the equation was the
modified value. The modified predicted values are shown in Table 7. After modification,
the precision of prediction was improved and the relative deviation was no more than 7%.15 of 17
Sustainability 2023, 15, x FOR PEER REVIEW
When the concentration was 0.5 and 1.0 ppm, the accuracy was improved by 8.01% and
17.29%, respectively.

Figure
Figure 13. Relationship
Relationshipbetween
between deviation
deviation andand toluene
toluene concentration.
concentration.

Table 7. Results
Table 7. Resultsofofimproved
improvedpredicted service
predicted life. life.
service

Toluene
Toluene Measured Service
Measured Modified Value
Service Modified Deviation
DeviationRelative Relative
Concentration (ppm)
Life (min) Value (min) (min) Deviation (%)
Concentration (ppm) Life (min) (min) (min) Deviation (%)
5.0 904 907 3 0.36
5.0 904 907 3 0.36
4.0 950 951 1 0.14
4.0
3.0 950
1094 1096 951 2 1 0.16 0.14
2.0
3.0 1452
1094 1398 1096 − 54 2 − 3.75 0.16
1.0 1869 1985 116 6.20
2.0
0.5
1452
2487 2419
1398 −68
−54 −2.72
−3.75
1.0 1869 1985 116 6.20
0.5
5. Conclusions 2487 2419 −68 −2.72
In this paper, the toluene adsorption by granular activated carbon at the ppm level
5.was
Conclusions
tested. The methods of toluene adsorption prediction in activated carbon, based on the
Wheeler–Jonas model,
In this paper, thewere compared
toluene and modified
adsorption using activated
by granular the experimental
carbondata. Some
at the ppm level
conclusions can be drawn as follows:
was tested. The methods of toluene adsorption prediction in activated carbon, based on
the Wheeler–Jonas model, were compared and modified using the experimental data.
Some conclusions can be drawn as follows:
1. The Langmuir and Freundlich models were both suitable for predicting the equilib-
Sustainability 2023, 15, 1555 15 of 16

1. The Langmuir and Freundlich models were both suitable for predicting the equilib-
rium adsorption capacity, and the Langmuir model was preferred under the consid-
ered conditions.
2. Using the equilibrium adsorption capacity obtained from the Langmuir model and the
mass transfer coefficient from the Wood–Lodewyckx correlation was the most accurate
method for the prediction of the breakthrough curve based on the Wheeler–Jonas model.
3. The service life was obtained based on the predicted breakthrough curve, which could
be significantly improved using the relationship with the inlet concentration of the
adsorbate to obtain a value close to the actual service life.
Usually, the adsorption of gaseous filters is conducted under complex conditions.
The methodology presented in this study requires more validation regarding different
adsorbents and operating conditions.

Author Contributions: Conceptualization, Y.S.; Methodology, Y.S. and Q.R.; Data curation, M.W.;
Writing—original draft, Y.S. and Q.D.; Writing—review & editing, Y.S. All authors have read and
agreed to the published version of the manuscript.
Funding: This research was funded by the National Natural Science Foundation of China grant
number 51908402.
Data Availability Statement: The data that support the findings of this study are available on request
from the corresponding author, Sheng Y., upon reasonable request.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. World Health Organization (WHO). Indoor Air Quality: Organic Pollutants (Euro Reports and Studies No. 111); WHO Regional Office
for Europe: Copenhagen, Denmark, 1989.
2. Sarigiannis, D.A.; Karakitsios, S.P.; Gotti, A.; Liakos, I.L.; Katsoyiannis, A. Exposure to major volatile organic compounds and
carbonyls in European indoor environments and associated health risk. Environ. Int. 2011, 37, 743–765. [CrossRef]
3. Zhou, C.B. Impact of Natural Ventilation on Indoor Air Quality in High-Rise Apartments. Ph.D. Thesis, Tianjin University, Tianjin,
China, 2016. (In Chinese)
4. Bastani, A.; Lee, C.S.; Haghighat, F.; Flaherty, C.; Lakdawala, N. Assessing the performance of air cleaning devices—A full-scale
test method. Build. Environ. 2010, 45, 143–149. [CrossRef]
5. Ramirez, D.; Qi, S.Y.; Rood, M.J. Equilibrium and heat of adsorption for organic vapors and activated carbons. Env. Sci. Technol.
2005, 39, 5864–5871. [CrossRef]
6. Haghighat, F.; Lee, C.S.; Pant, B.; Bolourani, G.; Lakdawala, N.; Bastani, A. Evaluation of various activated carbons for air
cleaning—Towards design of immune and sustainable buildings. Atmos. Environ. 2008, 42, 8176–8184. [CrossRef]
7. Zhu, L.L.; Shen, D.K.; Luo, K.H. A critical review on VOCs adsorption by different porous materials: Species, mechanisms and
modification methods. Hazard. Mater. 2020, 389, 122102. [CrossRef] [PubMed]
8. Vizhemehr, A.K.; Haghighat, F.; Lee, C.S. Gas-phase filters breakthrough models at low concentration—Effect of relative humidity.
Build. Environ. 2014, 75, 1–10. [CrossRef]
9. Zhi, Y.; Liu, J.J. Analysis of chemical filter performance and activated carbon microstructure at low concentration. Build. Environ.
2020, 169, 106563. [CrossRef]
10. Veksha, A.; Sasaoka, E.; Uddin, M.A. The influence of porosity and surface oxygen groups of peat-based activated carbons on
benzene adsorption from dry and humid air. Carbon 2009, 47, 2371–2378. [CrossRef]
11. ISO 10121-2; Test Methods for Assessing the Performance of Gas-Phase Air Cleaning Media and Devices for General Ventilation-
Part 2:Gas-Phase Air Cleaning Devices, Information Handling Services. International Organization for Standardization: Geneva,
Switzerland, 2013.
12. ASHRAE Standing Standard Project Committee 145; Laboratory Test Method for Assessing the Performance of Gas-Phase Air-
Cleaning Systems: Loose Granular Media. American Society of Heating, Refrigerating and Air-Conditioning Engineers, Inc.:
Atlanta, GA, USA, 2015.
13. Vizhemehr, A.K.; Haghighat, F. Modeling of gas-phase filter model for high- and low-challenge gas concentrations. Build. Environ.
2014, 80, 192–203. [CrossRef]
14. Shaverdi, G.; Haghighat, F.; Ghaly, W. Development and systematic validation of an adsorption filter model. Build. Environ. 2014,
73, 64–74. [CrossRef]
15. Ligotski, R.; Sager, U.; Schneiderwind, U.; Asbach, C.; Schmidt, F. Prediction of VOC adsorption performance for estimation of
service life of activated carbon based filter media for indoor air purification. Build. Environ. 2019, 149, 146–156. [CrossRef]
Sustainability 2023, 15, 1555 16 of 16

16. Yao, S.Y.; Li, Z.P.; Wang, S.L.; Yuan, Q.L.; Liu, L.L. Multi-model comparative analysis of nitrobenzene breakthrough curve in the
sandy strata. Geotech. Investig. 2020, 48, 40–46. (In Chinese)
17. Shiue, A.; Den, W.; Kang, Y.H.; Hu, S.C.; Jou, G.T.; Lin, C.H.; Hu, V.; Lin, S.I. Validation and application of adsorption breakthrough
models for the chemical filters used in the make-up air unit (MAU) of a cleanroom. Build. Environ. 2011, 46, 468–477. [CrossRef]
18. Chen, W.J.; Wan, W.R.; Xu, X.F.; Pu, Y.X.; Hu, Y.L.; Dai, H. Investigation on indoor air pollution of VOCs caused by indoor
decoration in city of Inner Mongolia. J. Med. Pest Control. 2017, 33, 201–206. (In Chinese)
19. Brown, S.K. Volatile Organic Pollutants in New and Established Buildings in Melbourne, Australia. Indoor Air 2002, 12, 55–63.
[CrossRef]
20. Jonas, L.A.; Rehrmann, J.A. Predictive equations in gas adsorption kinetics. Carbon 1973, 1, 59–64. [CrossRef]
21. Liu, Y.; Shen, L. From Langmuir kinetics to first- and second-order rate equations for adsorption. Langmuir 2008, 24, 11625–11630.
[CrossRef]
22. Ebato, M.; Yonebayashi, K. Identification of adsorption conditions fulfilling the correlation between freundlich isotherm coeffi-
cients, log K-F and 1/n by computer simulation. Soil Sci. Plant Nutr. 2004, 50, 171–179. [CrossRef]
23. Lin, D.R.; Hu, L.J.; Xing, B.S.; You, H.; Loy, D.A. Mechanisms of Competitive Adsorption Organic Pollutants on Hexylene-Bridged
Polysilsesquioxane. Materials 2015, 8, 5806–5817. [CrossRef]
24. Gimeno, O.; Plucinski, P.; Kolaczkowski, S.T.; Rivas, F.J.; Alvarez, P.M. Removal of the herbicide MCPA by commercial activated
carbons: Equilibrium, kinetics, and reversibility. Ind. Eng. Chem. Res. 2003, 42, 1076–1086. [CrossRef]
25. Pei, J.; Zhang, J.S.S. Determination of adsorption isotherm and diffusion coefficient of toluene on activated carbon at low
concentrations. Build. Environ. 2012, 48, 66–76. [CrossRef]
26. Seo, J.; Kato, S.; Atakab, Y.; Chino, S. Performance test for evaluating the reduction of VOCs in rooms and evaluating the lifetime
of sorptive building materials. Build. Environ. 2009, 44, 207–215. [CrossRef]
27. Mobasser, S.; Wager, Y.; Dittrich, T.M. Indoor Air Purification of Volatile Organic Compounds (VOCs) Using Activated Carbon,
Zeolite, and Organosilica Sorbents. Ind. Eng. Chem. Res. 2022, 61, 6791–6801. [CrossRef]
28. Elsayed, Y.; Dallas, A.; Beatty, P. Experimental and predicted adsorption isotherms of 2,2,4-trimethylpentane and toluene on
activated carbon for industrial applications. Res. Chem. Intermed. 2015, 41, 1267–1282. [CrossRef]
29. Baytar, O.; Sahin, O.; Horoz, S.; Kutluay, S. High-performance gas-phase adsorption of benzene and toluene on activated carbon:
Response surface optimization, reusability, equilibrium, kinetic, and competitive adsorption studies. Environ. Sci. Pollut. Res.
2020, 27, 26191–26210. [CrossRef]
30. Lee, M.G.; Lee, S.W.; Lee, S.H. Comparison of vapor adsorption characteristics of acetone and toluene based on polarity in
activated carbon fixed-bed reactor. Korean J. Chem. Eng. 2006, 23, 773–778. [CrossRef]
31. Agnihotri, S.; Rood, M.J.; Rostam-Abadi, M. Adsorption equilibrium of organic vapors on single-walled carbon nanotubes. Carbon
2005, 43, 2379–2388. [CrossRef]
32. Huang, S.S.; Deng, W.; Zhang, L.; Yang, D.Y.; Gao, Q.; Tian, Z.F.; Guo, L.M.; Ishihara, T. Adsorptive properties in toluene removal
over hierarchical zeolites. Microporous Mesoporous Mater. 2020, 302, 110204. [CrossRef]
33. Yoon, Y.H.; Nelson, J.H. Application of gas adsorption kinetics. I. A theoretical model for respirator cartridge service life. Am. Ind.
Hyg. Assocaition J. 1984, 45, 509–516. [CrossRef]
34. Lodewyckx, P.; Wood, G.O. An extended equation for rate coefficients for adsorption of organic vapors and gases on activated
carbons in air-purifying respirator cartridges. Am. Ind. Hyg. Assoc. J. 2003, 64, 646–650.
35. Álvarez-Gutiérrez, N.; Gil, M.V.; Rubiera, F.; Pevida, C. Kinetics of CO2 adsorption on cherry stone-based carbons in CO2 /CH4
separations. Chem. Eng. J. 2017, 307, 249–257. [CrossRef]
36. Choy, K.K.; Porter, J.F.; Mckay, G. A Film-Pore-Surface Diffusion Model for the Adsorption of Acid Dyes on Activated Carbon.
Adsorption 2001, 7, 231–247. [CrossRef]
37. Brito, J.; Zahn, A. An unheated permeation device for calibrating atmospheric VOC measurements. Atmos. Meas. Tech. 2011,
4, 2143–2152. [CrossRef]
38. Kim, W.K.; Younis, S.A.; Kim, K.H. A strategy for the enhancement of trapping efficiency of gaseous benzene on activated carbon
(AC) through modification of their surface functionalities. Environ. Pollut. 2021, 270, 116239. [CrossRef]

Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual
author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to
people or property resulting from any ideas, methods, instructions or products referred to in the content.

You might also like