You are on page 1of 9

Environmental Science and Pollution Research

https://doi.org/10.1007/s11356-018-1394-7

RESEARCH ARTICLE

Identification of functional groups of Opuntia ficus-indica


involved in coagulation process after its active part extraction
Omar Bouaouine 1,2 & Isabelle Bourven 1 & Fouad Khalil 2 & Michel Baudu 1

Received: 19 July 2017 / Accepted: 25 January 2018


# Springer-Verlag GmbH Germany, part of Springer Nature 2018

Abstract
Opuntia ficus-indica that belongs to the Cactaceae family and is a member of Opuntia kind has received increasing research
interest for wastewater treatment by flocculation. The objectives of this study were (i) to provide more information regarding the
active constituents of Opuntia spp. and (ii) to improve the extracting and using conditions of the flocculant molecules for water
treatment. A classic approach by jar test experiments was used with raw and extracted material by solubilization and precipitation.
The surface properties of solid material were characterized by FTIR, SEM, zeta potential measurement, and surface titration. The
splitting based on the solubility of the material with pH and the titration of functional groups completed the method. The optimal
pH value for a coagulation–flocculation process using cactus solid material (CSM) was 10.0 and a processing rate of 35 mg L−1.
The alkaline pH of flocculation suggests an adsorption mechanism with bridging effect between particles by water-soluble
extracted molecules. To validate this mechanism, an extraction water was carried out at pH = 10 (optimum of flocculation)
and the solution was acidified (pH = 7) to allow precipitation of so considered active flocculant molecules. The strong flocculant
property of this extract was verified, and titration of this solution showed at least one specific pKa of 9.0 ± 0.6. This pKa
corresponds to phenol groups, which could be assigned to lignin and tannin.

Keywords Flocculation . Opuntia ficus-indica . Biotreatment . Extraction . Colloids

Introduction neutralization of colloidal particles and strongly a physical floc-


culation. A chemical flocculation (coagulation–flocculation) with
The production of drinking water from most raw water resources natural or synthetic polymers may supplement the process (Fu
typically involves a coagulation–flocculation process with a de- and Wang 2011, Jiangya et al. 2017). However, several draw-
stabilization of colloidal solution (coagulation) and the floccula- backs of using inorganic coagulants are frequently reported.
tion (mechanical or chemical) of solid particles to obtain According to the literature, possible residual aluminum in treated
suspended material (Lee et al. 2014a). The treatment of both water may have human health implications, and the large quan-
synthetic water and industrial wastewater normally entails the tity of metallic sludge produced which disposal itself is another
use of conventional chemical coagulants, including Al2 (SO4)3, problem (Bodlund et al. 2014; Freitas et al. 2015).
Fe (SO4)3, FeCl3, and other derivatives of these materials (e.g., Recent studies, conducted on the flocculation process during
poly ferric chloride and poly aluminum chloride). These inorgan- synthetic water and industrial wastewater treatment, have shown
ic materials lead to efficient sludge settlement through charge that natural organic polymers are preferentially used over inor-
ganic and synthetic polymers in the flocculation reaction (Miller
Responsible editor: Philippe Garrigues et al. 2008; Nharingo and Moyo 2016; Li et al. 2014). Natural
polymers are in fact more advantageous than inorganic com-
* Michel Baudu pounds by virtue of producing denser and more compact flakes,
michel.baudu@unilim.fr
which are biodegradable and less toxic than metallic compounds.
1
Groupement de Recherche Eau, Sol et Environnement (GRESE),
A strong attention is given to starch-based flocculants for water
University of Limoges, 123 avenue Albert Thomas, treatment with many possibilities of hemisynthesis for the im-
87060 Limoges, France provement of these flocculation performances (Wu et al., 2016).
2
Laboratory of Applied Chemistry (LCA), University Sidi Mohamed Natural polymers are used in lower doses and produce smaller
Ben Abdellah of Fez, Immouzer Road, BP 2202 Fez, Morocco amounts of sludge (Renault et al. 2009; Freitas et al. 2015; Zhang
Environ Sci Pollut Res

et al. 2006). Jones and Bridgeman (2016) have shown the effi- thorns were removed and dried at 80 °C for 24 h. The material
ciency of addition of purified seed extracts of Hibiscus vitifolius was crushed and sieved through a set of 0.5–1.0-mm diameter
Linn as coagulant aids. Currently, it is discovered that all plants sieves; this cactus powder was then stored at 4 °C. The dry-
being studied for flocculants production have mucilaginous con- ness and volatile dry weight (%VDW) of the prepared cactus
stituents and it is predicted that some of the active ingredients in solid material (CSM) were 81 ± 1 and 77 ± 1%, respectively.
the mucilage (polysaccharides) are responsible for the flocculat- In order to isolate the active components, this cactus material
ing property (galactomannan and galactan for Nirmali seeds; Yin was fractionated based on its solubility in water. This ap-
2010), but the most studied plant of the natural coagulant proach relied on the water extraction of soluble compounds
(Moringa oleifera) shows that the active agents are dimeric cat- (water-extracted material, WEM) from 1 g of dry weight
ionic proteins (Ndabigengesere et al. 1995). (DW) powdered cactus in 100 mL of deionized water at two
Opuntia ficus-indica has received greater attention over the distinct pH values (7 and 10). After stirring for 24 h, the cactus
past few years and has been extensively investigated for its extracts were filtered through a 20-μm fiberglass filter
pharmaceutical properties (anti-cancer effect, anti-diabetic ef- (Whatman). The filter paper was washed several times by
fect, neuroprotective… (Feugang et al. 2006; Kaur et al. 2012) distilled water. Then, a second filtration at 0.45 μm (cellulose
and physicochemical constituents (Ibanez et al. 1983; Miller acetate—Whatman) was carried out before drying the
et al. 2008)). Numerous studies have revealed that Opuntia retentate at 105 °C.
ficus-indica contains different biochemical compounds (e.g., Afterwards, the solution of WEM carried out at pH = 10
carbohydrates, polyphenols, proteins), reporting that some of (WEMpH 10) was acidified to pH 7. The soluble extract in
these constituents contain active ingredients with coagula- alkaline solution precipitated with the acidification. The
tion–flocculation capabilities (Yin 2010; Miller et al. 2008; precipitate was recovered by filtration (0.45 μm). The active
Torres et al. 2014). molecules isolated in the precipitate were resolubilized in a
To date, information is still sorely lacking as regards the basic medium and named Bprecipitate.^ The filtrate and the
overall coagulation capabilities of Opuntia spp. and its action precipitate from WEMpH 10 were conserved to quantify the
mechanism. According to the literature, cactus operates pri- residual and efficient flocculation property respectively.
marily through a flocculation mechanism (Miller et al. 2008),
a polymer-like material originating from the cactus species. Characterization of materials
However, all studies investigating the flocculating property of
bioflocculants focused solely on the flocculation process, in The zeta potential and particle size distribution of the col-
analyzing the effects of (i) flocculant dose, (ii) contact time, loidal solution were determined by a Malvern Master
and (iii) pH on flocculation efficiency (Betatache et al. 2014; Zetasizer 3000 device. The pH of CSM or WEM was set
Bouaouine et al. 2017; Aboulhassan et al. 2016). Until now, at various values (3 to 12) for 2 h, followed by a 0.45-μm
no study has investigated the relationship existing between filtration (cellulose acetate) before conducting the measure-
soluble chemical species and flocculation. ments. The Fourier transform infrared (FTIR) spectroscopy
In this context, the objectives of this laboratory-scale work analysis was performed on the CSM by applying an infrared
program were to obtain more information about the active con- spectrometer (Betatache et al. 2014; M. Tatzber et al. 2007)
stituent of Opuntia ficus-indica and enhance knowledge of co- in the range of 400–4000 cm−1. The CSM was prepared
agulation mechanism of synthetic water by this cactus. To under high pressure (12–15 tons) in the presence of KBr
achieve these objectives, flocculation process efficiency was powder (KBr > 99% Acros organics). The cactus solid ma-
studied, and water extractions have been done under various terial was observed under a scanning electron microscope
pH conditions. Materials (raw material and extractible) were (SEM) (QUANTA FEG 450, FEI, USA) at different magni-
characterized using zeta potential, colloidal particle size distri- fications ranging from (× 300 to × 40,000).
bution, and spectroscopy (i.e., UV-visible, infrared (IR)). Dry weight (DW in g L−1) of the material was performed
Moreover, the ionizable functions of these materials were ana- at 105 °C during 24 h and the volatile dry weight content
lyzed by means of acid–base titration. (VDW expressed in percent (g L−1/g L−1: VDW/DW)) at
550 °C during 2 h. The total organic carbon (TOC) of the
WEM was determined using a TOC-meter (Analytik Jena
Materials and methods CLD-TOC multi N/C 2100S). The UV-vis analyses of or-
ganic matter from the WEM were performed with a UV-
Preparation of cactus materials visible spectrophotometer (UV 2300).
The ionizable functions of CSM or WEM were derived
Opuntia ficus-indica that belongs to the Cactaceae family and manually using acid–base titration. Fifty milliliters of solu-
is a member of Opuntia kind was collected in a wild plantation tions containing 10 g L−1 of solid at different pH values was
near Fez (Morocco). The cactus pads were washed, and all placed at a constant temperature of 22 °C and adjusted every
Environ Sci Pollut Res

Table 1 Ionic balance of


synthetic water Cations (μmol) Anions (μmol)

K+ Na+ Mg2+ Ca2+ Si4+ Cl− HCO3− SO42− NO3−


Values 29 283 81 108 140 171 283 51 134

interval at a pH of around 2 by HCl (1 M). The titrations (Sigma Aldrich for both). The preparation was left for
were performed under a nitrogen atmosphere. The pH was 24 h without stirring. This time is necessary for the swell-
detected by a pH Meter GLP 22 CRISON equipped with a ing of the clay but hydration was incomplete and kaolinite
CRISON codigo electrode: 5221 (pH 0–12 ± 0.1), coupled to settling was observed. The supernatant was used as a sta-
a 25 ± 0.1-mL burette (Hirschmann EM Techcolor) filled with ble solution. This stock solution (Table 1) was diluted to
NaOH (0.1 M). achieve the desired turbidity between 300 and 350 neph-
Microscopy was performed with an optical microscope elometric turbidity units (NTU); the pH value equaled 6.5
(NIKON Eclipse LV1100POL) at a magnification of × 10 ± 0.4, and the absorbance at 254 nm was 1.2 ± 0.2. The
coupled with a CCD camera (Nikon Japan P-I). Image treat- unit employed herein was composed of six stirred reactors
ment was performed with the software NIS-Element F2.20. (VELP Scientifica JTL6), each containing 1 L of synthetic
The viscosity was measured with a RM200 Lamy rheom- water. The first step of the jar test consisted of stirring the
eter in WEM solution and in flocculated suspended solution stock solution at 250 rotations per minute (rpm) for
for a shear rate between 250 and 500 s−1 to avoid settling of roughly 10 min while the flocculant was gradually added
solid. to the reactor. During the second step, the speed was re-
duced to 50 rpm for 30 min. The last step entailed sedi-
Jar test experiments mentation of the flocs that had formed over a 2-h period.
The efficiency of the cactus material was measured deter-
A synthetic water was introduced to draw a robust comparison mining the removal of turbidity or absorbance at 254 nm in
among all experiments. An ionic strength was established in synthetic water. The residual turbidity of suspended solids was
order to mimic natural water by salt solubilization (Table 1). then measured by a turbidimeter (Hanna HI 88713) and
The synthetic water was completed by thoroughly expressed in NTU in accordance with the standard nephelom-
mixing kaolin (1 g L−1) and humic acid (10 mg L−1) etry method. The absorbance of soluble substances was

Fig. 1 Observation by SEM of


cactus solid material (CSM)
Environ Sci Pollut Res

Fig. 2 Variation of zeta potential Colloidal granulometry CSM


350 0
and colloidal granulometry at Colloidal granulometry synthetic water
different values of pH for 10 g L−1 300
zeta potential CSM
-5
of cactus solid material (CSM)

Colloidal Granulometry (nm)


and synthetic water. *n = 3 (inde-
250 -10

Zeta Potential (mV)


pendent assays)
200 -15

150 -20

100 -25

50 -30

0 -35
3 4 5 6 7 8 9 10 11 12
pH

measured at a wavelength of 254 nm (Abs254 nm), which is The removal of turbidity or Abs254 nm was calculated using the
capable of absorbing aromatic groups like humic substances. following formula:

Initial Turb or Abs 254 nm−Residual Turb or Abs 254 nm


Removal ð%Þ ¼  100
Initial Turb or Abs 254 nm

Results and discussion pH 3 to − 30 mV at pH 12, as does the synthetic water from


− 10 to −37 mV, respectively. In parallel, the colloidal par-
Characterization of materials ticle size of CSM increases from 110 to 289 nm as pH from 3
to 10 and subsequently decreases (216 nm at pH 12). The
The observation of CSM by SEM (Fig. 1) showed irregular colloidal size variation can be explained by the presence of
colloidal particles with various sizes included between more organic monomers undergoing polymerization depending
of 1 mm and few hundred nanometers. on the number and characteristic variations of ionized
Figure 2 presents the zeta potential variation for the col- groups. Maximum polymerization was detected at pH 10.
loidal solution of CSM and synthetic water, as well as the In contrast (Fig. 3), FTIR has been implemented to identify
particle size distribution at different pH values (from 3 to the presence of functional groups on the CSM; a peak is found
12) for CSM. For CSM and the synthetic water, regardless between 3200 and 3500 cm−1, which could be correlated with
of the pH values, zeta potential is negative, and the synthetic the existence of carboxylic acids (Betatache et al. 2014). Band
water is more negative than CSM. For CSM, the evolution absorption between 2800 and 3000 cm−1 may be explained by
of zeta potential displays a gradual decrease from − 7 mV at the presence of CH3 and CH3–O. The vibrations of strips at

Fig. 3 FTIR spectrum for 2.5 mg 250


of cactus solid material (CSM)

200 1519
1320
Transmission (%)

150
1050
1430
100
1620

50 3460

0
500 1000 1500 2000 2500 3000 3500 4000
Wavelenght (cm-1)
Environ Sci Pollut Res

Fig. 4 Impact of the pH on the pH=3 pH=7 pH=10


efficiency of the coagulation– a
flocculation process (fixed dose
of 100 mg of cactus solid material
(CSM) for 1 L of synthetic water).
a Optic microscopic observations
of sludge formed after coagula-
tion–flocculation (pH 3, 7, and
10). b Viscosity (Pa s–1) of sludge
formed after coagulation–floccu-
lation (pH 3, 7, and 10) at 20 °C. c 180 µm 180 µm 180 µm
Variation of removal of turbidity Sludge (diluon 1/1 Sludge (diluon 1/1) Sludge (diluon ¼)
(%) for synthetic water + CSM
b Viscosity= (2.0 ± 0.1) 10-3 Pa.s-1 Viscosity= (4.8± 0.1) 10-3 Pa.s- Viscosity= (6.8± 0.1) 10-3 Pa.s-1
and synthetic water alone. *n = 3 1
(independent assays)
c
Synthetic water Synthetic water + CSM
100
90
80
Removal turbidity%

70
60
50
40
30
20
10
0
3 4 5 6 7 8 9 10 11 12
pH

1620 and 1430 cm−1 indicate the existence of groups C=O and and phosphoric groups, respectively (Cox et al. 1999). pKa4
phenol groups (M. Tatzber et al. 2007), respectively. In addi- (8.6–9.9) may be assigned to the phenol function (Martin
tion, the peaks at 1320 and 1050 cm−1 prove the appearance of Ragnar et al. 2000) from lignin (Herrero-Martínez et al.
–COOH groups, aromatic proteins, phosphoric groups, and 2005) or tannin, such as catechin (Tungjai et al. 2008). This
polysaccharides (Bouatay and Mhenni 2014; Thomas and pKa might also be assigned to some free amino acids. pKa5
Chittenden 1964). The FTIR analysis confirms the presence (11–12.2) may also be ascribed to arginine amino acids
of ionizable groups like hydroxyl, carboxyl, or phosphoric (Sjöholm and Stigbrand 1974). Table 4 summarizes, based
groups. on the literature, the possible ionizable groups, along with
Accordingly, acid–base titrations of CSM have been con- possible associated molecules and their correspondence to
ducted in order to identify these groups (Table 3), with five functional groups and molecules identified in FTIR for the
pKa being detected. The first sites identified with pKa1 (i.e., CSM.
2.1–2.9) likely correspond to the –COOH group from free The characteristics of water-extracted material (WEM), at
amino acids (Včeláková et al. 2004). pKa2 (4.5–5.5) and pH values of both 7 (WEMpH 7) and 10 (WEMpH 10), are
pKa3 (6–7) are most likely associated with carboxyl groups presented in Table 2.

Table 2 Characteristics of
WEMpH 7 and WEMpH 10 pH Extraction Colloidal VDW TOC g (C)/g Abs210 nm for Abs280 nm for
extraction yield granulometry % of dry weight 1 g dry weight 1 g dry weight
(DW/DW) nm
%

7, 0 34 ± 2 160 ± 2 16 ± 1 0.045 ± 0.001 0.53 ± 0.03 0.08 ± 0.01


10, 0 51 ± 2 208 ± 3 31 ± 1 0.16 ± 0.01 0.9 ± 0.1 0.13 ± 0.02

WEM water extract material


*n = 3 (independent assays)
Environ Sci Pollut Res

Fig. 5 Turbidity and Turbidity Abs 254 nm


100
absorbance254 nm removals upon
dose of cactus solid material 90
(CSM) at pH 10. *n = 3 (inde- 80
pendent assays) 70

Removal %
60
50
40
30
20
10
0
0 10 20 30 40 50 60
Concentration of CSM at mg L-1

Extraction yield increases with pH from 34 to 51% of dry function of pH value: turbidity removal, microscopic observa-
weight (factor = 1.5). Concerning the change in global organic tions of the sludge after coagulation–flocculation, and viscos-
matter of the WEM fraction, the absorbance at 210 and 280 nm ity observations of the sludge after coagulation–flocculation
(Abs210 nm and Abs280 nm for 1 g dry weight) increases (by a (Fig. 4). In Fig. 4C, results show the evolution of turbidity
factor > 1.5) with pH; this trend is also observed for TOC removal versus pH for the synthetic water alone or in a solu-
content (factor > 3). At pH 10, we have therefore obtained more tion with CSM.
organic material than at pH 7, with an organic component Three ranges can be observed: in acidic condition (from pH 3
slightly differentiated by absorbance (a more conjugated bond to 5), the removal turbidity is great but decreases with pH value
in WEMpH 10). The viscosity of WEM appeared not signifi- (63 to 40%, respectively). In the second range, from pH 5 to 6,
cantly different from pure water. An increase in particle size the turbidity decrease is weakest. Beyond pH 6 and when CSM
(from 160 to 208 nm) was also noticed from pH 7 to pH 10; is present in solution, removal turbidity and the settling increase,
two reasons can be cited: (i) modification of polymer confor- reaching a maximum of 92%. Lastly, from pH 10 to 12, removal
mation when considering possible new ionized groups and (ii) dips slightly from 92 to 78%. In the absence of CSM, the turbid-
the characteristic variation of organic WEM components based ity of synthetic water is slightly reduced. These results confirm
on pH of extracted molecules in alkaline conditions. that the optimal flocculation pH equals 10.0 and the minimum
occurs at around 6 (Miller et al. 2008; Zhang et al. 2006; Torres
Coagulation–flocculation efficiency et al. 2014). As shown in Fig. 2, as pH decreases, zeta potential
increases for the synthetic water (up to − 10 mV) and lies close to
In order to identify optimal condition pH for the flocculation, neutrality at pH 3. Colloid form of synthetic water can thus
the coagulant dose was set at 100 mg of CSM for 1 L of coagulate in the first range of acidic condition. Between pH 6
synthetic water and three parameters were measured as a and 7, the colloidal suspension of synthetic water is relatively

Fig. 6 Variation of removal of WEM pH 10 WEM pH 7


turbidity at different doses of dry 100 Residual solid pH 7 Residual solid pH 10
weight WEM (water-extracted 90
material) at pH = 7 and 10 for
synthetic water at pH 10. *n = 3 80
(independent assays) 70
Removal tubidity %

60
50
40
30
20
10
0
0 10 20 30 40 50 60 70 80
Dose of material (mg L-1 of WEM or residual)
Environ Sci Pollut Res

100
90 Given that regardless of pH value, the zeta potential of
80 colloids for both CSM and synthetic water is negative (Fig.
Removal turbidity%

70 2), then colloidal neutralization is infeasible with CSM. This


60
50
statement suggests that colloids and suspended CSM do not
40 operate through a colloidal neutralization of particles in the
30 synthetic water, but instead through an adsorption mechanism
20
10
and bridging between particles. Furthermore, the cactus at
0 pH 10.0 exhibits a negatively charged backbone (Fig. 2) due
WEM pH 10 WEM pH 7 Precipitate Filtrate to carboxyl and phenol groups (Fig. 3), which have allowed
Fig. 7 Comparison of removal of turbidity between various all cactus the polymer molecules to extend into solution and produce
preparations at an optimal dose of 35 mg L−1. *n = 3 (independent assays) loops and tails to promote the bridging of flocs, like other
bioflocculants: tannin and anionic cellulose (Lee et al. 2014b).
stable and more and more with the pH increase (with or without The optimal activity of CSM at pH = 10 may in fact be
CSM). On the other hand, a flocculation phenomenon is ob- correlated with quantitative aspects as well as with the nature
served in alkaline condition when CSM is introduced. of active molecules extracted and transferred into solution. To
Microscopic observation (Fig. 4A) and viscosity measurement better comprehend the pH influence, a water extraction for
(Fig. 4B) at pH values (3, 7, and 10) confirm the three two pH values was performed, one at the maximum floccula-
mechanisms: tion efficiency (pH = 10) the other at pH = 7, where floccula-
– mechanism I around pH 3: mechanical coagulation of tion appears to be minimal (Fig. 4). The flocculent activity of
particles in synthetic water: floc from sludge measuring the water-extracted material at pH 7 (WEMpH 7) and pH 10
between 10 and 30 μm (WEMpH 10) was investigated in the subsequent stage of this
– mechanism II around pH 7: stabilized colloidal suspen- study.
sion with poor concentration of suspended matter (parti- In order to confirm the efficiency of WEM as a flocculant,
cle size of floc from sludge less than 10 μm) Figure 6 examines the variation in turbidity removal with the
– mechanism III around pH 10: coagulation–flocculation WEM dry weight dose for both extraction pH values
by adsorption with particle size of floc from sludge rang- (WEMpH 7 and WEMpH 10). The coagulant effect has also
ing between 70 and 400 μm been tested for the residual solid after extraction at pH 7 and
10. It was observed that a 35-mg L−1 dry weight dose allows
Figure 5 depicts the variation in turbidity and Abs254 nm re- optimal removal with all extracted or residual solids. The best
movals as a function of CSM dose (between 0 and 60 mg L−1) at turbidity removal value was obtained for WEMpH 10, with a
pH 10.0. The highest turbidity and absorbance removal rate equal removal rate of 90%, whereas WEMpH 7 displayed a removal
96 and 83%, respectively, for 35 mg L−1 of CSM. The same rate of 48% for both. It was also found that the residual solids
tendency was observed for turbidity and Abs254 nm because hu- after extraction at pH 7 or 10 exhibit similar low efficiency (20
mic acids were absorbed into flocs during coagulation–floccula- to 30%) for turbidity removal. This settling with rare floccu-
tion process. These optimal conditions are comparable to results lant molecules with the residual solid can be associated to a
found by Miller et al. (2008), who worked on a synthetic water of sweep flocculation mechanism with the high concentration of
kaolin. In contrast, Zhang et al. (2006) achieved a higher turbidity colloids in the solution. Moreover, the removal of CSM (Fig.
removal with a dose of 60 mg L−1. Many studies have demon- 4) and WEMpH 10 is nearly identical (96 vs. 91% for turbidity
strated the efficiency of cactus in the field of wastewater treat- removal). These results prove that flocculant molecules are
ment (Bouaouine et al. 2017; Betatache et al. 2014; Nharingo mainly extracted at pH = 10, while the role of the residual
and Moyo 2016; Torres et al. 2014). solid phase during treatment with CSM remains very weak.

Table 3 pKas values for all cactus


preparations pKas CSM (cactus solid material) WEMpH 10 WEMpH 7 Precipitate Filtrate

pKa1 2.3 ± 0.6 2.9 ± 0.5 2.3 ± 0.6


pKa2 5.5 ± 0.6 6.0 ± 0.2 4.8 ± 0.4 4.7 ± 0.4 4.8 ± 0.5
pKa3 7.0 ± 0.6
pKa4 9.0 ± 0.6 9.0 ± 0.6 9.0 ± 0.6
pKa5 12.0 ± 0.6 12.0 ± 0.6 11.0 ± 0.5

WEM water extract material


*n = 2 (independent assays)
Environ Sci Pollut Res

Table 4 FTIR bands and ionizable groups of different pKa detected

Values of pKa Possible References (pKa) FTIR bands and correspondence References (FTIR) Possibility of the molecules
ionizable groups associated with the ionizable group

2.3–2.9 Carboxyl [21] 1050 cm−1 [19] Tryptophan, lysine, and glycine
4.5–5.5 Carboxyl [22] 3200–3500 cm−1 [15] Uronic acids
6–7 Phosphoric [22] 1320 cm−1 and 1600–2725 cm−1 [20] Phosphoric acids
8.6–9.9 Phenol [23] and [24] 1430–1620 cm−1 [18] Tannin (whose catechin) and
lignin and aromatic amino acid
11–12.2 Amine [26] 1499–1602 cm−1 [30] Tyrosine

To isolate the more active molecules contained in the solution WEMpH 10 and the precipitate are quite possibly tannins, like
at pH = 10, the WEMpH 10 (optimum flocculation) was acidified catechin and lignin.
at pH = 7, resulting in a precipitation. Turbidity removal at opti-
mal dose (i.e., 35 mg L−1) for a synthetic water at pH 10.0 is
compared (see Fig. 7); among BWEMpH 7^ and BWEMpH 10,^ a Conclusion
pellet from WEMpH 10 precipitated at pH 7 (precipitate), and the
WEMpH 10 supernatant precipitated at pH 7 (filtrate). The precip- The study explores that Opuntia ficus-indica can efficiently
itate contains (nearly) all of the active flocculant molecules com- remove colloid in water solution by a coagulation–floccula-
pared to WEMpH 10, with a turbidity removal rate equals to 91 ± tion mechanism with an optimal pH value of 10.0. The zeta
1% and 93 ± 2%, respectively. It was also determined that the potential has suggested that this material operates through a
turbidity removal rate of WEMpH 7 is actually greater than that of flocculation mechanism, with the adsorption of water-soluble
the filtrate, reaching 50 and 16%, respectively. biomolecules and a bridging between particles. A large quan-
To identify the components involved in the defined floccu- tity of material may be extracted in water at pH 10 and isolated
lation process, let us now compare the pKa of WEMpH 7, by precipitation at pH = 7. This fraction shows similar effi-
WEMpH 10, and both the precipitate and filtrate; the pKas ciency when compared to the raw material. The FTIR results
values of the various ionizable groups are listed in Table 3. and the identification of a specific function of water extractible
Among the five pKas values from CSM, the four cactus prep- material at pH 10 (pKa of 9.0 ± 0.6) suggest the presence of a
arations reveal values of less than 5 pKa. phenol group, possibly originating from lignin and tannin (es-
WEMpH 10 contains the same functional groups as CSM, pecially catechin). Bioflocculation properties of such mole-
except for phosphoryl groups (pKa3), which explains the im- cules stemming from plants had not been revealed yet.
proved extraction removal rate at pH 10. However, the most
striking result is that the precipitate contains just two pKa Funding information The authors would like to thank the bilateral
(pKa2 and 4), thus involving the more active ionizable groups. Toubkal PHC program for its financial support.
In sum, should the active constituents be ionizable, then it may
References
be confirmed that the major activity of ionizable groups is
associated with pKa4 since pKa2 is common to the filtrate
Aboulhassan MA, Souabi S, Yaacoubi A, Baudu M (2016) Coagulation
and precipitate, except that the turbidity removal rate of the efficacy of a tannin coagulant agent compared to metal salts for paint
filtrate is merely 16%. manufacturing wastewater treatment. Desalination Water Treat
pKa4, which is common to CSM, WEMpH 10, and the pre- 57(41):19199–19205. https://doi.org/10.1080/19443994.2015.
cipitate, is present in molecular form from the active com- 1101016
Betatache H, Aouabed A, Drouiche N, Lounici H (2014)
pounds. According to the literature (Table 4) , pKa4 (8.6– Conditioning of sewage sludge by prickly pear cactus
9.9) could correspond to phenol groups associated with lignin (Opuntia ficus Indica) juice. Ecolog Eng 70:465–469.
(Gross and Seybold 2001). Moreover, compared to groups https://doi.org/10.1016/j.ecoleng.2014.06.031
identified in FTIR (Fig. 3), we have observed that all of the Bodlund I, Pavankumar AR, Chelliah R, Kasi S, Sankaran K, Rajarao GK
(2014) Coagulant proteins identified in Mustard: a potential water
investigated CSM spectra shared certain spectral similarities
treatment agent. Inter J Environ Sci Techn 11(4):873–880. https://
with phenol groups. A band, due to the presence of a phenol doi.org/10.1007/s13762-013-0282-4
stretch, has been observed between 1430 and 1620 cm−1, es- Bouaouine O, Baudu M, Khalil F, Chtioui H, Zaitan H (2017)
pecially at 1519 cm−1, which could correspond to tannin (in- Comparative study between Moroccan cactus and chemicals coag-
cluding catechin) (Robb et al. 2002) or free aromatic amino ulants for textile effluent treatment. J Mater Environ Sci 8:2687–
2693
acids (Venyaminov and Kalnin 1990). In considering the floc-
Bouatay F, Mheni MF (2014) Use of the cactus cladodes mucilage
culation mechanism, the most active molecules in both (Opuntia Ficus Indica) as an eco-friendly flocculants: process
Environ Sci Pollut Res

development and optimization using stastical analysis. Int J Environ Ndabigengesere AI, Narasiah KS, Talbot BG (1995) Active agents and
Res 8:1295–1308 mechanism of coagulation of turbid waters using Moringa oleifera.
Cox JS, Smith DS, Warren LA, Ferris FG (1999) Characterizing hetero- Water Res 29(2):703–710. https://doi.org/10.1016/0043-1354(94)
geneous bacterial surface functional groups using discrete affinity 00161-Y
spectra for proton binding. Environ Sci Techn 33(24):4514–4521. Nharingo T, Moyo M (2016) Application of Opuntia ficus-indica in bio-
https://doi.org/10.1021/es990627l remediation of wastewaters. A critical review. J Environ Manag 166:
Feugang JM, Konarski P, Zou D, Stintzing FC, Zou C (2006) Nutritional 55–72. https://doi.org/10.1016/j.jenvman.2015.10.005
and medicinal use of Cactus pear (Opuntia spp.) cladodes and fruits. Ragnar M, Lindgren CT, Nilvebrant N-O (2000) pK a-values of guaiacyl
Front Biosci 11(1):2574–2589. https://doi.org/10.2741/1992 and syringyl phenols related to lignin. J Wood Chem Techn 20(3):
Freitas TKFS, Oliveira VM, de Souza MTF, Geraldino HCL, 277–305. https://doi.org/10.1080/02773810009349637
Almeida VC, Fávaro SL, Garcia JC (2015) Optimization of Renault F, Sancey B, Badot PM, Crini G (2009) Chitosan for coagulation/
coagulation-flocculation process for treatment of industrial flocculation processes—an eco-friendly approach. Europ Polym J
textile wastewater using okra (A. esculentus) mucilage as 45(5):1337–1348. https://doi.org/10.1016/j.eurpolymj.2008.12.027
natural coagulant. Indus Crops Prod 76:538–544. https://doi. Robb CS, Geldart SE, Seelenbinder JA, Brown PR (2002) Analysis of
org/10.1016/j.indcrop.2015.06.027 green tea constituents by HPLC-FTIR. J Liq Chrom Related Techn
Fu F, Wang Q (2011) Removal of heavy metal ions from wastewaters: a 25(5):787–801. https://doi.org/10.1081/JLC-120003036
review. J Environ Manag 92(3):407–418. https://doi.org/10.1016/j. Sjöholm I, Stigbrand T (1974) Circular dichroism studies on the copper
jenvman.2010.11.011 ligand structure of umecyanin by spectropolarimetric titration.
Gross KC, Seybold PG (2001) Substituent effects on the physical prop- Biochimica et Biophysica Acta (BBA) – Prot Structure 371(2):
erties and pKa of phenol. International J Quantum Chem 85(4-5): 408–416. https://doi.org/10.1016/0005-2795(74)90037-3
569–579. https://doi.org/10.1002/qua.1525 Tatzber M, Stemmer M, Spiegel H, Katzlberger C, Haberhauer G,
Gerzabek MH (2007) An alternative method to measure carbonate
Herrero-Martínez JM, Sanmartin M, Rosés M, Bosch E, Ràfols C (2005)
in soils by FT-IR spectroscopy. Environ Chem Letters 5(1):9–12.
Determination of dissociation constants of flavonoids by capillary
https://doi.org/10.1007/s10311-006-0079-5
electrophoresis. Electrophoresis 26(10):1886–1895. https://doi.org/
Thomas LC, Chittenden RA (1964) Characteristic infrared absorption
10.1002/elps.200410258
frequencies of organophosphorus compounds—I. The phosphoryl
Ibanez-Camacho R, Meckes-Lozoya M, Mellado-Campos V (1983) The
(P=O) group. Spectrochim Acta 20(3):467–487. https://doi.org/10.
hypoglucemic effect of Opuntia streptacantha studied in different
1016/0371-1951(64)80043-6
animal experimental models. J Ethnopharmacol 7(2):175–181.
Torres LG, Cadena G, Carpinteyro-Urbán S, Corzo LJ (2014) New
https://doi.org/10.1016/0378-8741(83)90019-3
galactomannans and mucilages with coagulant-flocculant activity
Jiangya M, Fu K, Fu XGQ, Ding L, Shi J, Zhu G, Zhang X, Zhang S,
for an environment friendly treatment of wastewaters. Current Adv
Jiang L (2017) Flocculation properties and kinetic investigation of
Environ Sci 2:52–58
polyacrylamide with different cationic monomer content for high
Tungjai M, Poompimon W, Chatchanok L, Kothan S (2008)
turbid water purification. Sep Purif Techn 182:134–143
Spectrophotometric characterization of behavior and the predomi-
Jones AN, Bridgeman J (2016) Investigating the characteristic strength of nant species of flavonoids in physiological buffer: determination of
flocs formed from crude and purified Hibiscus extracts in water solubility, lipophilicity and anticancer efficacy. The Open Drug
treatment. Water Res 103:21–29. https://doi.org/10.1016/j.watres. Delivery J 2(1):10–19. https://doi.org/10.2174/
2016.07.019 1874126600802010010
Kaur M, Kaur A, Sharma R (2012) Pharmacological actions of Opuntia Včeláková K, Zusková I, Kenndler E, Gaš B (2004) Determination of
ficus indica: a review. J of applied pharma Sci 07:15–18 cationic mobilities and pKa values of 22 amino acids by capillary
Lee CS, Chong MF, Robinson J, Binner E (2014a) A review on devel- zone electrophoresis. Electrophoresis 25(2):309–317. https://doi.
opment and application of plant-based bioflocculants and grafted org/10.1002/elps.200305751
bioflocculants. Indus Eng Chem Res 53(48):18357–18369. https:// Venyaminov SY, Kalnin NN (1990) Quantitative IR spectrophotometry
doi.org/10.1021/ie5034045 of peptide compounds in water (H2O) solutions. I. Spectral param-
Lee CS, Robinson J, Chong MF (2014b) A review on application of eters of amino acid residue absorption bands. Biopolymers 30(13-
flocculants in wastewater treatment. Proc Safety Environ Protect 14):1243–1257. https://doi.org/10.1002/bip.360301309
92(6):489–508. https://doi.org/10.1016/j.psep.2014.04.010 Wu H, Liu Z, Li A (2016) Evaluation of chain architectures and charge
Li Y, Li Q, Hao D, Hu Z, Song D, Yang M (2014) Characterization and properties of various starch-based flocculants for flocculation of
flocculation mechanism of an alkali-activated polysaccharide floc- humic acid from water. Water Res 96:126–135. https://doi.org/10.
culant from Arthrobacter sp. B4. Bioresour Technol 170:574–577. 1016/j.watres.2016.03.055
https://doi.org/10.1016/j.biortech.2014.07.112 Yin CY (2010) Emerging usage of plant-based coagulants for water and
Miller SM, Fugate EJ, Craver VO, Smith JA, Zimmerman JB (2008) wastewater treatment. Process Biochem 45(9):1437–1444. https://
Toward understanding the efficacy and mechanism of Opuntia doi.org/10.1016/j.procbio.2010.05.030
spp. as a natural coagulant for potential application in water treat- Zhang J, Zhang F, Luo Y, Yang H (2006) A preliminary study on cactus as
ment. Environ Sci Technol 42(12):4274–4279. https://doi.org/10. coagulant in water treatment. Process Biochem 41(3):730–733.
1021/es7025054 https://doi.org/10.1016/j.procbio.2005.08.016

You might also like