You are on page 1of 18

Adsorption Using Lime-Iron Sludge–Encapsulated

Calcium Alginate Beads for Phosphate Recovery


with ANN- and RSM-Optimized Encapsulation
Beverly S. Chittoo, Aff.M.ASCE 1; and Clint Sutherland, Ph.D., A.M.ASCE 2

Abstract: Excessive discharge of phosphates in municipal and industrial effluents into water bodies continues to amplify the rate and extent
Downloaded from ascelibrary.org by 41.102.31.224 on 08/24/21. Copyright ASCE. For personal use only; all rights reserved.

of eutrophication that is impairing aquatic ecosystems throughout the world. Consequently, research into technologies to combat the prob-
lem of eutrophication continues unabated. This study aimed to develop a protocol to encapsulate dewatered lime-iron sludge in calcium
alginate beads and assess and optimize its phosphate adsorption performance. Response surface methodology (RSM) and artificial neural
network (ANN) were used to optimize the encapsulation process through parameter variation. RSM was superior in capturing the nonlinear
behavior of the process. Numerical optimization in RSM revealed that maximum adsorption could be obtained from beads prepared using
0.25 g sodium alginate and 0.5 g lime-iron sludge in 25 mL of distilled water to produce a homogeneous mixture and added dropwise into a
solution of 0.31 g CaCl2 in 25 mL of distilled water. The accuracy of the RSM prediction was subsequently validated by laboratory experi-
ments that revealed a residual error of 2.9% and thus highlights the applicability of the model. Batch experiments were conducted and
modeled to expound the mechanisms of adsorption. Kinetic data were best simulated using the pseudo-second order model while equi-
librium data followed the Langmuir isotherm at room temperature and the Sips isotherm at higher temperatures. Physisorption, hydrogen
bonding, dipole interaction, and ligand exchange were the dominant attachment mechanisms while film and intraparticle diffusion were the
pertinent transport mechanisms. The beads exhibited a maximum monolayer adsorption capacity of 8.3 mg=g that compared well to other
phosphate-targeting adsorbents reported in the literature. DOI: 10.1061/(ASCE)EE.1943-7870.0001519. This work is made available
under the terms of the Creative Commons Attribution 4.0 International license, http://creativecommons.org/licenses/by/4.0/.
Author keywords: Response surface methodology; Artificial neural network; Phosphate; Adsorption; Lime-iron sludge; Alginate
hydrogel.

Introduction removal of phosphate from wastewater before its discharge into


water bodies is critical to curbing the problem of eutrophication
Anthropogenic activities have accelerated the rate and extent of (Akpor 2011).
eutrophication through the discharge of excessive nutrients such Removal of phosphate from water bodies can be accomplished
as phosphate in municipal and industrial effluents into aquatic through various treatment processes including ion exchange, dis-
ecosystems (Yang et al. 2008). This frequently stimulates the solved air floatation, membrane filtration, high-rate sedimentation,
extraordinary growth of blue-green algae, destroys aquatic life, and adsorption (Mohammed and Rashid 2012). Among these treat-
disrupts the natural food chain, and leads to deterioration of water ment techniques adsorption is drawing increasing attention and be-
quality (Chislock et al. 2013). The problem of eutrophication is coming an attractive technology because of its simplicity of design
of growing concern and affects several regions (Nixon et al. and operation, insensitivity to toxic pollutants, and its potential to
1986). The estimated cost of damages due to freshwater eutrophi- remove contaminants present in water even at low concentrations to
cation in England and Wales alone is estimated to be US$105– produce a high quality treated effluent (Chen et al. 2013). However,
160 million per annum (Pretty et al. 2003) while the cost of
the search for efficient and low-cost adsorbents that are readily
damages due to eutrophication in the United States is estimated
available remains a crucial issue of current research interest (Gadd
to be US$2.2 billion per annum (Dodds et al. 2008). It is also
2009).
reported that approximated 30% of the channel lengths of Irish
The efficiency of adsorbents for phosphate removal is controlled
rivers are polluted mainly from eutrophication (Walsh 2005). The
by the method of adsorption. According to Loganathan et al. (2014)
1 there are five methods of phosphate sorption: ion exchange (outer-
Instructor, Project Management and Civil Infrastructure Systems, Univ.
of Trinidad and Tobago, San Fernando Campus, Tarouba Link Rd., sphere surface complexation), ligand exchange (inner-sphere sur-
San Fernando, Trinidad and Tobago. Email: beverly.chittoo@utt.edu.tt face complexation), hydrogen bonding, surface precipitation,
2 and diffusion into the interior structure of the adsorbent. The au-
Assistant Professor, Project Management and Civil Infrastructure
Systems, Univ. of Trinidad and Tobago, San Fernando Campus, Tarouba thors went on to explain that ligand exchange is advantageous be-
Link Rd., San Fernando, Trinidad and Tobago (corresponding author). cause of its ability to remove large proportions of anions even from
ORCID: https://orcid.org/0000-0002-5569-325X. Email: clint.sutherland@ solutions of very low concentrations of target anions and in the
utt.edu.tt
presence of competing anions of lower selectivity. In ligand ex-
Note. This manuscript was submitted on July 6, 2018; approved on
October 23, 2018; published online on March 12, 2019. Discussion period change, covalent bonds form between anions such as phosphate
open until August 12, 2019; separate discussions must be submitted for and metallic cations on the surface of the adsorbent. It would, there-
individual papers. This paper is part of the Journal of Environmental fore, be advantageous to use metal oxide adsorbents for removing
Engineering, © ASCE, ISSN 0733-9372. phosphates from wastewater.

© ASCE 04019019-1 J. Environ. Eng.

J. Environ. Eng., 2019, 145(5): 04019019


Commercial adsorbents such as granular ferric oxide and efficient in RSM mainly because it requires few experiments (three
activated aluminum oxide have been reported by several authors levels per factor) and therefore saves time and resources. It has been
including Genz et al. (2004), Zhao et al. (2015), and Xie et al. successfully used to optimize encapsulation of several materials
(2015). However, a major drawback for the large-scale application including L-Asparaginase (Bahraman and Alemzadeh 2017), peat
of these commercial adsorbents is its high cost. An attractive alter- (Vecino et al. 2013), and α-Amylase (Dey et al. 2003) in calcium
native is the use of low-cost value-added adsorbents generated alginate beads.
from waste materials. Industrial by-products are a striking option ANN is a computational technique that attempts to simulate hu-
owing to their high oxide content, availability, and low cost. Its man intuition in making decisions and drawing conclusions given
reuse for phosphate removal not only aids in curbing the problem complex, irrelevant, and partial information (Saeh and Khairuddin
of eutrophication but is beneficial toward mitigating existing waste 2009). It has been successfully used to optimize encapsulation of
disposal issues. Low-cost materials such as blast furnace slag various materials including papain (Sankalia et al. 2005) and drugs
(Johansson and Gustafsson 2000), iron oxide tailings (Zeng et al. (Derakhshandeh et al. 2012) in calcium alginate beads.
2004), fly ash (Ugurlu and Salman 1998), and lime-iron sludge Adsorption from the aqueous phase onto the solid phase usually
(Chittoo and Sutherland 2017) have been reported to efficiently ad- involves three steps: (1) boundary layer mass transfer across the
sorb phosphate from aqueous solutions. The adsorption capacities liquid film surrounding the particle as film diffusion; (2) internal
Downloaded from ascelibrary.org by 41.102.31.224 on 08/24/21. Copyright ASCE. For personal use only; all rights reserved.

of these materials were reported to be influenced by its mineral diffusion/mass transport within the particle boundary as pore
composition, and the most effective were those with a high content and/or solid diffusion; and (3) adsorption within the particle and
of metal oxides such as aluminum hydroxide, iron oxide, and cal- on the external surface (Poots et al. 1976). The slowest of these
cium oxide. However, use of these low-cost adsorbents in pow- steps will be rate limiting and therefore determines the rate of the
dered form may not be practical for large-scale applications due reaction (Girish and Murty 2016). Consequently, knowledge of the
to the difficulty of separation from water, ultimate disposal, and mechanisms involved in the adsorption process is critical to under-
its potential reuse. It would, therefore, be worthwhile to encapsulate standing the suitability of the adsorbent as well as improving the
the adsorbent while at the same time maintaining its high adsorp- efficiency of the adsorption process (Sutherland and Venkobachar
tion capacity. 2013).
Alginate, a polysaccharide biopolymer composed of (1 → 4) The objectives of this investigation are: (1) to develop a protocol
linked α-L-guluronate and β-D-manuronate has attracted much and optimize the encapsulation of lime-iron sludge in calcium
attention for encapsulating materials due to its low cost and ability alginate beads using RSM and ANN; (2) to develop an empirical
to form stable hydrogels in the presence of divalent ions such as model to predict the phosphate adsorption capacity for different
Ba2þ , Ca2þ , and Fe2þ by cross linking (Lee and Mooney 2012). encapsulation conditions; and (3) to optimize the adsorption
The resulting porous hydrogels allow solutes to diffuse in and out process by varying operational parameters and by elucidating
consequently making contact with the encapsulated material. Addi- the transport and attachment mechanisms through batch kinetic,
tionally, encapsulation using alginate is advantageous because it is equilibrium, thermodynamic, and mass transfer studies.
nontoxic and biodegradable [A. Bezbaruah, T. B. Almeelbi, and
M. Quamme, US Patent No. 15/147,437 (2014)]. While entrapment
of adsorbents using alginate beads has been extensively investi- Materials and Methods
gated for the uptake of various metals (Wang et al. 2017), few stud-
ies have focused on the uptake of phosphate from solution. Sujitha Chemicals and Adsorbents
and Ravindhranath (2017) reported an exceptionally high phos-
phate adsorption capacity of 133.3 mg=g by calcium alginate beads Preparation of Adsorbent
doped with active carbon derived from the Achyranthes aspera Lime-iron sludge used in this study was obtained from a water
plant. However, activated carbon is expensive to produce and may treatment plant located in central Trinidad. After collection, the
not be the most viable option. Mahmood et al. (2015) studied the sludge was heated in an oven (ELE78-1215/01, ELE International,
use of alginate calcium-carbonate beads for phosphate adsorption UK) at 378 K for 24 h and then cooled to room temperature for
but reported a low phosphate adsorption capacity of only 0.62 mg=g. 72 h. It was then pulverized using a mortar and pestle and sieved
This study aims to encapsulate lime-iron sludge, a waste material, to pass through a 710-μm sieve.
in calcium alginate beads as a low-cost adsorbent and at the same The sludge was encapsulated in calcium alginate beads using the
time maintain a high adsorption capacity. Lime-iron sludge encap- ionic cross-linking technique previously reported by Mahmood
sulated alginate beads for phosphate adsorption is unreported. et al. (2015) with slight modification. Commercial grade sodium
Further, the extraordinarily high iron content of lime-iron sludge alginate [laboratory reagent (LR) grade, Breckland Scientific
obtained from central Trinidad makes this material an excellent Supplies, UK] was first dissolved in double distilled water and
candidate for investigating its encapsulation for phosphate removal. stirred to obtain a gel-like consistency. Varying masses of prepared
The process of encapsulation involves numerous variables that sludge were then added to the polymer solution and mixed thor-
can affect the adsorptive performance of the adsorbent. It is there- oughly to achieve a homogeneous mixture. The suspension was
fore important to select a suitable experimental technique that will subsequently injected dropwise into a calcium chloride (CaCl2 ,
evaluate the effects of significant variables along with possible in- anhydrous granular, LR, Som Datt Finance Corporation, New
teractions (Bhunia and Ghangrekar 2008). Response surface meth- Delhi, India) bath using a 30-mL syringe. After gelation, the beads
odology (RSM) and artificial neural network (ANN) have both were rinsed three times and kept in a double distilled water bath. All
been used to optimize and model processes such as adsorption experiments were carried out in duplicate.
(Shojaeimehr et al. 2014), fermentation (Desai et al. 2008), and air Detailed screening experiments were carried out to determine
drying (Karimi et al. 2012). the most optimal range of adsorbent, sodium alginate, and CaCl2
RSM is a collection of mathematical and statistical techniques for encapsulation. The selected ranges were 0.5%–5% (w/v)
that use specific experimental designs to develop mathematical adsorbent, equivalent to 0.125–1.25 g of adsorbent per 25 mL
models to improve and optimize processes (Yeniay 2014). Accord- of distilled water, 0.5%–5% (w/v) sodium alginate, equivalent
ing to Ferreira et al. (2007), the Box-Behnken design is most to 0.125–1.25 g of sodium alginate per 25 mL of distilled water,

© ASCE 04019019-2 J. Environ. Eng.

J. Environ. Eng., 2019, 145(5): 04019019


and 0.5–5% (w/v) CaCl2 , equivalent to 0.125–1.25 g of CaCl2 Adsorption Yield
per 25 mL of distilled water. Similar ranges for sodium alginate The adsorption yield was calculated using Eq. (1)
and CaCl2 were suggested for encapsulation by cross linking
Co − Ct
[A. Bezbaruah, T. B. Almeelbi, and M. Quamme, US Patent %Adsorption ¼ × 100 ð1Þ
No. 15/147,437 (2014)]. Other researchers including Ociński et al. Co
(2016), Derakhshandeh et al. (2012), and Mahmood et al. (2015)
where Co = initial adsorbate concentration in solution (mg=L); and
also experimented successfully within these ranges.
Ct = adsorbate concentration in solution (mg=L) at time t.
Preparation of Adsorbate Adsorption Capacity
Phosphate stock solution was prepared by dissolving preweighed The adsorption capacity was determined using Eq. (2)
amounts of potassium dihydrogen phosphate (KH2 PO4 , J.T Baker,
Mexico) in double distilled water. ðCo − Ce Þ
qe ¼ ×V ð2Þ
M
Characterization of Adsorbent
The morphological structure of the adsorbent was characterized where qe = mg of adsorbate adsorbed/g of adsorbent at equilibrium
Downloaded from ascelibrary.org by 41.102.31.224 on 08/24/21. Copyright ASCE. For personal use only; all rights reserved.

using a Hitachi S-3000N scanning electron microscope (SEM) (mg=g); Ce = equilibrium adsorbate concentration (mg=L); V =
(Hitachi, Tokyo), and the elemental composition was determined volume of adsorbate solution (L); and M = mass of adsorbent (g).
using an energy dispersive spectroscopy (EDS) analyzer (IXRF
Systems, Austin, Texas) at a voltage of 20 kV. The particle size Kinetic Models
was determined using a Vernier caliper and calculated as the aver-
age value of the diameter of 100 beads. Lagergren Model
Lagergren [as cited in Ho and McKay (1998a)], developed a first-
order rate equation to describe the kinetic process of oxalic acid and
Adsorption Studies
malonic acid onto charcoal. Ho and McKay (Ho and McKay
1998a), described the equation as pseudo-first order
Kinetic Studies
Batch studies were conducted using the parallel method according qt ¼ qe ð1 − exp−KPFO t Þ ð3Þ
to the EPA Office of Prevention, Pesticides and Toxic Substances
(OPPTS) method 835.1230 (USEPA 2008). Experiments were car- where K PFO = pseudo-first order rate constant (h−1 ); and qt = mg of
ried out in duplicate with 0.5 g of sludge encapsulated in calcium adsorbate adsorbed/g of adsorbent at time, t (mg=g).
alginate beads and spiked with 100 mL of synthetic phosphate sol- Pseudo-Second Order Model
ution (10 mg=L). Adsorbent masses were accurate to 0.001 g The pseudo-second order equation was developed for the adsorp-
and solution volumes to 0.5 mL. The effect of pH was studied tion of divalent metal ions onto peat moss (Ho and McKay 1998b).
in the range 4–10 and kept constant using appropriate amounts The model is based on pseudo-second order chemical reaction
of acetate buffers and sodium bicarbonate buffers and measured kinetics (Ho and McKay 1999) and is expressed as
using a HACH multiparameter meter (HQ430d, HACH, Loveland,
Colorado). Identical reaction mixtures were prepared for each time K PSO q2e t
qt ¼ ð4Þ
interval, agitated to maintain complete mixed conditions on a 1 þ K PSO qe t
mechanical shaker (SK-330-Pro Orbital Digital Shaker, Scilogex,
Rocky Hill, Connecticut) and removed at predetermined time in- h ¼ ðK PSO Þq2e ð5Þ
tervals (Weber and Miller 1988). The adsorbent was separated
by vacuum filtration using a Buchner funnel and Whatman No. where K PSO = pseudo-second order rate constant (g=mg · h); and
3 qualitative filter paper. The concentration of phosphate in the fil- h = initial rate of adsorption (mg=g · h).
trate was estimated by the molybdate/ascorbic acid method with
single reagent (Method 365.2) using an ultraviolet spectrophotom- Intraparticle Diffusion Model
eter (UV-1800 Recording Spectrophotometer, Shimadzu Scientific The intraparticle diffusion model (Weber and Morris 1963) as-
Instruments, Toyko). sumes that the rate of intraparticle diffusion varies proportionally
with the half power of time. The model has the following form:
Equilibrium Studies
qt ¼ K id ðt1=2 Þ ð6Þ
The effect of initial phosphate concentration was studied by equili-
brating 0.5 g of lime-iron sludge encapsulated in calcium alginate where K id = rate constant of intraparticle transport (mg=g · h1=2 ).
beads and 100 mL of synthetic phosphate solution of varying con-
centrations (10–100 mg=L). The effect of temperature was exam- Diffusion-Chemisorption Model
ined by agitating reaction mixtures in a shaking water bath (Julabo The diffusion-chemisorption model (Sutherland and Venkobachar
SW23, Julabo GmbH, Seelbach, Germany) at temperatures varying 2010) was developed to simulate biosorption of heavy metals onto
from 300  2 K to 328  2 K. Experiments were carried out in heterogeneous media. The model can be represented as follows:
duplicate.
1
qt ¼ 1 1
ð7Þ
Point of Zero Charge þK 0.5
qe DC ×t
The point of zero charge (pHPZC ) was evaluated using the
solid addition method described by Lee et al. (2012). Adsorbents K 2DC
were contacted with 100 mL KNO3 solution (50 mg=L) at various ki ¼ ð8Þ
qe
pH solutions. The graph of ΔpH versus the initial pH was
plotted, and the pHPZC was taken as the intersection of the initial where K DC = diffusion-chemisorption constant (mg=g · h1=2 ); and
pH with ΔpH. ki = initial adsorption (mg=g · h).

© ASCE 04019019-3 J. Environ. Eng.

J. Environ. Eng., 2019, 145(5): 04019019


Mass Transfer Studies where
π 2 De
External Diffusion Model K¼ ð16Þ
r2
The external diffusion model assumes that during the initial stages
of adsorption, intraparticle resistance is negligible and transport is A linear plot of − lnð1−X 2 ðtÞÞ versus t with an intercept of zero
mainly due to film diffusion (Aksu and Isoglu 2006). It was derived would suggest particle diffusion was the controlling step. If film
from an application of Fick’s second law and expresses the concen- diffusion controls the rate of adsorption, the following analogous
tration of solute in the solution as a function of the difference in expression can be used:
concentration of the solute in the solution, C, and at the particle  
surface, Ci , according to the following equation (Jansson-Charrier 3De Ce
XðtÞ ¼ 1 − exp ð17Þ
et al. 1996): rδCs
dq or
¼ −kf So ðC − Ci Þ ð9Þ
dt
− lnð1 − XðtÞÞ ¼ kf t ð18Þ
Downloaded from ascelibrary.org by 41.102.31.224 on 08/24/21. Copyright ASCE. For personal use only; all rights reserved.

where q = average solute concentration in the solid (mg=g);


C = concentration of the solute in the bulk of the liquid (mg=L); where
Ci = concentration of the solute in the liquid at the particle/liquid
3De Ce
interface (mg=L); kf = external mass transfer coefficient (m=s); and kf ¼ ð19Þ
So = surface area for mass transfer (cm−1 ). rδCs
Since Ci approaches zero and C approaches Co , as t → 0, where Cs = equilibrium concentration of adsorbate in the solid
Eq. (9) could be simplified [Eq. (10)] and kf can be determined phase (mg=L); and δ = thickness of liquid film.
from the slope of the curve C=Co versus t A linear plot of − lnð1−XðtÞÞ versus t with an intercept of zero
  would suggest film diffusion was the controlling step.
dðC=Co Þ
¼ −kf So ð10Þ
dt t¼o Particle Diffusion Model
The model assumes that particle diffusion is rate controlling and
Assuming the particles are spherical, the surface area for mass that Vermeulen’s approximation for particle diffusion [Eq. (14)]
transfer, So , can be obtained by the following equation (Furusawa could be simplified to cover most of the data points for calculating
and Smith 1973): effective particle diffusivity as follows (Vermeulen 1953):
6ms π2
So ¼ ð11Þ lnð1 − X 2 ðtÞÞ ¼ De t ð20Þ
dp ρð1 − εp Þ r2
where ms = mass of adsorbent particles per unit volume (g=cm3 ); qt
XðtÞ ¼ ð21Þ
dp = average particle diameter (cm); ρ = adsorbent density (g=cm3 ); qe
and εp = adsorbent porosity.
Biot Number
Homogeneous Particle Diffusion Model The Biot number, Bi, indicates the predominance of internal dif-
The homogeneous particle diffusion model (HPDM) assumes that fusion against external diffusion (Subash and Krishna Prasad 2016)
the rate controlling step of adsorption could be described by either
film or particle diffusion (El-Naggar et al. 2012). Particle diffusion kf r
Bi ¼ ð22Þ
was described by Boyd et al. (1947) as follows: De
 2 2 
6 X∞
1 −Z π De t Equilibrium Models
XðtÞ ¼ 1 − 2 exp ð12Þ
π Z¼1 Z2 r2
Langmuir Isotherm
where XðtÞ = fractional attainment of equilibrium at time, t; De = The Langmuir isotherm (Langmuir 1916) assumes that adsorption
effective diffusion coefficient of adsorbate in the sorbent phase sites on the adsorbent possess an equal affinity for molecules and
(m2 =s); r = radius of the adsorbent particle assumed to be spherical that each site is capable of adsorbing one molecule thereby forming
(m); and z = integer number. a monolayer. The model is expressed as
XðtÞ values could be determined as follows:
qm K L C e
qt qe ¼ ð23Þ
XðtÞ ¼ ð13Þ 1 þ K L Ce
qe
where K L = Langmuir equilibrium constant (L=mg); and qm =
Vermeulen (1953) approximation of Eq. (12) fits the whole maximum adsorption capacity (mg=g).
range 0 < XðtÞ < 1
  2 2 1 Freundlich Isotherm
π D t 2 Firth [as cited in Swan and Urquhart (1927)], explained that the
XðtÞ ¼ 1 − exp − 2 e ð14Þ
r equation of the form x ¼ kc1=n was first applied to adsorption
of gases by De Saussure in 1814. Its application was further ex-
This equation could be further simplified to the following tended to solutions by Boedecker in 1859 (Swan and Urquhart
expression (Lao-Luque et al. 2014): 1927). In 1906, Freundlich described the adsorption isotherm math-
ematically as a special case for nonideal and reversible adsorption
− lnð1 − X 2 ðtÞÞ ¼ 2Kt ð15Þ (Freundlich 1906). This equation is presented as

© ASCE 04019019-4 J. Environ. Eng.

J. Environ. Eng., 2019, 145(5): 04019019


PN P
qe ¼ K F ðCe Þ1=nF ð24Þ 2
i¼1 ððqei Þexp − qeexp;mean Þ − i¼1 ððqei Þexp − ðqei Þpred Þ
N 2
2 PN
R ¼ 2
where K F = Freundlich constant related to adsorption affinity i¼1 ððqei Þexp − ðqei Þpred Þ
[ðmg=gÞ ðL=mgÞ1=n ]; and nF = Freundlich constant related to ð32Þ
heterogeneity.
Redlich-Peterson Isotherm 1X N
½jðqei Þpred − ðqei Þexp j
RPE% ¼  100 ð33Þ
The Redlich-Peterson isotherm (Redlich and Peterson 1959) is a N i¼1 ðqei Þexp
hybrid isotherm that incorporates the features of the Langmuir
and the Freundlich isotherms. It is represented by vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
u N
u1 X
K RP Ce RMSE ¼ t ððqei Þexp − ðqei Þpred Þ2 ð34Þ
qe ¼ ð25Þ N i¼1
1 þ αRP CgeRP
where K RP = Redlich-Peterson equilibrium constant; gRP = sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 2
Redlich-Peterson exponent; and αRP = Redlich-Peterson constant. 1 X ðqei Þexp − ðqei Þpred
MPSD ¼ 100 ð35Þ
Downloaded from ascelibrary.org by 41.102.31.224 on 08/24/21. Copyright ASCE. For personal use only; all rights reserved.

N − P i¼1 ðqei Þexp


Sips Isotherm
The Sips isotherm (Sips 1948) is a combined form of the Langmuir
" #
and Freundlich isotherms developed for predicting heterogeneous 100 X N
ððqei Þexp − ðqei Þpred Þ2
adsorption systems and bypassing the limitation of the rising HYBRID ¼ ð36Þ
N − P i¼1 ðqei Þpred
adsorbate concentration associated with Freundlich isotherm
model. It is expressed as !
N h i2
qS ðαS Ce Þ
nS 1X
qe ¼ ð26Þ MSE ¼ ðqei Þexp − ðqei Þpred ð37Þ
1 þ ðαS Ce ÞnS N i¼1

where qS = Sips maximum adsorption capacity (mg=g); αs = Sips X


N ððq Þ 2
ei pred − ðqei Þexp Þ
affinity constant; and ns = Sips index of heterogeneity. χ2 ¼ ð38Þ
i¼1
ðqei Þpred
Thermodynamic Equations where N = number of experimental points; and P = number of
Gibbs free energy change ΔG° is given by the following equation parameters in the model.
(Nollet et al. 2003):
ΔG° ¼ −RTlnK L ð27Þ Experimental Design and Procedure

According to Eq. (28) the slope and intercept obtained from lin-
Response Surface Methodology
ear plots of ΔG° versus temperature, T, represents entropy change,
A three-level three-factor Box-Behnken experimental design in
ΔS°, and enthalpy change, ΔH°, respectively
RSM was generated using Design-Expert version 9.0 software
ΔG° ¼ ΔH° − TΔS° ð28Þ (Minneapolis, Minnesota) to determine the optimum conditions for
encapsulation of lime-iron sludge. Similar designs were used by
Activation energy Ea and sticking probability S can be esti- Chen et al. (2016) for optimizing microencapsulation of Bifidobac-
mated using a modified Arrhenius-type equation related to surface terium bifidum BB01, Narsaiah et al. (2014), for optimizing micro-
coverage as encapsulation of nisin with sodium alginate and guar gum and Ismail
and Madhavan Nampoothiri (2010) for optimizing the encapsulation
S ¼ ð1 − θÞ expð−Ea =RTÞ ð29Þ of probiotic Lactobacillus plantarum in calcium alginate beads. A
total of 17 runs were carried out to optimize the chosen variables,
The linearized form is given as
namely, sodium alginate dose, CaCl2 dose, and lime-iron sludge
lnð1 − θÞ ¼ ln S þ Ea =RT ð30Þ dose. For statistical computations, the three independent variables
were denoted as X1, X2, and X3, respectively. The ranges as deter-
where θ is the surface coverage expressed as mined by screening experiments and the levels of each variable are
given in Table 1. Phosphate adsorption capacity was taken as the re-
θ ¼ ð1 − Ce =Co Þ ð31Þ
sponse of the system. The optimum parameter values were obtained
According to Eq. (30), S and Ea can be determined from a plot by numeric optimization and by analyzing the response surface
of ln ð1 − θÞ versus 1=T (Nollet et al. 2003). 3D plots.

Error Analysis Table 1. Coded and actual values of variables for the Box-Behnken design
The goodness of fit by the various kinetic, isotherm, mass transfer, Coded levels
and predictive models to the experimental data were evaluated of variables
using the coefficient of determination (R2 ), the Relative Percent Variable w/v ratio (%) Code −1 0 1
Error (RPE), the Root Mean Square Error (RMSE), Marquardt’s
Percent Standard Deviation (MPSD), the Hybrid Error Function Sodium alginate (g=25 mL) 1–2 X1 0.25 0.38 0.5
(HYBRID), the Mean Square Error (MSE), and the Chi-square CaCl2 (g=25 mL) 1–2 X2 0.25 0.38 0.5
Lime-iron sludge (g=25 mL) 1–2 X3 0.25 0.38 0.5
function (χ2 ) presented as follows:

© ASCE 04019019-5 J. Environ. Eng.

J. Environ. Eng., 2019, 145(5): 04019019


Pareto Analysis Additionally, beads were left in the CaCl2 for 30 min after which
The Pareto analysis was carried out to check the percentage effect they were rinsed and stored in a distilled water bath. Beads isolated
(Pi ) of each variable on the response (Khuri and Cornell 1996) from CaCl2 in less than 15 min were not completely gelled. Longer
 2  curing time significantly improved the gelation and reduced bead
β diameter. However, beads cured for more than 24 h were easily
Pi ¼ P i 2 × 100 ði ≠ 0Þ ð39Þ
βi disintegrated when agitated.
where β i = regression coefficient of individual process variables. Determination of Optimum Range of Lime-Iron Sludge
for Encapsulation
Artificial Neural Network Increase of sludge content in calcium alginate beads from 0.125 to
The neural network toolbox of MATLAB version 7.14.0 (R2012a) 0.5 g resulted in 91% increase in adsorption capacity due to the
was used to develop a three-layer feed forward ANN model for increased number of available adsorption sites. However, increase
predicting phosphate adsorption capacity. Sodium alginate dose, in sludge dose beyond 0.5 g resulted in only marginal increments in
CaCl2 dose, and lime-iron sludge dose were used as inputs to adsorption (Fig. 1). This behavior may be due to the binding of
the network, and the corresponding adsorption capacity was used almost all phosphate anions to the sludge and the establishment
as the target. The data were first normalized in the range −1 to 1
Downloaded from ascelibrary.org by 41.102.31.224 on 08/24/21. Copyright ASCE. For personal use only; all rights reserved.

of equilibrium. Similar results were reported by Rezaei et al.


using the following equation: (2017) for dye adsorption using Fe3 O4 nanoparticles encapsulated
  in calcium alginate beads and Fiol et al. (2004) for chromium
X i − X min
Xnorm ¼ 2 −1 ð40Þ uptake using grape stalk waste encapsulated in calcium alginate
X max − X min beads.
where X i = input or output variable X; X min = minimum value of Determination of Optimum Range of CaCl2
variable X; and X max = maximum value of variable X. for Encapsulation of Lime-Iron Sludge
The data set was divided whereby 70% of the data was applied Increase in CaCl2 dose from 0.125 to 0.38 g resulted in 151%
to training the network, 15% for validation, and 15% for testing the increase in adsorption. This increase may be attributed to the in-
accuracy of the model and its prediction. The tangent sigmoid creased porosity of the beads. According to Barde (2015) the pore
transfer function (tansig) [Eq. (41)] and the linear transfer function size of calcium alginate beads depends on the amount of calcium
(purelin) [Eq. (42)] were used at the hidden and outer layers, ions present. The authors found that increasing the concentrations
respectively. of CaCl2 resulted in increased pore diameter. However, it was
Tansig observed that increase in CaCl2 beyond 0.5 g had an insignificant
fðnÞ ¼ ½2=ð1 þ expð−2  nÞÞ − 1 ð41Þ effect on the adsorption capacity (Fig. 2). This may be attributed to
the possible saturation of calcium binding sites in the guluronic
Purelin acid chains of the sodium alginate as a result of which no additional
cross linking could occur with higher CaCl2 dosages (El-Kamel
fðnÞ ¼ n ð42Þ et al. 2003).
Determination of Optimum Range of Sodium Alginate
Relative Importance Index for Encapsulation of Lime-Iron Sludge
The relative importance of the input variables on the adsorption Uniform-sized spherical beads were produced using 0.25–0.5 g of
capacity was determined by sensitivity analysis using the weight sodium alginate [Figs. 3(c and d)]. Sodium alginate doses less than
method proposed by Garson (Garson 1991) and is presented as 0.25 g resulted in flattened beads [Figs. 3(a and b)]. According to
Eq. (43) Senuma et al. (2000) this was due to the impact of the less viscous
Pm¼Nh PNi droplets against the surface of the CaCl2 bath. Doses more than
m¼1 ððjW jm j= k¼1 jW km jÞ × jW mn jÞ
ih ih ho
Ij ¼ Pk¼Ni Pm¼Nh PNi ð43Þ 0.5 g [Figs. 3(e and f)] resulted in larger deformed beads due to
k¼1 f m¼1 ðjW km j= k¼1 jW km jÞ × jW mn jg
ih ih ho
its high viscosity and denser matrix structure (Fundueanu et al.
1999). The selected parameter ranges for encapsulation are pre-
where I j = relative importance of the jth input variable on the out- sented in Table 2.
put variable; N i = number of input neurons; N h = number of hidden
neurons; and W = connection weight. The superscripts ‘i’, ‘h’, and
‘o’ refer to input, hidden, and output layers, respectively, and sub- 2
scripts ‘k’, ‘m’, and ‘n’ refers to input, hidden, and output neurons, 1.8
respectively. 1.6
1.4
qe (mg/g)

1.2
Results and Discussion 1
0.8
Protocol for Encapsulation of Lime-Iron Sludge Beads 0.6
Preliminary experiments were conducted to determine suitable 0.4
ranges for different variables to optimize the encapsulation of lime- 0.2
iron sludge. In all instances, the syringe was held approximately 0
0 0.2 0.4 0.6 0.8
6.0 cm above the CaCl2 bath as it was observed that the beads flat-
tened on contact with the CaCl2 if the syringe was held less than Lime-iron sludge dose (g)
6.0 cm above the bath. This may be attributed to the inability of
Fig. 1. Effect of lime-iron sludge content of encapsulated calcium
the droplet viscosity and surface tension forces to overcome the
alginate beads on phosphate adsorption capacity.
surface tension exerted by the CaCl2 solution (Chan et al. 2006).

© ASCE 04019019-6 J. Environ. Eng.

J. Environ. Eng., 2019, 145(5): 04019019


2
1.8 Y ¼ 1.70078 − 0.064994X 1 þ 0.0365795X 2 þ 0.105731X 3
1.6 þ 0.0285035X 1 X 2 þ 0.013836X 1 X 3 − 0.019121X 2 X 3
1.4
− 0.0162616X 21 − 0.0105021X 22 − 0.052366X 23 ð44Þ
qe (mg/g)

1.2
1
where Y is the adsorption capacity and X 1 , X 2 , and X 3 represent
0.8
the sodium alginate, CaCl2 , and lime-iron sludge dose, respec-
0.6 tively. The accuracy of the model was assessed by comparing
0.4 the predicted qe to the experimental qe . The results presented
0.2 in Fig. 4 reveal a significantly high correlation (R2 ¼ 0.9985)
0 and highlight the accuracy of the RSM prediction.
0 0.5 1 1.5 The analysis of variance (ANOVA) was conducted to test the
CaCl2 (g) significance and adequacy of the quadratic equation, and the results
are presented in Table 4. The calculated F-value (502.30) was
Fig. 2. Effect of CaCl2 content of encapsulated calcium alginate beads
greater than the tabulated F-value (3.68) at 95% significance
Downloaded from ascelibrary.org by 41.102.31.224 on 08/24/21. Copyright ASCE. For personal use only; all rights reserved.

on phosphate adsorption capacity.


and therefore confirms the adequacy of the model fits. The p-value
was < 0.0001, which confirms that the model was statistically
significant. Additionally, all the model terms were statistically
Determination of Optimum Combination of Parameters significant (p < 0.05) at the 95% confidence interval.
for Encapsulation The percentage effect of each independent variable on the ad-
The optimum ranges of individual parameters were used to con- sorption capacity was determined by Pareto analysis using Eq. (39).
struct a Box-Behnken design in RSM to determine the optimum The results shown in Fig. 5 indicate that sludge dose was the most
combination of parameters for lime-iron sludge encapsulation. influential parameter in the process, followed by sodium alginate,
The impact of parameter variation on the adsorption capacity and finally CaCl2 dose.
for each run performed according to the Box-Behnken design
is given in Table 3. By applying multiple regression analysis, Effect of Varying Bead Encapsulation Parameters
the following quadratic polynomial equation in coded terms was Three-dimensional (3D) surface plots were analyzed to gain insight
obtained: into the main and interactive effects of the independent variables

Fig. 3. Images of alginate beads formed using varying masses of sodium alginate: (a) 0.125 g; (b) 0.2 g; (c) 0.25 g; (d) 0.5 g; (e) 0.75 g; and (f) 1.0 g.

© ASCE 04019019-7 J. Environ. Eng.

J. Environ. Eng., 2019, 145(5): 04019019


Table 2. Parameter ranges used for encapsulation in CaCl2 concentration increases the porosity of the beads. Analysis
Parameter Range of the plots corroborates that lime-iron sludge was the most influ-
ential variable, which is in agreement with the results of the Pareto
Height of syringe above CaCl2 bath 6 cm
analysis.
Gelation time 30 mins
Sludge dose (=25 mL) 0.25–0.5 g Artificial Neural Network
CaCl2 dose (=25 mL) 0.25–0.5 g
A three-layer feedforward ANN was developed to predict the ad-
Sodium alginate dose (=25 mL) 0.25–0.5 g
sorption capacity of the sludge encapsulated beads. The number of
hidden neurons was varied from 1 to 15 and its impact on perfor-
mance assessed using the MSE. Using the Levenberg–Marquardt
algorithm to train the network with the tansig and purelin transfer
Table 3. Box-Behnken experimental and predicted adsorption capacities functions at the hidden and output layer, respectively, a minimum
Coded values qe (mg=g) MSE value of 0.000245 was obtained using three hidden neurons.
A comparison of the ANN predicted and the experimental qe
Run X1 X2 X3 Experimental Predicted (Fig. 7), reveals a high correlation (R2 ¼ 0.9405).
Downloaded from ascelibrary.org by 41.102.31.224 on 08/24/21. Copyright ASCE. For personal use only; all rights reserved.

1 −1 0 −1 1.6060 1.6053
2 1 0 −1 1.4506 1.4477 ANN Empirical Equation
3 0 −1 −1 1.4791 1.4765 An empirical equation correlating the inputs was developed to pre-
4 0 0 0 1.7057 1.7008 dict adsorption capacity without having to run the ANN software
5 −1 −1 0 1.7276 1.7309 (Shahryari et al. 2013). The equation derived using the weights
6 0 −1 1 1.7326 1.7262 (W i ) and biases (bi ) of the optimized network (Table 5) is presented
7 1 0 1 1.6860 1.6868 as follows:
8 1 −1 0 1.5383 1.5440
9 0 0 0 1.6981 1.7008 qt pred ¼ 0.17028F1 − 0.5279F2 þ 0.37552F3 þ 0.4983 ð45Þ
10 0 0 0 1.6990 1.7008
11 0 1 −1 1.5815 1.5879 where Fi is the tansig activation function used in the hidden layer
12 1 1 0 1.6775 1.6742 and is given as
13 0 1 1 1.7585 1.7611
14 0 0 0 1.6998 1.7008 2
15 −1 1 0 1.7528 1.7471 Fi ¼ − 1; i ¼ 1∶3 ð46Þ
½1 þ expð−2  EiÞ
16 0 0 0 1.7012 1.7008
17 −1 0 1 1.7860 1.7891 Ei is the weighted sum of the input, I, defined as

Ei ¼ W i1 × Sodium alginate þ W i2 × CaCl2


þ W i3 × Sludge dose þ bi ð47Þ
1.9
1.8 R2 = 0.9985
ANN Relative Importance Index
1.8
Predicted qe (mg/g)

The relative importance of the input variables on the adsorption


1.7 capacity was determined using the Garson’s equation [Eq. (43)].
1.7
Lime-iron sludge dose was found to be the most influential param-
eter with a relative importance of 35.38%. This was followed by
1.6 CaCl2 (34.31%) and finally sodium alginate dose (30.30%).
1.6
1.5
Comparison of RSM and ANN
The predictive accuracy of the RSM and ANN models was assessed
1.5
using R2 , χ2 , RMSE, and RPE functions. The results presented in
1.4 Table 6 indicate that the RSM model produced a higher R2
1.4 1.5 1.6 1.7 1.8 1.9
and significantly lower χ2, RMSE, and RPE and was, therefore,
Experimental qe (mg/g) superior to the ANN model.
Fig. 4. Comparison of experimental and RSM model predictions. Optimization Analysis
Numerical optimization in RSM was used to determine the opti-
mum parameters for bead preparation. The results revealed that
maximum adsorption could be achieved using beads prepared with
on phosphate adsorption capacity. Figs. 6(a and b) show that an 0.25 g of sodium alginate and 0.5 g of lime-iron sludge in 25 mL
increase in sludge dose resulted in a significant increase in adsorp- of distilled water to produce a homogeneous mixture and added
tion. This increase may be attributed to the availability of more sur- dropwise into a solution of 0.31 g CaCl2 in 25 mL of distilled water.
face area and accessibility to more adsorption sites (Ociński et al. Laboratory experiments were subsequently conducted to validate
2016). An increase in sodium alginate dose resulted in a reduction these findings. The resulting adsorption capacity was 1.8285 mg=g
in adsorption capacity [Figs. 6(b and c)]. This may be due to the producing a residual error of 2.9%, and thus confirms the validity of
high viscosity of the beads, which decreases the pore size, and con- the model.
sequently, impedes the diffusion of adsorbate into the alginate ma-
trix (Singh et al. 2012). Adinarayana et al. (2004) added that an
increase in sodium alginate dose also reduces the porosity of the Characterization of Adsorbent
beads. Increase in CaCl2 concentration increased adsorption capac- Optimization of the encapsulation process produced beads that
ity [Figs. 6(a and c)]. According to Saha and Ray (2013), increase were spherical with the sludge dispersed throughout the calcium

© ASCE 04019019-8 J. Environ. Eng.

J. Environ. Eng., 2019, 145(5): 04019019


Table 4. ANOVA results for response surface quadratic model
Source Sum of squares Degrees of freedom Mean square F value p-value Prob > F
Model 0.1500 9 0.0170 502.3000 <0.0001
A-Sodium alginate 0.0340 1 0.0340 996.2400 <0.0001
B-Calcium chloride 0.0110 1 0.0110 315.5700 <0.0001
C-Lime-iron sludge 0.0890 1 0.0890 2636.4500 <0.0001
AB 0.0033 1 0.0033 95.8000 <0.0001
AC 0.0008 1 0.0008 22.5700 0.0021
BC 0.0015 1 0.0015 43.1100 0.0003
A2 0.0001 1 0.0011 32.8200 0.0007
B2 0.0005 1 0.0005 13.6900 0.0077
C2 0.0120 1 0.0120 340.3800 <0.0001
Residual 0.0002 7 0.0000 — —
Pure error 0.0000 4 0.0000 — —
Cor total 0.1500 16 — — —
Downloaded from ascelibrary.org by 41.102.31.224 on 08/24/21. Copyright ASCE. For personal use only; all rights reserved.

Std. dev. 0.0058


Mean 1.6600
C.V. % 0.3500
PRESS 0.0033
R-squared 0.9985
Adj R-squared 0.9965
Pred R-squared 0.9786
Adeq precision 76.4390

60 Chen et al. (2007), the presence of Fe-OH functional groups


may result in adsorption of phosphate via ligand exchange. The
50 authors further explained that phosphate removal by Ca compo-
Percentage (%)

40 nents might have occurred by Ca orthophosphate precipitation


resulting in CaHPO4 . The EDS exposes the presence of small
30 concentrations of phosphate before adsorption. The increased
phosphate peak after adsorption confirms that phosphate was suc-
20 cessfully adsorbed. The high presence of oxygen shown presumes
its occurrence in the oxide and oxo-hydroxide form (Ociński et al.
10
2016).
0
X3 X1 X3X3 X2 X1X2 X2X3 X1X1 X1X3 X2X2
Kinetic Modeling
Factor

Fig. 5. Pareto analysis in coded terms. Modeling the Effects of Agitation


Kinetic data were fitted to four models: Lagergren model; pseudo-
second order model; Weber and Morris intraparticle diffusion
model, and the diffusion-chemisorption model. The goodness of
alginate matrix and had an average diameter of 4.0  1.0 mm fit was assessed using the RPE, MPSD, and HYBRID error func-
[Fig. 8(a)]. SEM images illustrated in Figs. 8(b–d) were used to tions presented as Eqs. (33), (35), and (36), respectively. The results
examine the surface morphology of the beads. SEM image of of nonlinear regression presented in Table 7 show the pseudo-
the lime-iron sludge prior to encapsulation revealed a rough surface second order model produced the best simulation of the data.
and particles appeared amorphous, with irregular flocs. After en- According to the pseudo-second order model, the increase in
capsulation, there was a reduction in roughness; however, an abun- agitation speed resulted in an increase in both the initial and overall
dant amount of protuberance was observed that according to Soni adsorption rate. This may have occurred because increased agita-
et al. (2012) may have resulted due to the agglomeration of the tion reduces the thickness of the liquid boundary layer and as a
encapsulated material. The SEM image after adsorption appears consequence reduces the mass transfer resistance. Additionally,
stable in morphology. The EDS was used to assess the change it increases the turbulence of the bulk liquid and therefore promotes
in elemental composition before and after encapsulation as well greater collision of adsorbate onto adsorption sites (Patil et al.
as before and after adsorption. The spectrum [Fig. 9(a)] reveals 2011).
that Fe, Ca, and O were the dominant elements present before For all agitation speeds tested, adsorption occurred rapidly and
encapsulation. Following encapsulation Fe, Ca, C, and O were then slowed as equilibrium approached. This may be attributed to
the dominant elements present both before and after adsorption the availability of a significant number of unoccupied adsorption
[Figs. 9(b and c)]. It was noted that after encapsulation the Fe sites early in the reaction, however, after a certain time, the avail-
and Ca peaks (particularly the Fe peak) diminished significantly. able sites become occupied by adsorbate and consequently create a
This was attributed to the finite mass of sludge (approximately repulsive force between the adsorbate on the adsorbent surface and
0.002 g per bead) encapsulated. A further significant reduction in bulk solution (Banerjee and Chattopadhyaya 2013). Equilibrium
in the Fe and Ca peaks was observed following adsorption was reached after 16 h of agitation. Makris et al. (2005) explained
[Fig. 9(c)]. These findings, therefore, suggest that Fe and Ca that relatively slow phosphate uptake might be due to intraparticle
were the main components for phosphate removal. According to diffusion within the micropores of an adsorbent.

© ASCE 04019019-9 J. Environ. Eng.

J. Environ. Eng., 2019, 145(5): 04019019


Design-Expert® Software Design-Expert® Software

qe qe
1.786 1.786

1.4506 1.4506
1.79 1.77
X1 = A: Sodium Alginate X1 = C: Lime-iron Sludge
X2 = C: Lime-iron Sludge X2 = B: Calcium Chloride
1.7025 1.695
Actual Factor Actual Factor
B: Calcium Chloride = 0.38 A: Sodium Alginate = 0.38
1.615 1.62

qe
qe

1.5275 1.545

1.44 1.47

0.5 0.50 0.50 0.50


0.4375 0.44 0.44 0.44
0.375 0.38 0.38 0.38
Downloaded from ascelibrary.org by 41.102.31.224 on 08/24/21. Copyright ASCE. For personal use only; all rights reserved.

0.3125 0.31 0.31 0.31


C: Lime-iron Sludge A: Sodium Alginate
0.25 0.25 B: Calcium Chloride 0.25 0.25 C: Lime-iron Sludge

(a) (b)

Design-Expert® Software

qe
1.786

1.4506
1.76
X1 = A: Sodium Alginate
X2 = B: Calcium Chloride
1.7025
Actual Factor
C: Lime-iron Sludge = 0.38
1.645
qe

1.5875

1.53

0.50 0.50
0.44 0.44
0.38 0.38
0.31 0.31
B: Calcium Chloride A: Sodium Alginate
0.25 0.25
(c)

Fig. 6. Surface plots for phosphate adsorption at (a) constant CaCl2 ; (b) constant sodium alginate; and (c) constant lime-iron sludge.

1.9 Table 6. Comparison of predictive accuracy of RSM and ANN models


R2 = 0.9405
1.85
Model
Predicted qe (mg/g)

1.8
1.75 Parameter RSM ANN
1.7 R2 0.9985 0.9405
1.65 χ2 0.0001 0.0079
1.6 RMSE 0.0037 0.0273
1.55 RPE 0.1930 0.9611
1.5
1.45
1.4
1.4 1.5 1.6 1.7 1.8 1.9 Effect of pH on Adsorption
Experimental qe (mg/g) The effect of solution pH was studied at an initial phosphate con-
centration of 10 mg=L and is presented in Fig. 10. Adsorption
Fig. 7. Comparison of experimental and ANN predictions.
capacity increased with increase in pH and reached a maximum
at a value of pH 7.5. Further increase in pH resulted in a significant
decrease in adsorption. Fig. 11 shows the pHPZC to occur at pH 8.9.
Table 5. ANN weight and bias values
Despite the surface being positively charged below the pHPZC , ad-
sorption decreased significantly with decreasing pH. This may be
I W i1 W i2 W i3 bi attributed to the presence of different phosphate species. At pH 3
1 −1.5598 4.0395 0.51599 −2.6509 the dominant species is H3 PO4 , however, as pH increases H2 PO− 4
2 2.3435 −0.85485 −4.4497 −2.014 and HPO2− 4 become the dominating, both of which are more easily
3 −0.9159 −1.5009 0.44498 −2.4417
adsorbed according to Liu et al. (2002). At pH 12, PO3− 4 is the

© ASCE 04019019-10 J. Environ. Eng.

J. Environ. Eng., 2019, 145(5): 04019019


Downloaded from ascelibrary.org by 41.102.31.224 on 08/24/21. Copyright ASCE. For personal use only; all rights reserved.

Fig. 8. (a) Diameter of encapsulated lime-iron sludge bead; (b) SEM of lime-iron sludge before encapsulation at 1,000×; (c) SEM before adsorption
of phosphate at 500×; and (d) SEM after adsorption of phosphate at 500×.

dominant species and phosphate adsorption reduces possibly due to of the adsorbate. The Sips index of heterogeneity, nS , deviated
competition between PO3− −
4 and OH for active sites on the adsorb- widely from unity indicating some degree of heterogeneity of the
ent. Additionally, the negatively charged carboxylic groups in cal- surface of the adsorbent.
cium alginate may impede the diffusion of anions into the beads, According to the Langmuir isotherm, lime-iron sludge encap-
therefore contributing to the decrease in adsorption capacity at pH sulated beads produced a maximum phosphate adsorption capacity
above the pHPZC (Ociński et al. 2016). of 8.31 mg=g. This was compared to that of other adsorbents re-
ported in the literature (Table 9). The reduction in adsorption
capacity when compared to that of powdered lime-iron sludge,
Equilibrium Modeling 15.30 mg=g, as reported by Chittoo and Sutherland (2017) may
be attributed to the presence of the polymer matrix that creates
Nonlinear Regression of Isotherm Data a diffusional barrier to adsorption sites on the lime-iron sludge.
Equilibrium data were fitted to the two-parameter Langmuir However, the presence of the polymer matrix provides beneficial
and Freundlich isotherms as well as the three-parameter Sips and properties that allow its recovery and reuse. Additionally, its
Redlich-Peterson isotherms. The goodness of fit was assessed adsorptive performance compares well to several adsorbents re-
using the RPE, MPSD, and HYBRID error functions presented ported in the literature; this further suggests its effectiveness as
as Eqs. (33), (35), and (36), respectively. The results of nonlinear an adsorbent.
regression (Table 8) showed the Langmuir isotherm was the best
performing two-parameter model. However, beyond room temper- Effect of Initial Phosphate Concentration on Adsorption
ature, a more robust simulation was observed by the three- The effect of initial phosphate concentration was studied at room
parameter Sips isotherm. It is postulated that at room temperature temperature using a constant sludge dose of 0.5 g. The results pre-
there was no transmigration of the adsorbate in the plane of the sented in Fig. 12 reveal a decrease in removal from 90% to 61%
surface (Foo and Hameed 2010). At higher temperatures, the in- as the concentration increased from 8 to 105 mg=L. At lower con-
creased kinetic energy of the adsorbate anions increases its colli- centrations, most phosphate anions could interact with an adsorp-
sion onto adsorption sites and may have influenced transmigration tion site on the adsorbent. However, as adsorbate concentration

© ASCE 04019019-11 J. Environ. Eng.

J. Environ. Eng., 2019, 145(5): 04019019


Downloaded from ascelibrary.org by 41.102.31.224 on 08/24/21. Copyright ASCE. For personal use only; all rights reserved.

Fig. 9. EDS of (a) lime-iron sludge before encapsulation; (b) encapsulated lime-iron sludge before adsorption of phosphate; and (c) encapsulated
lime-iron sludge after adsorption of phosphate.

Table 7. Analysis of kinetic models using nonlinear regression


Constants Error functions
RPM Model K PSO h RPE MPSD HYBRID
150 Pseudo-first order — — 4.1135 6.5949 0.4555
Pseudo-second order 0.2593 1.2468 1.1841 1.971 0.0511
Intraparticle diffusion — — 18.0584 24.3061 8.1625
Diffusion-chemisorption — — 5.953 8.9449 0.8511
250 Pseudo-first order — — 4.5018 7.3306 0.5714
Pseudo-second order 0.2776 1.3879 1.7276 2.5991 0.0842
Intraparticle diffusion — — 19.7641 26.4275 9.6994
Diffusion-chemisorption — — 5.1932 7.6453 0.7079
350 Pseudo-first order — — 5.1326 7.9981 0.8571
Pseudo-second order 0.374 1.9657 1.7431 2.6507 0.0983
Intraparticle diffusion — — 23.2589 31.4809 14.9196
Diffusion-chemisorption — — 3.281 4.7816 0.3588

increases the number of available adsorption sites becomes rela- generate a greater driving force sufficient enough to overcome
tively insufficient therefore reducing the percentage adsorption. the mass transfer resistance between the solid and liquid phases.
The results also show an increase in uptake capacity as the Hence the unit mass saturation of the adsorbent will increase with
phosphate concentration increased. Steeper concentration gradients higher initial concentration (Banerjee and Chattopadhyaya 2013).

© ASCE 04019019-12 J. Environ. Eng.

J. Environ. Eng., 2019, 145(5): 04019019


100
and the results are presented in Table 10. ΔG° values between 0
90
and −20 kJ=mol indicate that the adsorption process is controlled
Percentage Adsorption 80
by physisorption, whereas values within the range of −80 to
70
−400 kJ=mol indicate that the process is controlled by chemisorp-
60
tion (Yu et al. 2001). In this study ΔG° varied from −8.0847 to
50
−10.5722 kJ=mol indicating that physisorption was the dominant
40
mechanism. For all temperatures studied, ΔG° was negative, indi-
30 cating that the process was spontaneous. The positive value of ΔH°
20 suggests that the process was endothermic with the presence of an
10 energy barrier. According to Fu et al. (2012), ΔH° for Van der
0 Waals interactions, hydrogen bond, ligand exchange, dipole inter-
0 2 4 6 8 10 12 14
action, and chemical bond is 4–10, 2–40, ≈40, 2–29, and
pH
>60 kJ=mol, respectively. In this study, ΔH° was 19.585 kJ=mol
Fig. 10. Effect of pH on phosphate uptake. and therefore involves hydrogen bonding and dipole interaction.
The positive ΔS° reflects the high affinity of lime-iron sludge beads
Downloaded from ascelibrary.org by 41.102.31.224 on 08/24/21. Copyright ASCE. For personal use only; all rights reserved.

for phosphate anions. This may be due to the opening of pores


of the alginate matrix, overcoming the activation energy barrier
6 and enhancing the rate of intraparticle diffusion (Aravindhan
5
et al. 2007).
The Ea value was found to be 37.5826 kJ=mol. According to
4 Nollet et al. (2003), low Ea values (5–40 kJ=mol) denote physi-
Final pH- Initial pH

3 sorption and suggest that the energy barrier existing between the
2 reactants is relatively low. The low value of the sticking probability
(S < 1) indicates favorable sticking of adsorbate to the adsorbent
1
and further highlights physisorption as the dominant attachment
0 mechanism.
0 2 4 6 8 10 12
-1
-2 Mechanisms
-3
Initial pH
External Diffusion Model
Fig. 11. pHPZC for encapsulated beads. Kinetic data were first analyzed using the external diffusion model,
and the results are presented in Table 11. A slight increase in kf was
obtained as agitation increased from 150 to 250 rpm. Further in-
crease in agitation resulted in a marginal decrease in kf . Increase
Minimum change in adsorption capacity was obtained for phos- in agitation decreases the external film resistance allowing the
phate concentrations beyond 77 mg=L possibly due to the filling adsorbate to reach the adsorbent more rapidly and thus increases
of most of the adsorptions sites. kf . At higher agitation, the overall rate of adsorbate uptake may
cause blocking of the pores due to adsorbate ions competing for
available adsorption sites and as a result reduces kf (Allen et al.
Thermodynamic Analysis
2005). R2 obtained for all agitation speeds was well below 0.95
The thermodynamic effect of phosphate adsorption was investi- and therefore implies that external mass transfer may not be rate
gated at four different temperatures (300, 308, 318, and 328 K) controlling for the entire reaction.

Table 8. Analysis of isotherm models using nonlinear regression


Error functions
Temperature (K) Model Constants RPE MPSD HYBRID
300.15 Langmuir qm ¼ 8.3116; K L ¼ 0.2688 34.9966 59.9874 69.5136
Freundlich K F ¼ 3.2127; nF ¼ 4.509 49.7344 87.6678 143.1319
Sips αS ¼ 0.3417; qS ¼ 7.4404; nS ¼ 4.6424 61.1378 84.7882 497.5884
Redlich-Peterson αRP ¼ 0.0204; K RP ¼ 1.2433; gRP ¼ 1.1581 84.648 114.0732 601.5185
308.15 Langmuir qm ¼ 10.942; K L ¼ 0.307 35.5465 62.0614 85.5195
Freundlich K F ¼ 4.1639; nF ¼ 4.2207 53.9349 97.0906 196.8808
Sips αS ¼ 0.3445; qS ¼ 9.7485; nS ¼ 3.0483 10.4284 17.6354 8.1528
Redlich-Peterson αRP ¼ 0.0294; K RP ¼ 1.8954; gRP ¼ 1.4851 21.1128 37.7606 30.4237
318.15 Langmuir qm ¼ 14.584; K L ¼ 0.2364 32.5097 64.253 69.7454
Freundlich K F ¼ 3.8719; nF ¼ 2.8553 58.942 117.4239 215.6085
Sips αS ¼ 0.3272; qS ¼ 12.664; nS ¼ 1.984 15.8939 28.6884 20.9149
Redlich-Peterson αRP ¼ 0.0552; K RP ¼ 2.4541; gRP ¼ 1.3151 29.1539 55.9125 55.5898
328.15 Langmuir qm ¼ 17.2209; K L ¼ 0.5074 47.0785 97.5448 145.036
Freundlich K F ¼ 6.0827; nF ¼ 3.007 70.2111 144.1451 319.1363
Sips αS ¼ 0.7765; qS ¼ 14.5938; nS ¼ 2.3109 13.9427 22.6276 14.0724
Redlich-Peterson αRP ¼ 0.2598; K RP ¼ 7.0637; gRP ¼ 1.1576 42.1666 85.1678 113.1317

© ASCE 04019019-13 J. Environ. Eng.

J. Environ. Eng., 2019, 145(5): 04019019


Table 9. Comparison of the phosphate adsorption capacity by various adsorbents reported in the literature
Adsorbent Temperature Adsorption capacity,
Adsorbent dose (g) pH (K) qe (mg=g) Reference
Iron oxide tailings 2 6.7 266 7 Zeng et al. (2004)
Dolomite 10 9.5 266 4.76 Yuan et al. (2015)
Arundo donax reeds 0.1 6.5 269 16.4 Abdelhay et al. (2018)
Lime-iron sludge 0.5 8 298 15.30 Chittoo and Sutherland (2017)
Alginate calcium carbonate composite beads 1.5 10 298 0.62 Mahmood et al. (2015)
Lime-iron sludge encapsulated calcium alginate beads 0.5 7.5 300 8.31 This study

100 14 in the adsorption process. It was noted as agitation speed increased


the plots move further from the origin. The straight lines obtained

Adsorption Capacity (mg/g)


90
12 in the plot of Fig. 14(b) passed through the origin for all agitation
80
Downloaded from ascelibrary.org by 41.102.31.224 on 08/24/21. Copyright ASCE. For personal use only; all rights reserved.

10 speeds tested indicating that intraparticle diffusion is the dominant


70
% Adsorption

transport mechanism.
60 8
50 Particle Diffusion Model
40 6 The predominance of intraparticle diffusion was further assessed
30 4 using the particle diffusion model. The results (Table 11) show an
Percent Adsorption increase in De with increasing agitation. According to Ha et al.
20
Adsorption Capacity 2 (2008), as agitation increases the diffusion barrier would decrease
10
as particles move further apart thereby increasing the diffusion
0 0
0 50 100 150
coefficient. According to Walker and Weatherley (1999), the mag-
Co (mg/L) nitude of the diffusion coefficient is directly related to the nature
of the adsorption process. For physical adsorption, De ranges
Fig. 12. Effects of initial phosphate concentration. from 10−6 to 10−9 m2 =s and for chemical adsorption process
De ranges from 10−9 to 10−17 m2 =s. In this study, De was in the
order 10−8 m2 =s suggesting that physisorption was the dominant
attachment mechanism. R2 values were well above 0.95 indicating
Weber and Morris Intraparticle Diffusion Model that particle diffusion may be the rate controlling step in the
The data were then assessed using the Weber and Morris intrapar- process.
ticle diffusion model to determine if intraparticle diffusion was rate
controlling. The plots of qt versus t1=2 (Fig. 13) reveal two distinct Biot Number
linear phases during the adsorption process. The first linear phase Bi was determined using kf and De values presented in Table 11.
may be attributed to intraparticle diffusion. The second linear phase Crittenden et al. (2012) explained that for Bi values <1.0, external
represents the slowing down of intraparticle diffusion possibly due mass transfer dominates while for Bi > 30, surface diffusion con-
to low solute concentration in the solution (Aksu and Kabasakal trols and for values between 1 and 30, both external and intrapar-
2005). According to Nethajia et al. (2010), the intercept c is propor- ticle mass transfer contribute to the adsorption rate. The results
tional to the boundary layer thickness and the larger the intercept, (Table 11) therefore indicate that both external and intraparticle
the greater is the boundary layer effect. Patil et al. (2011) explained mass transfer contributes to the reaction rate.
that increasing the agitation speed increases turbulence, which
reduces the film boundary thickness. However, in this instance,
Design of Batch Adsorption System from
an increase in intercept c, was observed with increased agitation.
Isotherm Data
Similar observations were reported by McKay et al. (1983) and Sağ
and Aktay (2000). The authors attributed such effects to systems in Laboratory-scale equilibrium studies are used to predict batch ad-
which surface mass transfer was not significant. sorber size and performance. Fig. 15 shows the schematic of a
single-stage batch adsorber with a solution volume of V (L) and
Homogeneous Particle Diffusion Model an initial phosphate concentration, Co (mg=L), which is reduced
Kinetic data were further analyzed using the HPDM to determine to Ct (mg=L) as the reaction proceeds. The phosphate loading
the relative impact of film and intraparticle diffusion. The data on the adsorbent in the reactor of mass M (g), changes from qo
were fitted to the two model equations [Eqs. (15) and (18)] by to qt with increased reaction time. The mass balance for the reactor
plotting−(lnð1-X 2 ðtÞÞ versus t and − ln ð1-XðtÞÞ versus t, respec- is given by the following equation (McKay et al. 1985):
tively. The straight lines obtained in Fig. 14(a) do not pass through
the origin, indicating that film diffusion is not the rate-limiting step VðC0 − Ct Þ ¼ Mðqt − q0 Þ ¼ Mqt ð48Þ

Table 10. Thermodynamic parameters


Temperature (K) K a (L=mol) ΔG° (kJ=mol) ΔS (kJ=mol) ΔH (kJ=mol) Ea (kJ=mol) S
300 25.5280 −8.0847 0.092 19.585 37.5826 1.978 × 10−7
308 29.1558 −8.6406
318 40.7611 −9.8073
328 48.1878 −10.5722

© ASCE 04019019-14 J. Environ. Eng.

J. Environ. Eng., 2019, 145(5): 04019019


Table 11. Mass transfer coefficients and biot numbers
External diffusion Particle diffusion
model model
Agitation
RPM kf (m=s) R2 De (m2 =s) R2 Bi
150 5.08 × 10−4 0.7116 9.3 × 10−8 0.9953 21.8404
250 5.25 × 10−4 0.7316 9.4 × 10−8 0.9911 22.3532
350 4.57 × 10−4 0.7196 1.08 × 10−8 0.9343 16.9263

2.5

2
Downloaded from ascelibrary.org by 41.102.31.224 on 08/24/21. Copyright ASCE. For personal use only; all rights reserved.

qt (mg/g)

1.5 Fig. 15. Design of single-stage batch system.

1
150 RPM - First Slope
350 RPM - First Slope
0.5 60
150 RPM - Second Slope
350 RPM - Second Slope 50 60% Removal
0
70% Removal
0 1 2 3 4 5
40
t1/2 80% Removal

M (g)
30 90% Removal
Fig. 13. Intraparticle diffusion model for various agitation speed.
20

10
The adsorption process at 300 K was best represented by the
Langmuir isotherm. Therefore, the mass balance under equilibrium 0
condition (Ct → Ce and qt → qe ) is arranged as follows: 0 2 4 6 8 10 12
V (L)
M Co − Ce Co − Ce
¼ ¼ ð49Þ
V qe qm K L Ce =ð1 þ K L Ce Þ Fig. 16. Adsorbent mass (M) versus volume of phosphate solution
treated (V).
Fig. 16 illustrates a series of plots of the predicted values of M (g)
versus V (L) for 60%, 70%, 80%, and 90% phosphate ion removal at
the initial concentration of 15 mg=L and 300 K. For example, the to A. Bezbaruah, T. B. Almeelbi, and M. Quamme [US Patent No.
mass of adsorbent required for the 70% phosphate removal from 15/147,437 (2014)], alginate is biodegradable, and so phosphate sa-
aqueous solution was 2.3, 11.5, and 18.5 g, for phosphate solution turated beads can be used directly as fertilizer, without the need to
volumes of 1, 5, and 8 L, respectively. This evaluation becomes rel- desorb or extract the phosphate. Robalds et al. (2016) highlighted
evant for pilot-batch system design as well as large-scale batch ap- the importance of ensuring that the concentrations of heavy metals
plications. Following adsorption saturation, the spent adsorbent may are within safe limits if the spent adsorbent is to be used for land
be considered for agricultural applications as a fertilizer. According application. Elemental analysis before and after adsorption did

4 3
3.5
2.5
3
2.5 2
-ln(1-x2(t))
-ln(1-x(t))

2 1.5
1.5 150 RPM
150 RPM 1
1 250 RPM
250 RPM
0.5 350 RPM
0.5 350 RPM
0 0
0 2 4 6 8 10 12 0 5 10 15
(a) Time (hours) (b) Time (hours)

Fig. 14. HPDM plot for (a) film diffusion; and (b) intraparticle diffusion.

© ASCE 04019019-15 J. Environ. Eng.

J. Environ. Eng., 2019, 145(5): 04019019


not indicate the presence of heavy metals and therefore underscores Aravindhan, R., N. N. Fathima, J. R. Rao, and B. U. Nair. 2007. “Equi-
its potential for agricultural applications. librium and thermodynamic studies on the removal of basic black
dye using calcium alginate beads.” Colloids Surf. A 299 (1): 232–238.
https://doi.org/10.1016/j.colsurfa.2006.11.045.
Conclusion Bahraman, F., and I. Alemzadeh. 2017. “Optimization of L-Asparaginase
Immobilization onto calcium alginate beads.” Chem. Eng. Commun.
A protocol for encapsulating lime-iron sludge in calcium alginate 204 (2): 216–220. https://doi.org/10.1080/00986445.2015.1065821.
beads was successfully developed. RSM was found to be more ac- Banerjee, S., and M. C. Chattopadhyaya. 2013. “Adsorption characteristics
curate than ANN in optimizing the encapsulation process. An em- for the removal of a toxic dye, tartrazine from aqueous solutions by a
low cost agricultural by-product.” Arabian J. Chem. 10: S1629–S1638.
pirical model was developed, and numeric optimization indicated
https://doi.org/10.1016/j.arabjc.2013.06.005.
that maximum adsorption capacity could be obtained from beads
Barde, W. 2015. “Development and characterization of unidirectional
prepared using 0.25 g of sodium alginate and 0.5 g of lime-iron aligned macroporous alginate scaffold by ionotropic gelation using cal-
sludge in 25 mL of distilled water to produce a homogeneous mix- cium chloride.” Ph.D. dissertation. Dept. of Biotechnology and Medical
ture and added dropwise into a solution of 0.31 g CaCl2 in 25 mL Engineering, National Institute of Technology, Rourkela.
of distilled water. The accuracy of the RSM prediction was sub- Bhunia, P., and M. M. Ghangrekar. 2008. “Statistical modeling and
Downloaded from ascelibrary.org by 41.102.31.224 on 08/24/21. Copyright ASCE. For personal use only; all rights reserved.

sequently validated by laboratory studies that revealed a residual optimization of biomass granulation and COD removal in UASB reac-
error of 2.9%. tors treating low strength wastewaters.” Bioresour. Technol. 99 (10):
The beads exhibited a maximum phosphate adsorption capacity 4229–4238. https://doi.org/10.1016/j.biortech.2007.08.075.
of 8.3 mg=g that compared well to other reported adsorbents in the Boyd, G. E., A. W. Adamson, and L. S. Myers Jr. 1947. “The exchange
literature. EDS and thermodynamic analysis indicated that physi- adsorption of ions from aqueous solutions by organic zeolites. II:
sorption, ligand exchange, hydrogen bonding, and dipole interac- Kinetics.” J. Am. Chem. Soc. 69 (11): 2836–2848. https://doi.org/10
tion were the dominant attachment mechanisms. Mass transfer .1021/ja01203a066.
studies indicated that film diffusion was present; however, the re- Chan, L. W., H. Y. Lee, and P. W. Heng. 2006. “Mechanisms of external
and internal gelation and their impact on the functions of alginate as a
action was dominated by intraparticle diffusion.
coat and delivery system.” Carbohydr. Polym. 63 (2): 176–187. https://
At the given operational parameters, the solution pH influenced doi.org/10.1016/j.carbpol.2005.07.033.
the speciation of phosphate that in turn had a significant effect on Chen, H., D. Ma, Y. Li, Y. Liu, and Y. Wang. 2016. “Optimization the
the adsorption process. Kinetic data were best simulated using the process of microencapsulation of bifidobacterium bifidum BB01 by
pseudo-second order model while equilibrium data followed the Box-Behnken design.” Acta Universitatis Cibiniensis. Series E: Food
Langmuir isotherm at room temperature and the Sips isotherm Technol. 20 (2): 17–28. https://doi.org/10.1515/aucft-2016-0012.
at higher temperatures. Chen, J., Y. Cai, M. Clark, and Y. Yu. 2013. “Equilibrium and kinetic stud-
The findings of this study reveal that due to the high iron con- ies of phosphate removal from solution onto a hydrothermally modified
tent, this encapsulated waste material may be used as a novel oyster shell material.” PLoS One 8 (4): 1–8. https://doi.org/10.1371
adsorbent to remove phosphate (a limiting nonrenewable resource) /journal.pone.0060243.
from wastewater. It has a high adsorption capacity with the ability Chen, J., H. Kong, D. Wu, X. Chen, D. Zhang, and Z. Sun. 2007. “Phos-
to reduce synthetic phosphate concentrations from 45 mg=L to a phate immobilization from aqueous solution by fly ashes in relation to
final value of 4.5 mg=L (1.5 mg=L phosphorous) with an adsorb- their composition.” J. Hazard. Mater. 139 (2): 293–300. https://doi.org
ent dosage of 8.9 g=L. This makes it a promising adsorbent con- /10.1016/j.jhazmat.2006.06.034.
Chislock, M. F., E. Doster, R. A. Zitomer, and A. E. Wilson. 2013.
sidering the limit value of <2.0 mg=L phosphorus set by most
“Eutrophication: Causes, consequences, and controls in aquatic ecosys-
regulatory agencies for the discharge of treated effluents into
tems.” Nat. Educ. Knowl., 4 (4): 1–10.
sensitive waters. Chittoo, B. S., and C. Sutherland. 2017. “Phosphate removal and recovery
using lime-iron sludge: Adsorption, desorption, fractal analysis, mod-
eling and optimization using artificial neural network-genetic algo-
References rithm.” Desalin. Water Treat. 63: 227–240. https://doi.org/10.5004/dwt
.2017.20195.
Abdelhay, A., A. Al Bsoul, A. Al-Othman, N. M. Al-Ananzeh, I. Jum’h,
Crittenden, J. C., R. R. Trussell, D. W. Hand, K. J. Howe, and G.
and A. A. Al-Taani. 2018. “Kinetic and thermodynamic study of phos-
Tchobanoglous. 2012. MWH’s water treatment: Principles and design.
phate removal from water by adsorption onto (Arundo donax) reeds.”
Hoboken, NJ: Wiley.
Adsorpt. Sci. Technol. 36 (1–2): 46–61. https://doi.org/10.1177/026
Derakhshandeh, K., Z. Hamedi, M. Karimi, M. Amiri, and F. Ahmadi.
3617416684347.
2012. “Formulation optimization of low bioavailable drug loaded algi-
Adinarayana, K., K. B. Raju, and P. Ellaiah. 2004. “Investigations on
nate microparticles using artificial neural networks.” J. Rep. Pharm. Sci.
alkaline protease production with B. subtilis PE-11 immobilized in
calcium alginate gel beads.” Process Biochem. 39 (11): 1331–1339. 1 (1): 49–59.
https://doi.org/10.1016/S0032-9592(03)00263-2. Desai, K. M., S. A. Survase, P. S. Saudagar, S. S. Lele, and R. S. Singhal.
Akpor, O. B. 2011. “Wastewater effluent discharge: Effects and treatment 2008. “Comparison of artificial neural network (ANN) and response
processes.” Proc., 3rd International Conf., on Chemical, Biological and surface methodology (RSM) in fermentation media optimization: Case
Environmental Engineering, 85–91. Singapore: IACSIT Press. study of fermentative production of scleroglucan.” Biochem. Eng. J.
Aksu, Z., and I. A. Isoglu. 2006. “Use of agricultural waste sugar beet pulp 41 (3): 266–273. https://doi.org/10.1016/j.bej.2008.05.009.
for the removal of Gemazol turquoise blue-G reactive dye from aqueous Dey, G., S. Bhupinder, and R. Banerjee. 2003. “Immobilization of alpha-
solution.” J. Hazard Mater. 137 (1): 418–430. https://doi.org/10.1016/j amylase produced by Bacillus circulans GRS 313.” Braz. Arch. Biol.
.jhazmat.2006.02.019. Technol. 46 (2): 167–176. https://doi.org/10.1590/S1516-8913200300
Aksu, Z., and E. Kabasakal. 2005. “Adsorption characteristics of 2, 4-di- 0200005.
chlorophenoxyacetic acid (2, 4-D) from aqueous solution on powdered Dodds, W. K., W. W. Bouska, J. L. Eitzmann, T. J. Pilger, K. L. Pitts, A. J.
activated carbon.” J. Environ. Sci. Health B 40 (4): 545–570. https://doi Riley, J. T. Schloesser, and D. J. Thornbrugh. 2008. “Eutrophication of
.org/10.1081/PFC-200061533. US freshwaters: Analysis of potential economic damages.” Environ.
Allen, S. J., Q. Gan, R. Matthews, and P. A. Johnson. 2005. “Mass transfer Sci. Technol. 43 (1): 12–19. https://doi.org/10.1021/es801217q.
processes in the adsorption of basic dyes by peanut hulls.” Ind. Eng. El-Kamel, A. H., O. M. N. Al-Gohary, and E. A. Hosny. 2003. “Alginate-
Chem. Res. 44 (6): 1942–1949. https://doi.org/10.1021/ie0489507. diltiazem hydrochloride beads: Optimization of formulation factors, in

© ASCE 04019019-16 J. Environ. Eng.

J. Environ. Eng., 2019, 145(5): 04019019


vitro and in vivo availability.” J. Microencapsulation 20 (2): 211–225. Khuri, A. I., and J. A. Cornell. 1996. Response surfaces. 2nd ed. New York:
https://doi.org/10.3109/02652040309178063. Dekker, Inc.
El-Naggar, I. M., E. S. Zakaria, I. M. Ali, M. Khalil, and M. F. El-Shahat. Langmuir, I. 1916. “The adsorption of gases on plane surfaces of glass,
2012. “Kinetic modeling analysis for the removal of cesium ions from mica and platinum.” J. Am. Chem. Soc. 40 (9): 1361–1403. https://doi
aqueous solutions using polyaniline titanotungstate.” Arab. J. Chem. .org/10.1021/ja02242a004.
5 (1): 109–119. https://doi.org/10.1016/j.arabjc.2010.09.028. Lao-Luque, C., M. Solé, X. Gamisans, C. Valderrama, and A. D. Dorado.
Ferreira, S. L. C., et al. 2007. “Box-Behnken design: An alternative for 2014. “Characterization of chromium (III) removal from aqueous solu-
the optimization of analytical methods.” Anal. Chim. Acta 597 (2): tions by an immature coal (leonardite). Toward a better understanding
179–186. https://doi.org/10.1016/j.aca.2007.07.011. of the phenomena involved.” Clean Technol. Environ. Policy 16 (1):
Fiol, N., J. Poch, and I. Villaescusa. 2004. “Chromium (VI) uptake by grape 127–136. https://doi.org/10.1007/s10098-013-0610-x.
stalks wastes encapsulated in calcium alginate beads: Equilibrium and Lee, K. Y., and D. J. Mooney. 2012. “Alginate: Properties and biomedical
kinetics studies.” Chem. Speciation Bioavailability 16 (1–2): 25–33. applications.” Prog. Polym. Sci. 37 (1): 106–126. https://doi.org/10
https://doi.org/10.3184/095422904782775153. .1016/j.progpolymsci.2011.06.003.
Foo, K. Y., and B. H. Hameed. 2010. “Insights into the modeling of ad- Lee, X. J., L. Y. Lee, L. P. Y. Foo, K. W. Tan, and D. G. Hassell. 2012.
sorption isotherm systems.” Chem. Eng. J. 156: 2–10. https://doi.org/10 “Evaluation of carbon-based nanosorbents synthesised by ethylene de-
.1016/j.cej.2009.09.013. composition on stainless steel substrates as potential sequestrating ma-
Downloaded from ascelibrary.org by 41.102.31.224 on 08/24/21. Copyright ASCE. For personal use only; all rights reserved.

Freundlich, H. M. F. 1906. “Over the adsorption in solution.” J. Phys. terials for nickel ions in aqueous solution.” J. Environ. Sci. 24 (9):
Chem. 57: 385–470. 1559–1568. https://doi.org/10.1016/S1001-0742(11)60987-X.
Fu, J., Y. Li, C. Ye, and C. Lin. 2012. “Study on the adsorption kinetics and Liu, R. X., J. L. Guo, and H. X. Tang. 2002. “Adsorption of fluoride, phos-
thermodynamics of DMF on macroporous adsorbents.” Acta Scien. phate, and arsenate ions on a new type of ion exchange fiber.” J. Colloid
Circum. 32 (3): 639–644. Interface Sci. 248 (2): 268–274. https://doi.org/10.1006/jcis.2002.8260.
Fundueanu, G., C. Nastruzzi, A. Carpov, J. Desbrieres, and M. Rinaudo. Loganathan, P., S. Vigneswaran, J. Kandasamy, and N. S. Bolan. 2014.
1999. “Physico-chemical characterization of Ca-alginate microparticles “Removal and recovery of phosphate from water using sorption.” Crit.
produced with different methods.” Biomaterials 20 (15): 1427–1435. Rev. Environ. Sci. Technol. 44 (8): 847–907. https://doi.org/10.1080
https://doi.org/10.1016/S0142-9612(99)00050-2. /10643389.2012.741311.
Furusawa, T., and J. M. Smith. 1973. “Fluid-particle and intraparticle mass Mahmood, Z., S. Nasir, N. Jamil, A. Sheikh, and A. Akram. 2015. “Ad-
transport rates in slurries.” Ind. Eng. Chem. Fund. 12 (2): 197–203. sorption studies of phosphate ions on alginate-calcium carbonate
https://doi.org/10.1021/i160046a009. composite beads.” Afr. J. Environ. Sci. Technol. 9 (3): 274–281.
Gadd, G. M. 2009. “Biosorption: Critical review of scientific rationale, https://doi.org/10.5897/AJEST2014.1784.
environmental importance and significance for pollution treatment.” Makris, K. C., W. G. Harris, G. A. O’Connor, T. A. Obreza, and H. A.
J. Chem. Technol. Biotechnol. 84 (1): 13–28. https://doi.org/10.1002 Elliott. 2005. “Physicochemical properties related to long-term phos-
/jctb.1999. phorus retention by drinking-water treatment residuals.” Environ.
Garson, D. G. 1991. “Interpreting neural network connection weights.” Sci. Technol. 39 (11): 4280–4289. https://doi.org/10.1021/es0480769.
Artif. Intell. Expert. 6 (4): 46–51. McKay, G., H. S. Blair, and J. Gardner. 1983. “The adsorption of dyes in
Genz, A., A. Kornmüller, and M. Jekel. 2004. “Advanced phosphorus re- chitin. III: Intraparticle diffusion processes.” J. Appl. Polym. Sci. 28 (5):
moval from membrane filtrates by adsorption on activated aluminium 1767–1778. https://doi.org/10.1002/app.1983.070280519.
oxide and granulated ferric hydroxide.” Water Res. 38 (16): 3523–3530. McKay, G., M. S. Otterburn, and J. A. Aga. 1985. “Fuller’s earth and fired
https://doi.org/10.1016/j.watres.2004.06.006. clay as adsorbents for dyestuffs.” Water Air Soil Pollut. 24 (3):
Girish, C. R., and V. R. Murty. 2016. “Mass transfer studies on adsorption 307–322. https://doi.org/10.1007/BF00161790.
of phenol from wastewater using Lantana camara, forest waste.” Int. J. Mohammed, W. T., and S. A. Rashid. 2012. “Phosphorus removal from
Chem. Eng., 2016: 1–11. https://doi.org/10.1155/2016/5809505. wastewater using oven-dried alum sludge.” Int. J. Chem. Eng.
Ha, J., C. Engler, and S. J. Lee. 2008. “Determination of diffusion coef- 2012: 1–11. https://doi.org/10.1155/2012/125296.
ficients and diffusion characteristics for chlorferon and diethylthiophos- Narsaiah, K., S. N. Jha, R. A. Wilson, H. M. Mandge, and M. R.
phate in Ca-alginate gel beads.” Biotechnol. Bioeng. 100 (4): 698–706. Manikantan. 2014. “Optimizing microencapsulation of nisin with
https://doi.org/10.1002/bit.21761. sodium alginate and guar gum.” J. Food Sci. Technol. 51 (12):
Ho, Y. S., and G. McKay. 1998a. “A comparison of chemisorption kinetic 4054–4059. https://doi.org/10.1007/s13197-012-0886-6.
models applied to pollutant removal on various sorbents.” Process Saf. Nethajia, S., A. Sivasamya, G. Thennarasua, and S. Saravanan. 2010.
Environ. Prot. 76 (4): 332–340. https://doi.org/10.1205/095758298 “Adsorption of malachite green dye onto activated carbon derived
529696. from Borassus aethiopum flower biomass.” J. Hazard. Mater. 181 (1):
Ho, Y. S., and G. McKay. 1998b. “Sorption of dye from aqueous solution 271–280. https://doi.org/10.1016/j.jhazmat.2010.05.008.
by peat.” Chem. Eng. J. 70 (2): 115–124. https://doi.org/10.1016/S0923 Nixon, S. W., C. A. Oviatt, J. Frithsen, and B. Sullivan. 1986. “Nutrients
-0467(98)00076-1. and the productivity of estuarine and coastal marine ecosystems.”
Ho, Y. S., and G. McKay. 1999. “Pseudo-second order model for sorption J. Limnol. Soc. South. Afr. 12 (1–2): 43–71. https://doi.org/10.1080
processes.” Process Biochem. 34 (5): 451–465. https://doi.org/10.1016 /03779688.1986.9639398.
/S0032-9592(98)00112-5. Nollet, H., M. Roels, P. Lutgen, P. Van der Meeren, and W. Verstraete.
Ismail, B., and K. Madhavan Nampoothiri. 2010. “Exopolysaccharide 2003. “Removal of PCBs from wastewater using fly ash.” Chemosphere
production and prevention of syneresis in starch using encapsulated 53 (6): 655–665. https://doi.org/10.1016/S0045-6535(03)00517-4.
probiotic Lactobacillus plantarum.” Food Technol. Biotechnol. Ociński, D., I. Jacukowicz-Sobala, and E. Kociołek-Balawejder. 2016.
48 (4): 484. “Alginate beads containing water treatment residuals for arsenic re-
Jansson-Charrier, M., E. Guibal, J. Roussy, B. Delanghe, and P. Le Cloirec. moval from water—Formation and adsorption studies.” Environ. Sci.
1996. “Vanadium (IV) sorption by chitosan: Kinetics and equilibrium.” Pollut. Res. 23 (24): 24527–24539. https://doi.org/10.1007/s11356
Water Res. 30 (2): 465–475. https://doi.org/10.1016/0043-1354(95) -016-6768-0.
00154-9. Patil, S., V. Deshmukh, S. Renukdas, and N. Patel. 2011. “Kinetics of ad-
Johansson, L., and J. P. Gustafsson. 2000. “Phosphate removal using blast sorption of crystal violet from aqueous solutions using different natural
furnace slags and opoka-mechanisms.” Wat. Res. 34 (1): 259–265. materials.” Int. J. Environ. Sci. 1 (6): 1116–1134.
https://doi.org/10.1016/S0043-1354(99)00135-9. Poots, V. J. P., G. Mckay, and J. J. Healy. 1976. “The removal of acid dye
Karimi, F., S. Rafiee, A. Taheri-Garavand, and M. Karimi. 2012. “Optimi- from effluent using natural adsorbents. I: Peat.” Water Res. 10 (12):
zation of an air drying process for Artemisia absinthium leaves using 1061–1066. https://doi.org/10.1016/0043-1354(76)90036-1.
response surface and artificial neural network models.” J. Taiwan Inst. Pretty, J. N., C. F. Mason, D. B. Nedwell, R. E. Hine, S. Leaf, and R. Dils.
Chem. Eng. 43 (1): 29–39. https://doi.org/10.1016/j.jtice.2011.04.005. 2003. “Environmental costs of freshwater eutrophication in England

© ASCE 04019019-17 J. Environ. Eng.

J. Environ. Eng., 2019, 145(5): 04019019


and Wales.” Environ. Sci. Technol., 37 (2): 201–208. https://doi.org/10 Sutherland, C., and C. Venkobachar. 2010. “A diffusion-chemisorption
.1021/es020793k. kinetic model for simulating biosorption using forest macro-fungus,
Redlich, O. J. D. L., and D. L. Peterson. 1959. “A useful adsorption iso- fomes fasciatus.” Int. Res. J. Plant Sci. 1 (4): 107–117.
therm.” J. Phys. Chem. 63 (6): 1024–1026. https://doi.org/10.1021 Sutherland, C., and C. Venkobachar. 2013. “Equilibrium modeling of
/j150576a611. Cu (II) biosorption onto untreated and treated forest macro-fungus
Rezaei, H., M. Haghshenasfard, and A. Moheb. 2017. “Optimization of dye Fomes fasciatus.” Int. J. Plant Anim. Environ. Sci. 3 (1): 193–203.
adsorption using Fe3O4 nanoparticles encapsulated with alginate beads Swan, E., and A. R. Urquhart. 1927. “Adsorption equations: A review of
by Taguchi method.” Adsorpt. Sci. Technol. 35 (1–2): 55–71. https://doi the literature.” J. Phys. Chem. 31: 251–276. https://doi.org/10.1021
.org/10.1177/0263617416667508. /j150272a008.
Robalds, A., L. Dreijalte, O. Bikovens, and M. Klavins. 2016. “A novel Ugurlu, A., and B. Salman. 1998. “Phosphorus removal by fly ash.”
peat-based biosorbent for the removal of phosphate from synthetic Environ. Int. 24 (8): 911–918. https://doi.org/10.1016/S0160-4120
and real wastewater and possible utilization of spent sorbent in land (98)00079-8.
application.” Desalin. Water Treat. 57 (28): 13285–13294. https://doi USEPA. 2008. Fate, transport, and transformation test guidelines. Adsorp-
tion/desorption. OPPTS 835.1230. Washington, DC: USEPA.
.org/10.1080/19443994.2015.1061450.
Vecino, X., R. Devesa-Rey, J. M. Cruz, and A. B. Moldes. 2013.
Saeh, I. S., and A. Khairuddin. 2009. “Implementation of artificial intelli-
“Entrapped peat in alginate beads as green adsorbent for the elimination
gence techniques for steady state security assessment in pool market.”
Downloaded from ascelibrary.org by 41.102.31.224 on 08/24/21. Copyright ASCE. For personal use only; all rights reserved.

of dye compounds from vinasses.” Water Air Soil Pollut. 224 (3):
Int. J. Eng. 3 (1): 1–11.
1448–1456. https://doi.org/10.1007/s11270-013-1448-x.
Sağ, Y., and Y. Aktay. 2000. “Mass transfer and equilibrium studies for Vermeulen, T. 1953. “Theory for irreversible and constant-pattern solid
the sorption of chromium ions onto chitin.” Process Biochem. 36 (1): diffusion.” Ind. Eng. Chem. 45 (8): 1664–1670. https://doi.org/10
157–173. .1021/ie50524a025.
Saha, A. K., and S. D. Ray. 2013. “Effect of cross-linked biodegradable Walker, G. M., and L. R. Weatherley. 1999. “Biological activated carbon
polymers on sustained release of sodium diclofenac-loaded micro- treatment of industrial wastewater in stirred tank reactors.” Chem. Eng.
spheres.” Braz. J. Pharm. Sci. 49 (4): 873–888. https://doi.org/10.1590 J. 75 (3): 201–206. https://doi.org/10.1016/S1385-8947(99)00109-6.
/S1984-82502013000400028. Walsh, A. 2005. “Reference conditions and eutrophication impacts in Irish
Sankalia, M. G., R. C. Mashru, J. M. Sankalia, and V. B. Sutariya. 2005. Rivers: Meeting the requirements of the water framework directive.”
“Papain entrapment in alginate beads for stability improvement and site- Accessed August 24, 2017. http://epa.ie/pubs/reports/research/water
specific delivery: Physicochemical characterization and factorial opti- /FS2-M1-2000-Final%20Report.pdf.
mization using neural network modeling.” AAPS Pharm. Sci. Technol. Wang, A., K. Zhou, X. Liu, F. Liu, and Q. Chen. 2017. “Development of
6 (2): E209–E222. https://doi.org/10.1208/pt060231. Mg–Al–La tri-metal mixed oxide entrapped in alginate for removal of
Senuma, Y., C. Lowe, Y. Zweifel, J. G. Hilborn, and I. Marison. 2000. fluoride from wastewater.” RSC Adv. 7 (50): 31221–31229. https://doi
“Alginate hydrogel microspheres and microcapsules prepared by spin- .org/10.1039/C7RA02566A.
ning disk atomization.” Biotechnol. Bioengin. 67 (5): 616–622. https:// Weber, W. J., and C. T. Miller. 1988. “Modeling the sorption of hydropho-
doi.org/10.1002/(SICI)1097-0290(20000305)67:5<616::AID-BIT12>3 bic contaminants by aquifer materials—I. Rates and equilibria.” Water
.0.CO;2-Z. Res. 22 (4): 457–464. https://doi.org/10.1016/0043-1354(88)90040-1.
Shahryari, Z., A. Sharifi, and A. Mohebbi. 2013. “Artificial neural network Weber, W. J., and J. C. Morris. 1963. “Kinetics of adsorption on carbon
(ANN) approach for modeling and formulation of phenol adsorption from solution.” J. Sanit. Eng. Div. 89 (2): 31–60.
onto activated carbon.” J. Eng. Thermophys. 22 (4): 322–336. https:// Xie, J., Y. Lin, C. Li, D. Wu, and H. Kong. 2015. “Removal and recovery of
doi.org/10.1134/S181023281304005X. phosphate from water by activated aluminum oxide and lanthanum
Shojaeimehr, T., F. Rahimpour, M. A. Khadivi, and M. Sadeghi. 2014. “A oxide.” Powder Technol. 269: 351–357. https://doi.org/10.1016/j
modeling study by response surface methodology (RSM) and artificial .powtec.2014.09.024.
neural network (ANN) on Cu 2+ adsorption optimization using light Yang, X. E., X. Wu, H. L. Hao, and Z. L. He. 2008. “Mechanisms and
expended clay aggregate (LECA).” Ind. Eng. Chem. Res. 20 (3): assessment of water eutrophication.” J. Zhejiang Univ. Sci. B 9 (3):
197–209. https://doi.org/10.1631/jzus.B0710626.
870–880. https://doi.org/10.1016/j.jiec.2013.06.017.
Yeniay, Ö. 2014. “Comparative study of algorithms for response surface
Singh, S., B. K. Gogoi, and R. L. Bezbaruah. 2012. “Calcium alginate as a
optimization.” Math. Comput. Appl. 19 (1): 93–104. https://doi.org/10
support material for immobilization of L-amino acid oxidase isolated
.3390/mca19010093.
from Aspergillus fumigatus.” IIOAB J. 3 (5): 7–11.
Yu, Y., Y. Y. Zhuang, and Z. H. Wang. 2001. “Adsorption of water-soluble
Sips, R. 1948. “On the structure of a catalyst surface.” J. Chem. Phys. dye onto functionalized resin.” J. Colloid Interface Sci. 242 (2):
16 (5): 490–495. https://doi.org/10.1063/1.1746922. 288–293. https://doi.org/10.1006/jcis.2001.7780.
Soni, A., A. Tiwari, and A. K. Bajpai. 2012. “Adsorption of o-nitrophenol Yuan, X., W. Xia, J. An, J. Yin, X. Zhou, and W. Yang. 2015. “Kinetic
onto nano iron oxide and alginate microspheres: Batch and column and thermodynamic studies on the phosphate adsorption removal by
studies.” Afr. J. Pure Appl. Chem. 6 (12): 161–173. dolomite mineral.” J. Chem. 2015: 1–8. https://doi.org/10.1155/2015
Subash, N., and R. Krishna Prasad. 2016. “Kinetics and mass transfer /853105.
models for sorption of titanium industry effluent in activated carbon.” Zeng, L., X. Li, and J. Liu. 2004. “Adsorptive removal of phosphate
Desalin. Water Treat. 57 (16): 7254–7261. https://doi.org/10.1080 from aqueous solutions using iron oxide tailings.” Water Res. 38 (5):
/19443994.2015.1016458. 1318–1326. https://doi.org/10.1016/j.watres.2003.12.009.
Sujitha, R., and K. Ravindhranath. 2017. “Extraction of phosphate from Zhao, B., Y. Zhang, X. Dou, H. Yuan, and M. Yang. 2015. “Granular ferric
polluted waters using calcium alginate beads doped with active carbon hydroxide adsorbent for phosphate removal: Demonstration preparation
derived from A. aspera plant as adsorbent.” J. Anal. Methods Chem. and field study.” Water Sci. Technol. 72 (12): 2179–2186. https://doi
2017: 1–13. https://doi.org/10.1155/2017/3610878. .org/10.2166/wst.2015.438.

© ASCE 04019019-18 J. Environ. Eng.

J. Environ. Eng., 2019, 145(5): 04019019

You might also like