You are on page 1of 12

Biochemical Engineering Journal 149 (2019) 107241

Contents lists available at ScienceDirect

Biochemical Engineering Journal


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

Regular article

Effect of alginate concentration in wastewater nutrient removal using T


alginate-immobilized microalgae beads: Uptake kinetics and adsorption
studies
Shantanu Banerjeea, Palas Balakdas Tiwadea, Kumar Sambhava, Chiranjib Banerjeeb,
Soubhik Kumar Bhaumika,

a
Department of Chemical Engineering, Indian Institute of Technology (Indian School of Mines), 826004, Dhanbad, India
b
Department of Environmental Science and Engineering, Indian Institute of Technology (Indian School of Mines), 826004, Dhanbad, India

HIGHLIGHTS

• Effect of alginate concentration on nutrient removal was investigated.


• Uptake kinetics and adsorption of nutrients onto alginate beads was evaluated.
• 3% alginate concentration optimal both for nutrient removal and adsorption.
• Adsorption of NH +
and PO 3−
- P onto alginate matrix through hydrogen bonding.
• 3% alginate concentration bear optimal pore size for nutrient access.
4 4

ARTICLE INFO ABSTRACT

Keywords: Effect of alginate concentration on wastewater nutrient removal by alginate-immobilized microalgae bead
Alginate concentration systems (AIMS) of Chlorella vulgaris was investigated to delineate the underlying role of alginate matrix in the
Nutrient removal removal process. The study includes kinetic evaluation of nutrient removal in AIMS of different alginate con-
Uptake kinetics centration (1.5–4.5% w/v%) followed by an assessment of abiotic adsorption characteristics of blank alginate
Adsorption
beads systems (BABS) of same alginate concentrations. Results establish 3% alginate concentration as optimal
Charge-interaction
Pore morphology
both for nutrient removal in AIMS and adsorption in BABS. The adsorption data in the latter fitted well with
Langmuir isotherms, suggesting monolayer physisorption, which was later corroborated through Fourier-
Transform Infrared Spectroscopy (FTIR) and Energy Dispersive X-ray (EDX) analysis. A scheme for adsorption of
NH4+ - N and PO43−- P on alginate matrix through hydrogen bonding was illustrated. The optimal trends were
explained based on pore morphology and bead structural integrity.

1. Introduction synthetic polymers [3]. A popular immobilization medium is the algi-


nate gel, due to its lower toxicity, higher porosity, transparency, and
Since its inception over a century ago, microalgae-based treatment milder physical environment to the cells [4]. Alginate-bead based algal
routes continue to be the most popular means for nutrient removal in nutrient removal has been studied with respect to various parameters
tertiary water treatment in order to prevent eutrophication. The re- such as bead concentration, bead cell count, type of alginate material,
moval of secondary metabolites and pollutants is achieved through in bead stability, manufacture cost etc. [5]. Some studies have highlighted
situ photosynthesis occurring ‘in parallel’ with the biological nutrient the role of alginate concentration on bead deterioration and nutrient
assimilation [1]. The processes boost the oxygen levels and offer a sy- removal [6,7]. Yet the underlying physicochemical interaction of algi-
nergistic advantage of algal biofuel production [1]. Algae-immobiliza- nate matrix with the nutrient species has not been exclusively ad-
tion techniques play a crucial role in harvesting the biomass without dressed.
the use of mechanical devices [2]. One popular method is the entrap- Compared to the free cells, the nutrient removal efficiency of algi-
ment that encloses the algal cells in a three-dimensional matrix of nate-immobilized microalgae bead system (AIMS) is considerably high


Corresponding author at: Department of Chemical Engineering, Indian Institute of Technology (Indian School of Mines), Dhanbad, 826004, Jharkhand, India.
E-mail address: soubhikge@gmail.com (S.K. Bhaumik).

https://doi.org/10.1016/j.bej.2019.107241
Received 8 January 2019; Received in revised form 18 May 2019; Accepted 23 May 2019
Available online 24 May 2019
1369-703X/ © 2019 Elsevier B.V. All rights reserved.
S. Banerjee, et al. Biochemical Engineering Journal 149 (2019) 107241

Nomenclature Sna Unassimilable substrate concentration (mg L−1)


t Time (h)
Ce Nutrient concentration at equilibrium(mg L−1) Va Volume of the adsorbate in solution (L)
Co Initial nutrient concentration(mg L−1) X Momentary concentration of microorganisms (mg L−1)
De Effective diffusivity (mm2 h−1) X0 Initial biomass concentration (mg L−1)
h Initial adsorption rate(mg mg−1 min−1) Xm Final biomass concentration (mg L−1)
k Kinetic constant (L mg−1 s−1) Y Microalgae yield coefficient; ratio of biomass amount
m Mass of the adsorbent (mg) produced per amount of substrate consumed (mg mg−1)
P Volumetric productivity (mg L−3 h−1)
q Adsorption capacity(mg g−1) Greek letters
qe,c Isotherm model calculated adsorption capacity(mg mg−1)
qe Adsorption capacity at equilibrium(mg g−1) μ Maximum specific growth rate (h−1)
Qm Maximum monolayer coverage capacity (mg g−1) μN Maximum specific NH4+ - N uptake rate (h−1)
r Radius of the bead (mm) μP Maximum specific PO43−- P uptake rate (h−1)
S Total substrate (nutrient) concentration at an instant ‘t’ Ƞ Effectiveness factor
(mg L−1) ϕ Thiele modulus
S0 Initial substrate concentration (mg L−1)

[8]. Immobilized Chlorella vulgaris exhibits uptake rates 15% and 31% 2. Experimental
higher than suspended free algae, for NH4+ - N and PO43−- P respec-
tively [9]. Nutrient removal by the same species in immobilized form is 2.1. Materials
reported to be over 95% for NH4+ - N ions and 99% for PO43−- P ions in
three days, compared to only 50% of NH4+ - N and 40% PO43−- P 2.1.1. Microalgae culture preparation
removal by their free counterparts [10]. One factor responsible for A stock suspension of unicellular microalgae Chlorella vulgaris was
enhanced removal is the protective environment provided by the sur- cultured in TRIS-Acetate-Phosphate (TAP) medium and incubated at
face to the indigenous microalgae species by acting as a physical barrier 27 ± 2 °C under controlled illumination (10,000 lx) with 16 h/8 h
against native microorganisms (through smaller pore size) in waste- light/dark photoperiods [15]. After the 12-day cultivation period, algal
water while allowing unlimited access of nutrients from the wastewater cells were harvested by centrifugation at 3000 g-forces for 10 min. The
[11]. Hydraulic Retention Time (HRT) of nutrient removal is lower for cell residues were washed and resuspended in deionized water resulting
AIMS (0.3–0.8 days) compared to high rate algal ponds (HRAP) (2–6 in a concentrated algal suspension with a cell density of 106 cells mL−1.
days) [12], rendering AIMS highly suitable for continuous wastewater
treatment [13].
Other factors for enhanced nutrient removal in AIMS exist through 2.1.2. Microalgae bead preparation
the possible chemical interactions between alginate matrix and nutrient Algal suspension was mixed with sodium alginate solution in 1:1 v/
species. Alginate is shown to have a high affinity for cations because of v ratio to prepare algal alginate suspension of (w/v %) 1.5%, 2%, 3%,
the carboxyl groups (COO−) [14]. Physical adsorption (through elec- 4%, 4.5%. The different suspension was allowed to drop precisely into a
trostatic interaction) of NH4+ and PO43−- P ions on the free binding calcium chloride solution (5.5%w/v%). After 12 h of hardening, Ca-
sites present on alginate structure is anticipated to be a precursor step Alginate beads of diameter 2–5 mm were formed. The beads were
before assimilation by the entrapped algae. The adsorption rates would rinsed with deionized water and saline solution (0.85% NaCl) and fi-
vary with the alginate concentration due to the proportionate con- nally starved in a fresh, sterile saline solution for about 3 days at
centration of binding sites available in the matrix. The hypothesis could 27 ± 2 °C under illumination of 12000 lx [16]. BABS of different al-
be cross-examined by analyzing the effect of alginate concentration in ginate concentrations were prepared in the same way using deionized
the removal process in AIMS as well as in blank alginate bead systems water in place of algal suspension.
(BABS). The mechanism of binding of nutrient ions on the alginate
matrix would be explored along with the effects of alginate con-
centration on matrix pore morphology and bead integrity. Thus, algi- 2.1.3. Wastewater sampling
nate concentration could be evaluated as a potential design optimiza- Wastewater was collected from a municipal secondary effluent
tion parameter in nutrient removal using AIMS. stream at Bekar bandh pond, Dhanbad, Jharkhand, India (23°48′1″N,
Herein, the effects of alginate concentration on wastewater nutrient 86°25′44″E). The composition of the wastewater is given as (mg L−1):
removal by immobilized Chlorella vulgaris were systematically analyzed NH4+ - N, 66.23 and PO43−- P, 10.85.
in AIMS and BABS. The nutrient removal of AIMS of different alginate
concentrations (w/v%): 1.5%, 2%, 3%, 4%, 4.5% was investigated ex-
2.2. Experimental procedure
perimentally. Kinetic parameters were evaluated through models for
biomass growth and nutrient removal with quantification of the diffu-
2.2.1. Nutrient removal experiments
sional resistance in the matrix. To delineate the role of alginate, in-
AIMS of different concentration along with control BABS (3%) and
dependent batch adsorption studies were conducted on BABS of the
free algae were treated with wastewater in conical flasks (1000 mL
same concentrations, and optimum conditions were evaluated. The
capacity) for 5 days under continuous sparging (2 L min−1), controlled
experimental data were fitted to adsorption isotherm and kinetic
temperature 27 ± 2 °C and illumination cycles of 16 h/8 h light/
models. The nutrient mechanisms were examined through Fourier-
darkness (10800 lx) (Fig. S1 in Supplementary material). pH was
Transform Infrared Spectroscopy (FTIR) analysis and Energy Dispersive
maintained below 8.5 (except in aerated wastewater) to prevent phy-
X-ray (EDX) analysis. The influence of alginate concentration on the
sical process such as NH3 stripping and precipitation of PO43−- P, which
pore morphology and bead stability was explored. The optimum algi-
ensured that the removal occurred solely by algal uptake [10]. For
nate concentration was assessed based on nutrient removal efficiency,
analysis, 25-mL samples were collected at increasing time intervals:
adsorption capacity, pore restrictive nutrient uptake and bead stability.
8 h–24 h.

2
S. Banerjee, et al. Biochemical Engineering Journal 149 (2019) 107241

2.2.2. Adsorption studies chamber using a vacuum pump (Tarsons, USA) by decreasing pressure
Batch adsorption experiments of NH4+ - N and PO43−- P onto BABS slowly and steadily to 80 mm Hg for 24 h and were examined using an
were conducted in 100 mL conical flasks on a thermostatic shaker at FE-SEM (Supra 55, Carl Zeiss, Germany) at 10 kV. The diameter of the
120 rpm at 27 °C. The adsorption capacity of BABS was calculated from random pores was measured using the integrated SmartSEM™ software.
the equation: Experiments were conducted to analyze the leakage of algal cells from
the AIMS into the aqueous medium. Leakage of cells was checked by
(Co Ce ) Va
qe =
(1) collecting 10 mL well-mixed medium samples and measuring their
m
turbidity (Turbiquant 3000 T, Merck, USA) at regular intervals.
Adsorption kinetics were evaluated as a function of contact time at
an initial nutrient concentration of 40 mg L−1 NH4+ - N and 20 mg L−1 2.3.4. Statistical analysis
PO43−- P at 7 pH. The effect of pH was examined for a range of 4 to 9 Each experiment was performed in triplicates. The mean, con-
for an initial nutrient concentration: NH4+ - N: 15 mg L−1; PO43−- P: fidence interval and standard deviation values of the triplicates were
5 mg L−1. For the same initial nutrient concentration and 7 pH, the calculated. Data were analyzed using parametric one-way analysis of
effect of adsorbent dosage (of 3% alginate beads) for a range of alginate variance (ANOVA) method. The statistical significance of the result was
mass from 0.3 g to 1.5 g was evaluated. The effect of initial nutrient evaluated using Student’s paired t-test. For probability p-value ≤ 0.05,
concentration on sorption was determined using concentration of NH4+ the effect of alginate gel concentration on nutrient removal and biomass
- N: 9.17 mg L−1 to 60 mg L−1 and corresponding PO43−- P: growth was considered significant, and otherwise not.
5.35 mg L−1 to 25 mg L−1 using 1000 mg sorbent dosage for 600 min.
The effect of calcium chloride concentration: (w/v%) 2%, 3%, 4%, 5% 3. Kinetic model
and 6% was observed for adsorption of nutrients on alginate. The nu-
trient medium used for adsorption studies consisted of both NH4+ - N 3.1. Algae growth kinetics
and PO43−- P and their adsorption were studied simultaneously.
The microbial growth is governed by Verhulst logistic kinetic model
2.2.3. Alginate characterization equation [21]:
The removal process was characterized both in terms of charge
X0 Xm e µt
based interaction and morphological structure. The former was studied X=
through FTIR and EDX for samples before and after treatment with
Xm X 0 + X0 e µt (2)
wastewater. The morphological study was conducted based on a Field- The growth is characterized by productivity, which is defined as the
Emission Scanning Electron Microscope (FE-SEM) to characterize the rate of change in biomass concentration from 10% higher than the in-
average pore size distribution. Finally, the durability and strength of the itial concentration to 90% of the final [22].
alginate was studied through leakage experiments.
µ (0.9Xm 1.1X0 )
P=
2.3. Analytical methods ln ( 9(Xm 1.1X0 )
1.1X0 ) (3)

2.3.1. Nutrient and biomass measurement


Concentrations of NH4+ - N and PO43−- P were measured according 3.2. Nutrient removal kinetics
to standard analysis methods: 4500-NH3 F [17] and 4500-P E [17] re-
spectively using UV–vis spectrophotometer (ThermoFisher Scientific, The relation for nutrient consumption is given as [22]:
Massachusetts, USA). Biomass of AIMS was measured by immersing the
beads in 0.1 M sodium tri-citrate solution. The released cell con- S=
( X0
Y
+ S0 (S0) Sna) (
Sna S0 ( X0
Y
+ S0 ))e µt

centration was evaluated spectrophotometrically (Dry weight, mg L−1


= (OD680 ×11.62) ̶ 3.9695; R2 = 0.996). The chlorophyll concentra-
(S0 Sna ) (S (
0
X0
Y
+ S0 ))e µt
(4)
tion inside the AIMS was analyzed according to a previous study [18]. Where Y is the microalgae yield coefficient defined as:
Ten beads were dissolved in 5 mL of 90% acetone and kept in darkness
(Xm X0 )
for around 24 h. Chlorophyll a and Chlorophyll b concentration level Y=
(S0 Sna ) (5)
were measured using UV–vis spectrophotometer (ThermoFisher Scien-
tific, Massachusetts, USA). The internal diffusion coefficient (De) for
NH4+ - N and PO43−- P inside the beads was determined using effusion 4. Results and discussion
experiments according to a previous study [19].
4.1. AIMS uptake batch kinetics
2.3.2. FTIR and EDX analysis
FTIR analysis for characterization was performed using FTIR spec- 4.1.1. Growth
trophotometer (MB3000, ABB) equipped with a MIRacle three reflec- The growth of microalgae was analysed for a period of 120 h
tion diamond attenuated total reflectance device (Pike Technologies, (Fig. 1(a)). The algal cell count increased from an initial of 7 × 106 cells
USA). The elemental composition of the bead was evaluated with bead−1 to approximately 90 × 106 cells bead−1. Chlorophyll levels of
Energy-Dispersive X-ray Spectroscopy (EDX) using a scanning electron different AIMS increased almost linearly with time (Fig. 1(b)). The
microscope (Zeiss EVO50, Oberkochen, Germany). trend of decreasing final chlorophyll levels were in the order:
3% > 4% > 2% > 4.5% > 1.5%. The increase in cell count and levels
2.3.3. Pore size and leakage measurements of chlorophyll confirmed photosynthesis [23,24]. The growth of algae
Beads for Scanning Electron Microscopy (SEM) imaging were pre- in AIMS and free suspended system followed similar pattern, differing
pared using the method described in a previous study [20]. Two beads slightly in the final biomass concentration (Xm):
of each alginate concentration were fixed in 5% (v/v) glutaraldehyde 3% > 4% > 2% > 4.5% > 1.5% > free cells.
solution in 1 M HEPES buffer at pH 7.2. They were washed twice using
the same buffer and then dehydrated using increasing ethanol con- 4.1.2. Nutrient removal
centrations: 30% for 10 min, 50% for 30 min, 75% for 10 h, 100% for Fig. 2 shows the evolution curves of NH4+ - N and PO43−- P con-
30 min and 100% for 60 min at 4 °C. The beads were vacuum dried in a centration in free algae systems, AIMS, and control BABS along with bar

3
S. Banerjee, et al. Biochemical Engineering Journal 149 (2019) 107241

39.7% PO43−- P in free algae.


The results of 3% alginate are compared against existing literature
in Table 1. The % removal obtained using 3% alginate is higher com-
pared to other concentrations for specific retention time. The remark-
able feature is that for 3% AIMS, the NH4+ - N (at 8 h) and PO43−- P (at
4 h) removal was 12% more than its closest alginate concentration.

4.1.3. Kinetic parameters


The latest generic approach for modelling multi-nutrient removal
accounts for growth and removal kinetics for each nutrient exclusively.
Verhulst model [21] was adopted for algal growth kinetics while Pho-
tobiotreatment model (PhBT) [22] was used for removal kinetics of
individual nutrient.

4.1.3.1. Growth kinetics. Fitting Eq. 2, the different kinetic growth


parameters of microalgae for all the concentrations are summarized
in Table 2. No significant difference was found in the comparison of the
specific growth between concentrations (p > 0.05) [10].The values for
the specific growth rates compared well with other studies (0.03-
0.04 h−1) having relatively higher levels of NH4+ - N and PO43−- P
than others [8,22,25]. No influence of alginate matrix on the growth
characteristics was observed [10,24,26].
An optimum value of productivity was observed for 3% alginate
(7.10 mg L−1 h−1). This value was significantly different from the
productivity value of 4% alginate (6.68 mg L−1 h−1), with a p-value of
0.0462. The value of productivity compared well with that obtained in
previous studies (6.25 mg L−1 h−1) [25]. However, it was slightly
higher than that obtained by other researchers (4.365 mg L−1 h−1) [27]
and (4.16 mg L−1 h−1) [28] for Chlorella vulgaris. The higher value of
productivity was attributed to the higher initial nutrient concentration
in the wastewater.

4.1.3.2. Nutrient removal kinetics. Photobiotreatment (PhBT) model


Fig. 1. (a) Evolution of biomass concentration in free and immobilized mi-
croalgae (symbol represent experimental values; lines represent model values)
[22] treats the growth of microalgae in the presence of nutrient-
(b) Evolution of chlorophyll concentration in different immobilized config- substrate as an autocatalytic reaction. PhBT model (Eq. 4) is
urations. applicable to simple reactive systems as in free algae. Nevertheless, it
can also be applied to AIMS for negligible diffusional resistances, as
confirmed through the estimates of the effectiveness factor : unity for
plot correlating the % removal with HRT. In both the cases, % removal
both NH4+ - N and PO43−- P (Table 3).
by control experiments with BABS was minimal, establishing algal as-
The summary of uptake rates is presented in Table 4. In the case of
similation as the primary mechanism of removal than chemical and
PO43−- P, due to the fast uptake by algae, a model could not be ob-
physical processes. Non-control of pH in aerated wastewater reflected
tained. Hence, the kinetic data for PO43−- P were obtained using ex-
the enhanced removal of NH4+ - N and PO43−- P, due to increased pH
perimental values. The highest maximum specific uptake rate was ob-
(> 8) resulting in ammonia stripping and phosphorus precipitation
served in 3% alginate: 0.137 h−1 and 0.47 h−1 for NH4+ - N and
[24].
PO43−- P respectively. The maximum specific NH4+ - N uptake rate was
similar to obtained in a study [22], specifically in the case of NH4+ - N
4.1.2.1. Ammonium removal. In AIMS, the NH4+ - N concentration rich wastewater medium. The maximum uptake rate of PO43−- P was
reduced rapidly during the initial 24 h, slowed down between 36–72 h, much higher than that of NH4+ - N in each case [29]. This agrees with
and almost settled to zero afterwards (Fig. 2a). For free cells, it previous studies that reported the maximum specific uptake rate for
continued to decrease steadily although the % removal was phosphorus to be up to three-four times that of nitrogen [22].
significantly lesser (28% difference in removal) compared to AIMS A correlation of kinetics between NH4+ - N and that of PO43−- P
within 48 h (Fig. 2b). After 48 h, the trend for highest NH4+ - N exists through the sequence in which they are required in the cell
removal using AIMS was (in decreasing order) as: metabolism: Demand for phosphorus depends on the level of nitrogen
3% > 4% > 2% > 4.5% > 1.5% > free algae. However, net removal imbibed [30]. P is stored in the ribosomal RNA and participates in the
after 120 h was nearly equal for all alginate concentrations (> 98%), synthesis of cellular molecules such as phospholipids and nucleic acids.
slightly differing from free algae (94.31%). The higher performance for Being a precursor for other uptakes, uptake of NH4+ - N serves as the
AIMS (3% Alginate) is also evident from the removal achieved in 8 h rate-limiting step in the overall removal kinetics. Hence,understanding
HRT: 46.9% compared to 17.8% removal by free algae. the mechanism of NH4+ - N uptake is sufficient for overall kinetics.

4.1.2.2. Phosphorus removal. The PO43−- P concentrations declined 4.2. Adsorption studies on BABS
sharply in all alginate concentrations within 8 h and eventually
stabilized (Fig. 2c). Removal after 16 h was observed in the following 4.2.1. Effect of contact time
trend (in decreasing order): 3% > 4% > 2% > 1.5% > 4.5% > free The effect of time was observed from 0 to 720 min and is shown in
algae. The net removal after 120 h was > 95% for all alginate Fig. 3 (a)-(b). The equilibrium time was approximately 600 min for both
concentrations and 90.36% for free algae (Fig. 2d). Within 8 h, 86.2% NH4+ - N and PO43−- P. The adsorption capacity increased to
PO43−- P removal was achieved in AIMS (3% Alginate) compared to 4.97 mg g−1 for NH4+ - N and 2.16 mg g−1 for PO43−- P. The difference

4
S. Banerjee, et al. Biochemical Engineering Journal 149 (2019) 107241

Fig. 2. (a) Evolution of NH4+ - N uptake in free algae, AIMS and BABS (symbol represent experimental values; lines represent model values except for A and B) (b) %
NH4+ - N removal for AIMS and free algae (c) Evolution of PO43−- P uptake in free algae, AIMS and BABS (symbol and line both represent experimental values) (d) %
PO43−- P removal for AIMS and free algae.

Table 1
Nutrient removal efficiency of different alginate-immobilized microalgae nutrient removal systems.
Culture Microalgae Algae- Alginate Retention NH4+ - N PO43−- P References
condition concentration (%) time (d)
Initial Concentration Removal Initial Concentration Removal
(mg L−1) Efficiency (%) (mg L−1) Efficiency (%)

Batch C. vulgaris 3% 1 66 70 11 89 Present Work


Batch C. vulgaris (P. 2% 1 50 49 10 36 [53]
Putida*)
Batch C. vulgaris FACHB- 3% 1 49 98 13 100 [54]
30
Continuous C. vulgaris FACHB- 3% 1 49 97 13 87 [54]
30 (P. Putida*)
Batch S. obliquus 3% 1 10 30 1 28 [7]
Batch C. vulgaris 2% 1 66 60 11 83 Present work
Batch C. vulgaris 2% 2 – 80 – 53 [23]
Batch C. sorokiniana 2% 2 48 63 3 53 [11]
Batch C. vulgaris 2% 1 – – 39 4 [16]
Batch C. vulgaris 2% 1 37.5 - 47.5 40 – 94 [10]
Batch C. vulgaris 2% 1 9.8 33 35.7 1.5 [55]

* Co-immobilized microorganism.

among the adsorption capacities of different alginate concentrations was 4.2.2. Effect of pH
appreciable for NH4+ - N but negligible for PO43−- P. Almost all alginate The effect of pH on NH4+ - N and PO43−- P adsorption was eval-
concentrations reached equilibrium at the same time. The difference was uated (Fig. 3 (c)–(d)). Maximum adsorption capacity for NH4+ - N and
observed in adsorption capacities, which depended upon the number of PO43−- P adsorption was observed at 7.0 pH (2.21 mg g-1) and at 9.0 pH
active sites available. 3% alginate showed relatively higher adsorption (1.9 mg g-1) tively, for 3% alginate. Guluronic and mannuronic acid
capacity for both nutrients. As per these results, rest of batch adsorption residues of alginate have pKa values of 3.38 and 3.65 respectively.
experiments on the physical properties were conducted for a time in- When the alginate is immersed in pH greater than its pKa, the hydrogel
terval of 600 min, to ensure that equilibrium was reached. becomes entirely negative charged through the COO− ions, which

5
S. Banerjee, et al. Biochemical Engineering Journal 149 (2019) 107241

Table 2
Kinetic parameters for microalgae growth.
Xo (mg L−1) Xm (mg L−1) μ (h−1) R2 P (mg L−1 h−1)

Free algae 144.33 ± 25.02 732.67 ± 16.16 0.036 ± 0.0004 0.980 5.25 ± 0.10
1.5% 143.00 ± 18.08 853.33 ± 41.63 0.037 ± 0.0010 0.992 6.13 ± 0.15
2% 137.33 ± 16.62 900.10 ± 30.00 0.038 ± 0.0007 0.988 6.53 ± 0.14
3% 143.67 ± 17.78 960.00 ± 45.82 0.039 ± 0.0010 0.986 7.10 ± 0.17
4% 144.67 ± 13.01 911.67 ± 43.10 0.038 ± 0.0009 0.982 6.68 ± 0.19
4.5% 139.00 ± 12.52 891.33 ± 47.59 0.037 ± 0.0008 0.987 6.39 ± 0.22

Values represent means of triplicates; ± shows standard deviation from mean.

Table 3
Diffusion-Reaction model parameters: for NH4+ - N and PO43−- P uptake.
Bead (% alginate) Radius (mm) k × 10−5 (L mg−1 h-1) k × 10−3 (L mg−1 h-1) De (mm2 h−1) ϕ × 10−4 ϕ × 10−3 η

NH4+ - N PO43−- P NH4+ - N PO43−- P NH4+ - N PO43−- P NH4+ - N PO43−- P

1.5% 2.080 9.47 ± 1.43 8.21 ± 1.71 1.5191 0.9630 1.00 ± 0.07 1.17 ± 0.13 0.999 0.999
2% 2.175 10.4 ± 1.70 8.00 ± 1.50 1.3242 0.9384 1.31 ± 0.10 1.36 ± 0.13 0.999 0.999
3% 2.210 10.9 ± 1.62 8.05 ± 1.74 1.2483 0.9035 1.30 ± 0.09 1.31 ± 0.15 0.999 0.999
4% 2.250 10.11 ± 2.14 7.24 ± 1.54 1.1460 0.8826 1.44 ± 0.15 1.40 ± 0.15 0.999 0.999
4.5% 2.375 10.3 ± 1.90 7.22 ± 1.15 1.0297 0.8821 1.78 ± 0.16 1.61 ± 0.13 0.999 0.999

Values represent means of triplicates; ± shows standard deviation from mean.

Table 4
Kinetic parameters for NH4+ - N and PO43−- P removal.
So (mg L−1) Sna (mg L−1) μN (h−1) R2 Y k × 10−5 (L mg−1 h−1)

NH4+ - N Free algae 66.23 ± 3.52 3.76 ± 1.57 0.061 ± 0.0004 0.975 13.85 ± 0.79 7.06 ± 0.13
1.5% 66.23 ± 3.52 0.89 ± 0.17 0.095 ± 0.0053 0.969 15.53 ± 0.71 9.47 ± 1.43
2% 66.23 ± 3.52 0.06 ± 0.09 0.112 ± 0.0073 0.983 16.34 ± 0.78 10.40 ± 1.70
3% 66.23 ± 3.52 0.06 ± 0.02 0.137 ± 0.0075 0.986 19.05 ± 0.79 10.90 ± 1.62
4% 66.23 ± 3.52 0.15 ± 0.003 0.119 ± 0.0080 0.977 18.02 ± 0.77 10.11 ± 2.14
4.5% 66.23 ± 3.52 0.21 ± 0.02 0.107 ± 0.0090 0.966 15.84 ± 0.91 10.34 ± 1.90

So (mg L−1) Sna (mg L−1) μP (h−1) R2 Y k × 10−3 (L mg−1 h−1)

PO43−- P Free algae 10.85 ± 1.03 6.53 ± 0.66 0.24 ± 0.022 0.972 3.10 ± 0.26 8.10 ± 2.07
1.5% 10.85 ± 1.03 1.98 ± 1.00 0.27 ± 0.025 0.961 3.83 ± 0.76 8.21 ± 1.71
2% 10.85 ± 1.03 1.90 ± 0.88 0.34 ± 0.053 0.984 4.76 ± 0.63 8.00 ± 1.50
3% 10.85 ± 1.03 1.50 ± 0.94 0.47 ± 0.046 0.984 6.42 ± 1.53 8.05 ± 1.74
4% 10.85 ± 1.03 1.81 ± 0.94 0.35 ± 0.044 0.975 5.58 ± 1.39 7.24 ± 1.54
4.5% 10.85 ± 1.03 2.21 ± 0.77 0.28 ± 0.012 0.965 4.61 ± 0.54 7.22 ± 1.15

Values represent means of triplicates; ± shows standard deviation from mean.

attract the oppositely charged cation NH4+ - N [31]. As the pH in- 4.2.4. Effect of adsorbent dosage
creases above 4 up to 7, the COO− ions in the matrix increases which The % adsorption extent increased with increase in the sorbent
leads to strong interactions between the cationic NH4+ at 7 pH [8]. In (alginate) mass in the solution for NH4+ - N and decreased for PO43−- P
the case of PO43−- P, the adsorption capacity increased as the pH in- (Fig. S2 in Supplementary material). In the case of NH4+ - N, the in-
creased from 6 onwards. At pH higher than 7.0, the calcium ions inside crease in alginate mass increases the number of active sites which leads
the alginate favoured the precipitation of PO43- ions as calcium phos- to an increase in adsorption. For PO43−- P, the adsorption decreased
phate [10,24], hence, the adsorption of PO43−- P was higher at elevated due to the increasing repulsion forces (COO− ions) which are similar to
pH. Wan’s result [32]. Adsorption capacity for both NH4+ - N and PO43−- P
decreased with increasing alginate dosage. The increase in adsorbent
dosage could possibly lead to unsaturation of active sites and ag-
4.2.3. Effect of initial nutrient concentration gregation, which reduces the available total surface area available for
For both NH4+ - N and PO43−- P, at a fixed amount of adsorbent adsorption [33,34].
dosage, with the increase in their initial nutrient concentration the
adsorption capacity increased but the % extent of adsorption decreased.
In the case of NH4+ - N, the capacity reached up to 6.47 mg g−1 for 3% 4.2.5. Effect of CaCl2 concentration
BABS. The increase in adsorption capacity can be linked to an increase Cross-linking solution is an important parameter determining nu-
in affinity for charge interactions for an increase in cations (NH4+) and trient adsorption for a specific alginate concentration. For NH4+ - N,
the enhanced concentration gradient. For PO43−- P the increase in increased CaCl2 concentration lead to increase in % NH4+ - N removal
concentration led to an increase of adsorption capacity by (Fig. S3 in Supplementary material). Increase in CaCl2 concentration
1.328 mg g−1. The enhanced concentration gradient at higher PO43−- P promotes additional inter crosslinking inside alginate [35], which in-
concentrations increases the amount adsorbed onto alginate. creases the pore formation inside the matrix and the active COO− sites.
At lower CaCl2 (1–2 w/v%) concentration, NH4+ - N removal decreased
due to a lower degree of crosslinking (lower COO− sites). In the case of

6
S. Banerjee, et al. Biochemical Engineering Journal 149 (2019) 107241

Fig. 3. Sorption studies experiment effect on BABS (a) Effect of contact time on NH4+ - N (b) Effect of contact time on PO43−- P (c) pH on NH4+ - N (d) pH on
PO43−- P (e) Initial nutrient concentration on NH4+ - N (f) Initial nutrient concentration on PO43−- P.

PO43−- P, the % removal declined with increasing concentration of [36] (models described in Table S1 in Supplementary material). The
CaCl2 due to an increase in anionic charge due to the increase of COO− models were fitted in their non-linear form with the experimental data
sites. Therefore, an optimum concentration of 3% CaCl2 is desirable to in Fig. 4(a)-(b) with the fitted parameters listed in Table S2 (Supple-
achieve proper crosslinking of the alginate matrix [7]. mentary material). The best fit was observed for the Langmuir model
In the BABS experiments, 3% alginate concentration exhibited the (NH4+ - N: R2 = 0.990 and PO43−- P: R2 = 0.974) followed by the
highest adsorption capacity for all cases. Hence, this concentration was R − P model (NH4+ - N: R2 = 0.988 and PO43−- P: R2 = 0.973) and the
used to evaluate the adsorption isotherms and kinetics. Freundlich model (NH4+ - N: R2 = 0.973 and PO43−- P: R2 = 0.958).
The value of exponent β in R − P approached unity for both NH4+ - N
and PO43−- P implying conformity to Langmuir isotherm model. This
4.2.6. Adsorption isotherm models implies that the adsorption of the nutrients occurs at specific homo-
The adsorption equilibrium data for 3% BABS was interpreted using genous sites as mono-layer adsorption onto alginate matrix and surface.
Langmuir, Freundlich and Redlich-Peterson (R − P) isotherm models

7
S. Banerjee, et al. Biochemical Engineering Journal 149 (2019) 107241

Fig. 4. Adsorption isotherms obtained through experimental data nonlinear fitting (a) for NH4+ - N (b) for PO43−- P. Adsorption kinetics of BABS using Pseudo first
order model and Pseudo-second order model (c) for NH4+ - N (d) for PO43−- P.

4.2.7. Adsorption kinetics


The adsorption process of N and P onto alginate was characterized
based on Lagergren pseudo-first-order model and Ho pseudo-second-
order model [40,41] (models described in Table S3 in Supplementary
material). The non-linear constants obtained from pseudo-first-order
and pseudo-second-order models were calculated from non-linear ki-
netic plots given in Fig. 4(c)-(d) and summarized in Table S4 (Supple-
mentary material). The pseudo-first order model (NH4+ - N: R2 > 0.996
and PO43−- P: R2 > 0.991) fitted the model better than pseudo-second-
order model (NH4+ - N: R2 > 0.98 and PO43−- P: R2 > 0.98). The
calculated values of qe,c from the pseudo-first-order model closely ap-
proximates the experimental qe values than the qe,c obtained from the
pseudo-second-order model. This implies that the adsorption process
for NH4+ - N and PO43−- P was physisorption. The high adsorption
capacity of 3% alginate BABS compared to other alginate concentra-
tions promotes nutrient removal, and this effect is amplified in the case
of AIMS due to the presence of microalgae.

Fig. 5. FTIR spectra of AIMS and BABS, before and after adsorption.
4.3. Alginate characterization and mechanism investigation

The adsorption of NH4+ - N and PO43−- P onto alginate was mainly 4.3.1. FTIR and EDX analysis
because of electrostatic attraction between functional groups and the The mechanism was explored from FTIR (Fig. 5) and EDX studies
nutrient ions in the aqueous solution as seen in other studies [37–39]. conducted on AIMS and BABS before and after treatment with waste-
The maximum monolayer adsorption capacities calculated from Lang- water. The FTIR spectra of untreated BABS and AIMS exhibited peaks at
muir isotherm model of NH4+ - N and PO43−- P are 15.58 mg g−1 and 3330 cm−1 (NH stretching), 1641 cm−1 (COO − asymmetric stretching
3.29 mg g−1, respectively. (strong)), and 1033 cm−1 (CO − stretching of alcoholic groups).
Spectra of treated AIMS and BABS beads revealed the predominant
peaks 3360 cm−1 (NH stretching and Hydrogen bond (broad);

8
S. Banerjee, et al. Biochemical Engineering Journal 149 (2019) 107241

Fig. 6. Mechanism for NH4+ - N and PO43−- P adsorption and algal assimilation on AIMS.

overlapping with OH stretching), 1641 cm-1 (COO asymmetric Supplementary material) showed the emergence of the peak for ni-
stretching), 1423 cm-1 (NH deformation vibration) and at 1035 cm-1 trogen and phosphorous after treatment. The nitrogen content (wt%)
(P − O stretching (strong)) [42–44]. The EDX analysis (Fig S4 in (N) increased from an initial value of 0% to 7.234% in BABS and

9
S. Banerjee, et al. Biochemical Engineering Journal 149 (2019) 107241

Fig. 7. SEM images of pores in different alginate concentration beads (a) Typical morphological surface of a bead (b) 1.5% Alginate bead (c) 2% Alginate bead (d)
3% Alginate bead (e) 4% Alginate bead (f) 4.5% Alginate bead.

Fig. 8. Cell leakage from AIMS into a nutrient-deficient medium (a) Evolution of leakage concentration (turbidity) with time (b) Turbid solutions of AIMS, obtained
after the 5th day of the experiment.

4.118% in AIMS. Phosphorus content (wt%) increased from an initial interaction of the PO43− ion with OH is through hydrogen bonding
value of 0% to 5.095% in BABS and 1.295% in AIMS. The above ana- [48].
lysis showed a similar behaviour of both AIMS and BABS during ad- The above trends comply with the adsorption isotherm results as
sorption of nutrients into the bead, except for the difference in N and P well. Langmuir isotherm model assumes monolayer adsorption, char-
content, ruling out any influence of microalgae on the adsorption of acterized by homogenous adsorption, i.e. all sites have equal affinity for
nutrients onto alginate. adsorbate [36]. The NH4+ ion adsorption is based on electrostatic at-
traction between the protonated amino group and the negatively
charged COO− ions. This interaction involves constant and equal ad-
4.3.2. Mechanism of nutrient adsorption
sorption energy which agree with the correlation coefficients obtained
The interaction of nutrient with the alginate structure is illustrated
from Langmuir isotherm. In the case of AIMS, these ions (NH4+ and
in Fig. 6. An interesting decline in the peak at 3361 cm−1 to 3330 cm−1
PO43−), which are weakly bonded to the alginate structure, are as-
is observed after the adsorption of nutrients. Hydrogen bonding causes
similated by the microalgae in the course of time. The assimilation
significant band broadening and lowering of the mean adsorption fre-
results in unoccupied vacancies in the alginate chain at COO− and OH
quency as noted in the region [44]. In the presence of an electrostatic
groups, which causes more uptake of nutrients from outside the algi-
environment within the polymeric matrix, the NH4+ ion forms a weak
nate bead environment. This process continues until the microalgae
hydrogen bond with the carbonyl oxygen of the free COO− ions
limit the uptake of nutrients (possibly due to the stagnant phase) and
[45,46]. The observations were in line with Etter’s rule, which states
consequently saturating the bead with adsorbed nutrients.
that best hydrogen pair donor (NH4+, in our study) forms a hydrogen
bond with the best available hydrogen bond acceptor (COO−) [47]. In
case of PO43−- P, the emergence of a new peak at 1035 cm−1 is at- 4.3.3. Morphological studies
tributed to P − O stretching, confirming the presence of PO43−- P being Apart from adsorption characteristics, the concentration of the al-
adsorbed. The broadening of the peak suggests that the physical ginate gel affects the removal indirectly through varying pore size. SEM

10
S. Banerjee, et al. Biochemical Engineering Journal 149 (2019) 107241

images of BABS are shown in Fig. 7. The average surface pore size of References
alginate in the decreasing order is as follows: 1.5% alginate
(0.542 ± 0.094 μm), 2% alginate (0.506 ± 0.087 μm), 3% alginate [1] C. Alcántara, E. Posadas, B. Guieysse, R. Muñoz, Microalgae-based wastewater
(0.394 ± 0.076 μm), 4% alginate (0.240 ± 0.180 μm) and 4.5% algi- treatment, Handb. Mar. Microalgae, Elsevier, 2015, pp. 439–455.
[2] N. Mallick, Biotechnological potential of immobilized algae for wastewater N, P and
nate (0.119 ± 0.180 μm). The increase in alginate gel concentration metal removal: a review, Biometals 15 (2002) 377–390.
causes the matrix to become denser and reduces the pore size, leading [3] L.E. De-Bashan, Y. Bashan, Immobilized microalgae for removing pollutants: review
to restrictive uptake of nutrients or other macro-molecules [49]. The of practical aspects, Bioresour. Technol. 101 (2010) 1611–1627.
[4] I. Moreno-Garrido, Microalgae immobilization: current techniques and uses,
porosity of the matrix also decreases with increasing alginate con- Bioresour. Technol. 99 (2008) 3949–3964.
centration, which reduces the total void volume inside the matrix [5] M. Kube, A. Mohseni, L. Fan, F. Roddick, Impact of alginate selection for waste-
[50,51]. water treatment by immobilised Chlorella vulgaris, Chem. Eng. J. 358 (2019)
1601–1609.
[6] I. Cruz, Y. Bashan, G. Hernàndez-Carmona, L.E. De-Bashan, Biological deterioration
of alginate beads containing immobilized microalgae and bacteria during tertiary
4.3.4. Durability studies of alginate bead wastewater treatment, Appl. Microbiol. Biotechnol. 97 (2013) 9847–9858.
The leakage tendency was examined through separate experiments [7] P. Wang, Z. Li, J. Bai, Y. Lang, H. Hu, Optimization of microalgal bead preparation
with Scenedesmus obliquus for both nutrient removal and lipid production, Ecol.
on AIMS conducted for 15 days (Fig. 8). Leakage increased with de-
Eng. 92 (2016) 236–242.
creasing alginate concentration and was highest in 1.5% alginate. After [8] P. Chevalier, J. de la Noüe, Wastewater nutrient removal with microalgae im-
the 8th day, the 1.5% alginate beads lost their integrity while the 2% mobilized in carrageenan, Enzyme Microb. Technol. 7 (1985) 621–624.
and 3% were seen to disintegrate on the 11th day and 14th day, re- [9] N. Mallick, L.C. Rai, Removal of inorganic ions from wastewaters by immobilized
microalgae, World J. Microbiol. Biotechnol. 10 (1994) 439–443.
spectively. Increasing alginate concentration increases the crosslink [10] P.S. Lau, N.F.Y. Tam, Y.S. Wong, Wastewater nutrients (N and P) removal by car-
density and solid volume fraction, resulting in higher mechanical rageenan and alginate immobilized Chlorella vulgaris, Environ. Technol. 18 (1997)
strength and structural integrity [52]. Higher alginate concentration 945–951, https://doi.org/10.1080/09593331808616614.
[11] S.A. Covarrubias, L.E. de-Bashan, M. Moreno, Y. Bashan, Alginate beads provide a
(> 3%) were observed to retain their integrity until the end of the beneficial physical barrier against native microorganisms in wastewater treated
experiment, and the cell leakage was significantly lesser than lower with immobilized bacteria and microalgae, Appl. Microbiol. Biotechnol. 93 (2012)
alginate concentration beads. 2669–2680.
[12] J.P. Hoffmann, Wastewater treatment with suspended and nonsuspended algae, J.
The findings indicate that a balance in the alginate concentration is Phycol. 34 (1998) 757–763.
vital to obtain maximum nutrient adsorption and uptake and mini- [13] G. Mujtaba, K. Lee, Treatment of real wastewater using co-culture of immobilized
mizing the subsequent leakage of cells, which can result in higher nu- Chlorella vulgaris and suspended activated sludge, Water Res. 120 (2017) 174–184.
[14] J.P. Chen, L. Hong, S. Wu, L. Wang, Elucidation of interactions between metal ions
trient removal. Alginate concentration of 3% was seen to be the best
and Ca alginate-based ion-exchange resin by spectroscopic analysis and modeling
among all concentrations, in terms of adsorption on alginate and pore simulation, Langmuir 18 (2002) 9413–9421.
restrictive nutrient uptake. A previous study also reported 3% alginate [15] D.S. Gorman, R.P. Levine, Cytochrome f and plastocyanin: their sequence in the
photosynthetic electron transport chain of Chlamydomonas reinhardi, Proc. Natl.
concentration to be an optimum for nutrient removal and lipid pro-
Acad. Sci. 54 (1965) 1665–1669.
ductivity [7]. Since immobilization increases cost to wastewater treat- [16] J.-P. Hernandez, L.E. de-Bashan, Y. Bashan, Starvation enhances phosphorus re-
ment, hence, for the overall process to be economical, the im- moval from wastewater by the microalga Chlorella spp. Co-immobilized with
mobilization process should be optimized. As such, the careful selection Azospirillum brasilense, Enzyme Microb. Technol. 38 (2006) 190–198.
[17] A.P.H. Association, APHA, Stand. Methods Exam. Water Wastewater, 21st ed., Am.
of sodium alginate concentration used to prepare the bead is imperative Public Heal. Assoc., Washingt. DC, 2005 1220p. (n.d.).
in enhancing the economic capability of the algal wastewater treat- [18] D.I. Arnon, Copper enzymes in isolated chloroplasts. Polyphenoloxidase in Beta
ment. Due to the advantage of obtaining higher removal in lesser HRT, vulgaris, Plant Physiol. 24 (1949) 1.
[19] P. Grunwald, Determination of effective diffusion coefficients—an important
3% alginate shall be used as packings in packed bed reactor for con- parameters for the efficiency of immobilized biocatalysts, Biochem. Educ. 17 (1989)
tinuous nutrient removal in our future application of this batch study. 99–102.
[20] Y. Bashan, H. Levanony, E. Klein, Evidence for a weak active external adsorption of
Azospirillum brasilense Cd to wheat roots, Microbiology 132 (1986) 3069–3073.
[21] M. Vogels, R. Zoeckler, D.M. Stasiw, L.C. Cerny, PF Verhulst’s “notice sur la loi que
5. Conclusion la populations suit dans son accroissement” from correspondence mathematique et
physique. Ghent, vol. X, 1838, J. Biol. Phys. 3 (1975) 183–192.
The present study analyzes the effect of alginate concentration on [22] J. Ruiz, Z. Arbib, P.D. Álvarez-Díaz, C. Garrido-Pérez, J. Barragán, J.A. Perales,
Photobiotreatment model (PhBT): a kinetic model for microalgae biomass growth
wastewater nutrient removal using alginate immobilized microalgae
and nutrient removal in wastewater, Environ. Technol. 34 (2013) 979–991.
beads. Within the alginate systems, AIMS of 3% alginate exhibited [23] A. Ruiz-Marin, L.G. Mendoza-Espinosa, T. Stephenson, Growth and nutrient re-
optimal values for specific nutrient uptake rate and productivity. moval in free and immobilized green algae in batch and semi-continuous cultures
treating real wastewater, Bioresour. Technol. 101 (2010) 58–64, https://doi.org/
Adsorption studies on BABS characterized the enhanced nutrient ad-
10.1016/j.biortech.2009.02.076.
sorption by the immobilizing matrix. Elaborate FTIR and EDX analysis [24] N.F.Y. Tam, Y.S. Wong, Effect of immobilized microalgal bead concentrations on
on BABS confirmed the role of alginate in providing adsorption sites for wastewater nutrient removal, Environ. Pollut. 107 (2000) 145–151, https://doi.
catalyzing the nutrient uptake. The optimal trends were explained org/10.1016/S0269-7491(99)00118-9.
[25] J. Ruiz, P. Álvarez, Z. Arbib, C. Garrido, J. Barragán, J.A. Perales, Effect of nitrogen
through elaborate morphology analysis that showed 3% to strike the and phosphorus concentration on their removal kinetic in treated urban wastewater
balance of adsorption on alginate and nutrient access. by Chlorella vulgaris, Int. J. Phytoremediation 13 (2011) 884–896.
[26] P.K. Robinson, K.H. Goulding, A.L. Mak, M.D. Trevan, Factors affecting the growth
characteristics of alginate-entrapped Chlorella, Enzyme Microb. Technol. 8 (1986)
729–733, https://doi.org/10.1016/0141-0229(86)90160-2.
Acknowledgement [27] C. Yoo, S.-Y. Jun, J.-Y. Lee, C.-Y. Ahn, H.-M. Oh, Selection of microalgae for lipid
production under high levels carbon dioxide, Bioresour. Technol. 101 (2010)
CB highly acknowledges the Department of Science and Technology S71–S74, https://doi.org/10.1016/j.biortech.2009.03.030.
[28] F.Z. Mennaa, Z. Arbib, J.A. Perales, Urban wastewater treatment by seven species of
(DST) Govt. of India for providing financial support as well as project microalgae and an algal bloom: biomass production, N and P removal kinetics and
grant from INSPIRE Faculty award scheme (DST/INSPIRE/04/2014/ harvestability, Water Res. 83 (2015) 42–51.
002322). [29] A.C. Redfield, The biological control of chemical factors in the environment, Am.
Sci. 46 (1958) 230A–221.
[30] A. Beuckels, E. Smolders, K. Muylaert, Nitrogen availability influences phosphorus
removal in microalgae-based wastewater treatment, Water Res. 77 (2015) 98–106,
Appendix A. Supplementary data https://doi.org/10.1016/j.watres.2015.03.018.
[31] H. Cruz, P. Luckman, T. Seviour, W. Verstraete, B. Laycock, I. Pikaar, Rapid removal
Supplementary material related to this article can be found, in the of ammonium from domestic wastewater using polymer hydrogels, Sci. Rep. 8
(2018) 2912.
online version, at doi:https://doi.org/10.1016/j.bej.2019.107241.

11
S. Banerjee, et al. Biochemical Engineering Journal 149 (2019) 107241

[32] J. Wan, T. Tao, Y. Zhang, X. Liang, A. Zhou, C. Zhu, Phosphate adsorption on novel Chem. (2000).
hydrogel beads with interpenetrating network (IPN) structure in aqueous solutions: [45] S.Y. Reddy, T.C. Bruice, Mechanisms of ammonia activation and ammonium ion
kinetics, isotherms and regeneration, RSC Adv. 6 (2016) 23233–23241. inhibition of quinoprotein methanol dehydrogenase: a computational approach,
[33] W.S.W. Ngah, S. Fatinathan, Adsorption of Cu (II) ions in aqueous solution using Proc. Natl. Acad. Sci. 101 (2004) 15887–15892.
chitosan beads, chitosan–GLA beads and chitosan–alginate beads, Chem. Eng. J. [46] P.I. Nagy, P.W. Erhardt, On the interaction of aliphatic amines and ammonium ions
143 (2008) 62–72. with carboxylic acids in solution and in receptor pockets, J. Phys. Chem. B 116
[34] N.M. Mahmoodi, Equilibrium, kinetics, and thermodynamics of dye removal using (2012) 5425–5436.
alginate in binary systems, J. Chem. Eng. Data 56 (2011) 2802–2811. [47] M.C. Etter, Encoding and decoding hydrogen-bond patterns of organic compounds,
[35] L.Q. Wan, J. Jiang, D.E. Arnold, X.E. Guo, H.H. Lu, V.C. Mow, Calcium con- Acc. Chem. Res. 23 (1990) 120–126.
centration effects on the mechanical and biochemical properties of chondrocyte- [48] D. Das, S. Zhang, I. Noh, Synthesis and characterizations of alginate-α-tricalcium
alginate constructs, Cell. Mol. Bioeng. 1 (2008) 93–102. phosphate microparticle hybrid film with flexibility and high mechanical property
[36] K.Y. Foo, B.H. Hameed, Insights into the modeling of adsorption isotherm systems, as a biomaterial, Biomed. Mater. 13 (2018) 25008.
Chem. Eng. J. 156 (2010) 2–10. [49] A. Martinsen, I. Storrø, G. Skjårk‐Bræk, Alginate as immobilization material: III.
[37] X. Wang, S. Lü, C. Gao, C. Feng, X. Xu, X. Bai, N. Gao, J. Yang, M. Liu, L. Wu, Diffusional properties, Biotechnol. Bioeng. 39 (1992) 186–194.
Recovery of ammonium and phosphate from wastewater by wheat straw-based [50] N. Wang, G. Adams, L. Buttery, F.H. Falcone, S. Stolnik, Alginate encapsulation
amphoteric adsorbent and reusing as a multifunctional slow-release compound technology supports embryonic stem cells differentiation into insulin-producing
fertilizer, ACS Sustain. Chem. Eng. 4 (2016) 2068–2079. cells, J. Biotechnol. 144 (2009) 304–312.
[38] C. Shen, Y. Zhao, R. Liu, Y. Mao, D. Morgan, Adsorption of phosphorus with calcium [51] S. Mohanty, Y. Wu, N. Chakraborty, P. Mohanty, G. Ghosh, Impact of alginate
alginate beads containing drinking water treatment residual, Water Sci. Technol. 78 concentration on the viability, cryostorage, and angiogenic activity of encapsulated
(2018) 1980–1989. fibroblasts, Mater. Sci. Eng. C. 65 (2016) 269–277.
[39] H. Siwek, A. Bartkowiak, M. Włodarczyk, Adsorption of phosphates from aqueous [52] M.A. LeRoux, F. Guilak, L.A. Setton, Compressive and shear properties of alginate
solutions on Alginate/Goethite hydrogel composite, Water 11 (2019) 633. gel: effects of sodium ions and alginate concentration, J. Biomed. Mater. Res. An
[40] Y.-S. Ho, G. McKay, Pseudo-second order model for sorption processes, Process Off. J. Soc. Biomater. Japanese Soc. Biomater. Aust. Soc. Biomater. Korean Soc.
Biochem. 34 (1999) 451–465. Biomater. 47 (1999) 46–53.
[41] B.S. De, K.L. Wasewar, V.R. Dhongde, S.S. Madan, A.V. Gomase, Recovery of acrylic [53] G. Mujtaba, M. Rizwan, K. Lee, Removal of nutrients and COD from wastewater
acid using calcium peroxide nanoparticles: synthesis, characterisation, batch study, using symbiotic co-culture of bacterium Pseudomonas putida and immobilized
equilibrium, and kinetics, Chem. Biochem. Eng. Q. 32 (2018) 29–39. microalga Chlorella vulgaris, J. Ind. Eng. Chem. 49 (2017) 145–151.
[42] X. Liu, L. Zhang, Removal of phosphate anions using the modified chitosan beads: [54] Y. Shen, J. Gao, L. Li, Municipal wastewater treatment via co-immobilized micro-
adsorption kinetic, isotherm and mechanism studies, Powder Technol. 277 (2015) algal-bacterial symbiosis: microorganism growth and nutrients removal, Bioresour.
112–119. Technol. 243 (2017) 905–913.
[43] Y. Vijaya, S.R. Popuri, V.M. Boddu, A. Krishnaiah, Modified chitosan and calcium [55] J.-P. Hernandez, L.E. de-Bashan, D.J. Rodriguez, Y. Rodriguez, Y. Bashan, Growth
alginate biopolymer sorbents for removal of nickel (II) through adsorption, promotion of the freshwater microalga Chlorella vulgaris by the nitrogen-fixing,
Carbohydr. Polym. 72 (2008) 261–271. plant growth-promoting bacterium Bacillus pumilus from arid zone soils, Eur. J.
[44] J. Coates, Interpretation of infrared spectra, a practical approach, Encycl. Anal. Soil Biol. 45 (2009) 88–93.

12

You might also like