You are on page 1of 6

Ecotoxicology and Environmental Safety 190 (2020) 110140

Contents lists available at ScienceDirect

Ecotoxicology and Environmental Safety


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

D151 resin preloaded with Fe3+ as a salt resistant adsorbent for glyphosate T
from water in the presence 16% NaCl
Guqing Xiao∗, Qiudong Meng
College of Materials and Chemical Engineering, Hunan City University, Yiyang, 413000, PR China

ARTICLE INFO ABSTRACT

Keywords: D151 resin preloaded with Fe3+ [denoted as R–Fe3+] was to investigate R–Fe3+ as an adsorbent for glyphosate
D151 resin preloaded with Fe3+ from water in the presence high concentration of salt. The adsorption mechanism revealed the coordination of
Salt resistance Fe3+ inside R–Fe3+ with O atoms of P–O and N atoms in glyphosate molecule. The adsorption capacity of
Glyphosate glyphosate by R–Fe3+ was much larger than that of D151 resin preloaded with Ni2+, Cu2+, Na+ and H+. Even
Coordination
in glyphosate solutions containing 16% NaCl, R–Fe3+ showed the constant adsorption capacity of glyphosate.
Adsorption
The result provided the first evidence of R–Fe3+ as a salt resistant adsorbent for glyphosate. The adsorption
capacity of glyphosate was the maximum at pH 3.35. The adsorption thermodynamics showed that the ad-
sorption of glyphosate by R–Fe3+ was the ligand exchange of glyphosate and water. The maximum coordination
ratio of glyphosate to Fe3+ inside R–Fe3+ was 1:1. The maximum adsorption capacity of glyphosate by R–Fe3+
was up to 481.85 mg/g, which is much higher than that of other reported adsorbents in the presence 16% NaCl.
2 mol/L NaOH, 2 mol/L H2SO4 and 2 mol/L Fe2(SO4)3 could all be used to achieve over 97% regeneration of
R–Fe3+.

1. Introduction containing 10% glyphosate for sale. The disadvantage of the method is
that the glyphosate reagent contains a large amount of salt, which can
Glyphosate, as a non-selective herbicide, is the most commonly used easily lead to soil salinization. Of course, the method has been banned
herbicide in the world (Skeff et al., 2018) [1]. About 82.6 × 104 tons of in China (Shen et al., 2013). The evaporation solvent is another tradi-
glyphosate have been used only in 2014 (Zhang et al., 2018). Glypho- tional method to recover glyphosate from the two mother liquors in
sate can be detected in surface water and soil because of its widespread China (Shen et al., 2013), which consumes a great deal of energy. There
use (Zhou et al., 2017; Arroyave et al., 2016). The US Environmental are other methods such as electrochemical oxidation (Rubí-Juárez
Protection Agency (US EPA) has set the maximum concentration of et al., 2016), biodegradation (Wang et al., 2016) and visible light cat-
glyphosate in water at 0.28 mg/L (Lan et al., 2016). Once glyphosate alytic degradation (Huo et al., 2017), these treatments result in the
enters the human body through the food chain and water, it can in- waste of glyphosate. Ore adsorption is not suitable for large scale gly-
terfere with the function of the nervous and endocrine systems (Zavareh phosate wastewater due to its defects of low efficiency (Yang et al.,
et al., 2018). Glyphosate can also cause genetic mutation (Jiang et al., 2018). The recovery of glyphosate in the presence high concentration of
2018) and cancer (Yamaguchi et al., 2016). There are two methods to salt has been a scientific problem for many years.
produce glyphosate in the industry. The mother liquor used to produce -NH- groups exhibit alkalinity, while the groups [-COOH and (HO)2-
glyphosate with glycine contains about 1.5% glyphosate and 13–15% PO-] exhibit acidity. So glyphosate is an amphoteric substance.
NaCl (Liu et al., 2013; Shen et al., 2013), while the mother liquor of Glyphosate can exhibit four levels of ionization constants (Fiorilli et al.,
glyphosate by IDA method contains about 2.0% glyphosate and 4.5% 2017). Glyphosate, as a weak electrolyte, is mainly in the first ioniza-
NaCl (Shen et al., 2013). Direct discharge of the two mother liquors can tion equilibrium without adding acid and base. The first ionization
cause severe environmental pollution and the loss of glyphosate (Hu equation of glyphosate (pKa1) is described in Scheme 1. Glyphosate
et al., 2011). Obviously, the recovery of glyphosate from the two mo- exhibits a high solubility in water and a small solubility in organic
ther liquors has important environmental and economic benefits. solvents (Mata et al., 2014; Divisekara et al., 2018). Therefore, it is
One way to treat glyphosate mother liquor is to neutralize it, then difficult to adsorb glyphosate from aqueous solution by polystyrene
evaporate and concentrate to produce glyphosate pesticide reagent because of the weak hydrophobic action as the adsorption driving force.


Corresponding author.
E-mail address: xiaoguqing2005@163.com (G. Xiao).

https://doi.org/10.1016/j.ecoenv.2019.110140
Received 12 November 2019; Received in revised form 25 December 2019; Accepted 26 December 2019
Available online 03 January 2020
0147-6513/ © 2019 Elsevier Inc. All rights reserved.
G. Xiao and Q. Meng Ecotoxicology and Environmental Safety 190 (2020) 110140

Scheme 1. Resin transformation and the adsorption mechanism of glyphosate by R–Fe3+.

–NH– groups of glyphosate turns into NH2+ in the first ionization concentrations (Ce mg/mL) of glyphosate in the aqueous phase were
equilibrium. The force of the weak acid groups on NH2+ is weak. determined with UV-3010 spectrophotometer at 242 nm. According to
Hence the weak acid resin exhibits very small adsorption capacity for Chinese national standard GB12686-2004, glyphosate reacted with
glyphosate. sodium nitrite to turn into glyphosate nitroso derivatives under the
Our team is inspired by the modification of zeolite 4A as an ad- acidic condition (Song et al., 2013). Glyphosate nitroso derivatives had
sorbent for glyphosate (reported in the literature of Ecotoxicol. Environ. the maximum absorption peak at the wavelength of 242 nm (Song
Saf. 2018, 155, 1–8). D151 resin preloaded with Fe3+ [denoted as et al., 2013). The adsorption capacity of glyphosate q(mg/g) was cal-
R–Fe3+] as a new adsorbent for glyphosate from water in the presence culated according to (1).
16% NaCl is carried out in the work. It's a surprise that the maximum
q = 25(Ci-Ce)/W (1)
adsorption capacity of glyphosate by R–Fe3+ is much higher than that
of other reported adsorbents in the presence 16% NaCl.

3. Results and discussion


2. Experimental
3.1. Characterization of the resins
2.1. Material and chemicals
1.0 g of R–Fe3+ are added into a cone-shaped flask with 25 mL of
D301 resin, D151 resin and 330 resin were supplied by Anhui 2.0 mol/L H2SO4, then the flask is continuously shaken for 48 h at
Sanxing resin technology co., Ltd (Anhui Province, China). KBr, NaNO2, 298 K so that R–Fe3+ is converted into R–H+ to release of Fe3+ com-
FeCl3, CuCl2 and NiCl2 were analytical reagents. Ultra-pure water was pletely. In the transformation of the two resins, the color of the two
used to prepare various solutions in the work. The standard glyphosate resins changes from brown to white. Fe3+ in desorption solution is
was purchased from SigmaeAldrich (Steinheim, Germany). determined at 478 nm by KSCN coloring. Fe3+ inside R–Fe3+ is de-
termined to be 2.85 mmol/g. Zero charge point of the resin (denoted as
2.2. D151 resin preloaded with Fe3+ pHPZC) is pH value at which the net charge on the resin surface is equal
to zero. Zero charge point of resin is determined by inert electrolyte
As the resin transformation is described in Scheme 1, 30 g of D151 titration (Lazarević et al., 2007). pH values of glyphosate solution be-
resin were rinsed with 2 mol/L HCl, then rinsed with deionized water to fore and after adsorption are denoted as pHi and pHf respectively. The
the neutral and D151-COOH resin [denoted as R–H+] was obtained. difference of pHi and pHf is denoted as ΔpH. ΔpH against pHi is plotted
R–H+ was rinsed with 2 mol/L NaOH. The resin was rinsed with as shown in Fig. S1. Zero point of ΔpH corresponds to the zero charge
deionized water to neutral further and D151-COONa resin [denoted as point of the resin. Zero charge points of R–Na+, R–Fe3+ and R–H+ are
R–Na+] was obtained. R–Na+ was rinsed with 2 mol/L FeCl3. The resin 8.26, 3.35 and 2.05, respectively. Zero charge points of R–Na+, R–Fe3+
was rinsed with deionized water to no outflow of Cl−. D151 resin and R–H+ decrease gradually, which proves that the acidity of the resin
preloaded with Fe3+ [denoted as R–Fe3+] was obtained. Fe3+ was surface increases (Lazarević et al., 2007). Total exchange capacities of
represented with Ni2+ and Cu2+ as the same operation by turns, the resins are determined by acid-base neutralization titration (Xiao
R–Ni2+ and R–Cu2+ were obtained accordingly. The method is simple and Wen, 2016). The main characteristic parameters of the resins are
to operate, low cost and beneficial to industrial production. listed in Table S1.

2.3. Static adsorption of glyphosate 3.2. Spectral analysis of the adsorption mechanism of glyphosate by R–Fe3+

About 0.200 g (W g) of resin were weighed out and placed into the 3.2.1. FTIR analyses
flasks containing 25 mL of glyphosate solution. Its initial concentrations As shown in Fig. S2, Fourier transform infrared spectra of R–H+,
(Ci mg/mL) of glyphosate ranged from 0.5 to 1.1 mg/mL pH value and R–Na+ and R–Fe3+ were measured by ATR method. 2930 cm−1 and
the content of salt in glyphosate solution were regulated. Adsorption 1700 cm−1 can be assigned to C–H stretching and C]O stretching of
equilibrium was reached in a shaker at 150 rpm. The equilibrium R–H+, respectively. 1450 cm−1 and 1160 cm−1 are related to –COO–

2
G. Xiao and Q. Meng Ecotoxicology and Environmental Safety 190 (2020) 110140

asymmetric stretching and C–O stretching of R–H+. After H+ of R–H+


is replaced by Na+, C]O and C–O of COONa are homogenized into two
coupling equal carbon oxygen bonds, which brings the result that
strong anti-symmetric stretching and symmetric stretching appear at
1550 cm−1 and 1400 cm−1. The characteristic absorption peak of Fe–O
in R–Fe3+ appears at 648 cm−1. The stretching vibration peaks of C–H
(2933 cm−1) and C]O (1704 cm−1) in R–Fe3+ show “red shift”
compared with the infrared spectra of R–H+. Compared with the in-
frared spectrum of R–Na+, the symmetrical stretching vibration peak of
COO− (1415 cm−1) appears “red shift”, while its anti-symmetric
stretching vibration peak (1548 cm−1) exhibits “blue shift”, indicating
that Fe3+ is partially coordinated with COO−. Due to the coordination
of Fe3+ with COO−, Fe3+ inside R–Fe3+ is fixed by the coordination
bond and are not to fall off from R–Fe3+ into aqueous solution.
As for the infrared spectra of R–Fe3+ adsorbed glyphosate [denoted
as R–Fe3+/glyphosate], symmetrical stretching vibration peak of
COO− shows “red shift” from 1415 cm−1 to 1420 cm−1. The anti-
symmetric vibration peak (1548 cm−1) of R–Fe3+ and in-plane bending
vibration peak of N–H (1558 cm−1) of glyphosate exhibit “red shift”, Fig. 1. The comparative adsorption of glyphosate by the five resins
and merge into the broad absorption peak (1590 cm−1) of R–Fe3+/ (m = 0.200 g, V = 25 mL, Ci = 0.5–1.1 mg/mL, T = 298 K, t = 24 h).
glyphosate. The vibration peak of P]O (1155 cm−1), P–OH
(908 cm−1) and the asymmetric stretching vibration of P–O
usually 6 (Ramezanpour et al., 2014). Compared with Ni2+ and Cu2+,
(1070 cm−1) of glyphosate all merge into the broad absorption peak
the vacant orbitals of Fe3+ are partially coordinated by lone pair
(1060 cm−1) of R–Fe3+/glyphosate. All the spectral changes indicate
electrons of COO− in R–Fe3+, and more remaining vacant orbitals can
that R–Fe3+ and glyphosate have been coordinated and integrated into
form coordination bonds with glyphosate. Therefore, adsorption capa-
R–Fe3+/glyphosate after the adsorption of glyphosate.
city of glyphosate by R–Fe3+ is much greater than that of R–Ni2+ and
R–Cu2+. R–Fe3+ exhibits the eximious adsorption performance of
3.2.2. XPS analyses
glyphosate.
XPS analyses before and after adsorption have been carried to elu-
cidate the adsorption mechanism of glyphosate by R–Fe3+ further. As
shown in Fig. S3. (a), N1s and P2p emerge in the spectrum of R–Fe3+/ 3.4. Excellent salt resistance of R–Fe3+
glyphosate, indicating glyphosate molecules are connected on the sur-
face of R–Fe3+ after the adsorption of glyphosate. XPS spectrum of N1s Industrial wastewater used to produce glyphosate contains a lot of
of glyphosate is presented in Fig. S3. (b). The high binding energy (BE) NaCl, and the maximum content of NaCl is 13–15% (Liu et al., 2013;
at 401.9 eV ascribes to NH2+, and the low BE at 399.9 eV assigns to Shen et al., 2013). Therefore, NaCl is selected to investigate the salt
NH (Sheals et al., 2007). N1s of R–Fe3+/glyphosate (Fig. S3. (c)) is resistance of R–Fe3+ for the adsorption of glyphosate, and the max-
divided into three peaks at 401.9, 400.6, and 399.9 eV, which are at- imum content of NaCl is controlled 16%. The experimental results are
tributed to NH2+, Fe–N, and NH, respectively (Fiorilli et al., 2017). shown in Fig. 2. Large adsorption capacities of glyphosate by 330 resin
Fe–N peak provides the direct evidence that Fe3+ inside R–Fe3+ has and D301 resin can be observed in the absence of NaCl. The adsorption
been coordinated with N of glyphosate. P2p BE of glyphosate (Fig. S3. capacities of glyphosate by 330 resin and D301 resin are very small at
(d)) occurs at 133.9 eV (P]O) and 132.9 eV (P–O), and that of R–Fe3+/ 0.3% NaCl and at 2% NaCl respectively. Adsorption of glyphosate by
glyphosate (Fig. S3. (e)) adds a new BE peak at 133.4 eV (P–O–Fe) 330 resin and D301 resin is based on the mechanism of anion exchange.
indicating that Fe3+ inside R–Fe3+ has been coordinated with O of P–O Cl− of NaCl will compete with glyphosate anion for the anion exchange
(Xie et al., 2017). O1s of R–Fe3+ (Fig. S3. (f)) has three chemical states, sites of 330 resin and D301 resin, resulting in significantly lessened
and their corresponding binding energies are 530.1 eV (Fe–O), 531.6 eV adsorption capacity of glyphosate by the two resins. According to Fig. 2,
(C–O) and 533.0 eV (C]O), respectively (Li et al., 2018). The char- NaCl exhibits no significant effect on the adsorption of glyphosate by
acteristic peak of O1s of R–Fe3+/glyphosate (Fig. S3. (g)) occurs at R–Fe3+. Adsorption of glyphosate by R–Fe3+ is not anion exchange
531.3 ev assigning to Fe–O–P and P]O after the adsorption of gly- mechanism, but that Fe3+ inside the resin can form coordination bonds
phosate (Xie et al., 2017). As described adsorption mechanism of gly- with glyphosate. Even if the content of NaCl in glyphosate solution is as
phosate by R–Fe3+ in Scheme 1, XPS spectra reveal the coordination of high as 16%, the salt shows no significant effect on the adsorption of
Fe3+ inside R–Fe3+ with O atoms of P–O and N atoms in glyphosate glyphosate by R–Fe3+. The paper provide the first evidence of Fe3+
molecule. inside D151 resin as an excellent salt resistant adsorbent for glyphosate
from water in the presence 16% NaCl.
3.3. Comparative adsorption of glyphosate
3.5. Effect of pH on the adsorption of glyphosate by R–Fe3+
3+ 2+ 2+
Comparative adsorption of glyphosate by R–Fe , R–Ni , R–Cu ,
R–Na+ and R–H+ is shown in Fig. 1. The adsorption capacity of gly- As shown in Fig. 3, pH exerts a significant effect on the adsorption of
phosate by R–Fe3+ is much larger than that of R–Ni2+, R–Cu2+, R–Na+ glyphosate by R–Fe3+. At pH 3.35, which is the zero charge point of
and R–H+. pKa1 of glyphosate is 0.78 (Fiorilli et al., 2017). Phosphate R–Fe3+, R–Fe3+ shows the net charge of zero and the maximum ad-
group of glyphosate ionizes H+, and H+ binds to amino group of gly- sorption capacity of glyphosate. At pH < 3.35, H+ increases with the
phosate. As there is no added acid or base, glyphosate exits mainly in decrease of pH values of glyphosate solution, and R–Fe3+ gradually
the form of HO–PO2–CH2–NH2+-CH2-COOH. The interaction force turns into R–H+. Fe3+ is desorbed from R–Fe3+, and the adsorption
between the functional groups of glyphosate and Na+ or H+ is small, so capacity of glyphosate decreases. At pH > 3.35, many negative
R–Na+ or R–H+ can hardly adsorb glyphosate. The central ion co- charges exist on the surface of R–Fe3+. The negative charge of R–Fe3+
ordination numbers of Ni2+ and Cu2+ are 4 (Takjoo et al., 2014; Jóvári increase with pH values of glyphosate solution. According to pKa and
et al., 2014), while the central ion coordination number of Fe3+ is the ionization equation of glyphosate, the ionization degree of

3
G. Xiao and Q. Meng Ecotoxicology and Environmental Safety 190 (2020) 110140

Fig. 2. The salt resistance of the resins (m = 0.200 g, V = 25 mL, Ci = 0.5–1.4 mg/mL, T = 298 K, t = 24 h).

Fig. 3. Effect of pH on the adsorption of glyphosate by R–Fe3+ (m = 0.200 g,


Fig. 4. The adsorption isotherms of glyphosate by R–Fe3+ (m = 0.200 g,
V = 25 mL, Ci = 1.1 mg/mL, T = 298 K, t = 24 h).
V = 25 mL, Ci = 0.5–1.1 mg/mL, t = 24 h, 16%NaCl).

glyphosate increases with pH values of glyphosate solution, and the


glyphosate by R–Fe3+ is accordance with Freundlich equation, and the
negative charge of glyphosate anion increase. The repulsive force of the
adsorption of glyphosate by R–Fe3+ is heterogeneous (Xiao et al.,
negative charges between R–Fe3+ and glyphosate anion increase, re-
2017). n > 2 indicates that the adsorption of glyphosate by R–Fe3+ is
sulting in the adsorption capacity of glyphosate by R–Fe3+ decreases.
favorable (Xiao and Wen, 2016). The higher values of KF and n are
observed at the higher temperature for the adsorption of glyphosate by
3.6. Adsorption thermodynamics of glyphosate by R–Fe3+ R–Fe3+, which is consistent with the trend that the adsorption capa-
cities of glyphosate increase with the increase of temperature.
Fig. 4 describes the adsorption isotherms of glyphosate by R–Fe3+. The adsorption thermodynamic parameters such as adsorption en-
The greater adsorption capacity of glyphosate by R–Fe3+ is observed thalpy ΔH (kJ/mol), adsorption entropy ΔS (J/(mol K)), and adsorption
with the higher temperature, indicating an endothermic process in- free energy ΔG (kJ/mol) can be calculated based on the following
volving a chemical reaction. Fe3+ inside D151 resin can form the co- equations (Zhang and Huang, 2017; Yamaguchi et al., 2016; Xiao et al.,
ordination bonds with glyphosate, and the increasing temperature is 2016):
conducive to the formation of the ligand between Fe3+ and glyphosate.
d ln Ce H
=
Freundlich equation q = KF Ce n
1
(2) dT RT 2 (4)

1 1 1
Langmuir equation = + ΔG = -RTlnKd (5)
q qm qm KL Ce (3)
ΔS=(ΔH-ΔG)/T (6)
Freundlich equation and Langmuir equation are used for the ad-
sorption isotherms of glyphosate by R–Fe3+. KF and n are the para- where qe and Ce are equilibrium adsorption capacity and equilibrium
meters of Freundlich equation to reflect the adsorption capacity (Peng concentration, respectively. Kd stands the equilibrium solid-liquid dis-
et al., 2019), while qm and KL are the parameters of Langmuir equation tribution coefficient, and which can be obtained from the intercept of
to reflect the adsorption capacity (Wang et al., 2019). The fitting results the ordinate by plotting ln(qe/Ce) versus qe (Yamaguchi et al., 2016). T
of the adsorption of glyphosate by R–Fe3+ are listed in Table S2. The (K) stands the temperature, and R (8.314 J/mol K) is the ideal gas
correlation coefficients (R2 > 0.99) indicate that the adsorption of constant. ΔH can be calculated by plotting ln Ce versus 1/T based on Eq.

4
G. Xiao and Q. Meng Ecotoxicology and Environmental Safety 190 (2020) 110140

the results are shown in Fig. 5. With the increase of the mole ratio of
glyphosate to Fe3+ inside R–Fe3+, the coordinated glyphosate in-
creases. As the mole ratio of glyphosate to Fe3+ inside R–Fe3+ increases
to 2.4: 1, even if glyphosate is added, the coordinated glyphosate re-
mains unchanged. The maximum coordination ratio of glyphosate to
Fe3+ inside R–Fe3+ is 1:1. The maximum coordination ratio is also
described in the adsorption mechanism in Scheme 1. Fe3+ inside
R–Fe3+ is 2.85 mmol/g, and the maximum adsorption capacity (qmax)
of glyphosate by R–Fe3+ is calculated to be 481.85 mg/g.
As compared qmax of glyphosate by R–Fe3+ with the adsorbents in
the literature (Zhou et al., 2017; Zavareh et al., 2018; Jiang et al., 2018;
Yamaguchi et al., 2016; Hu et al., 2011; Xiao and Wen, 2016; Herath
et al., 2016; Yang et al., 2017; Khenifi et al., 2010; Vasiljević et al.,
2019; Shen et al., 2006; Zhou et al., 2014; Milojević -Rakić et al., 2013;
Mayakaduwa et al., 2016) (Table 1), the adsorption capacity of gly-
phosate by R–Fe3+ is much higher than that of the other adsorbents,
except for UiO-67/GO (Yang et al., 2017) and alkalescent fiber FFA-1
(Zhou et al., 2014). However, it is exciting that the adsorption capacity
of glyphosate by R–Fe3+ in the presence 16% NaCl is much higher than
Fig. 5. The coordination ratio of glyphosate to Fe3+ inside R–Fe3+ (the mole
ratio of glyphosate to Fe3+ inside R–Fe3+ = 0.12:1, 0.24:1, 0.36:1, 0.48:1, that of the other reported adsorbents. R–Fe3+ is proven to be a pro-
0.72:1, 0.96:1, 1.2:1, 2.4:1, 3.6:1, 4.8:1, 6.0:1, 12:1, T = 298 K, t = 24 h). mising salt resistant adsorbent for the efficient recovery of glyphosate
in the high concentration of salt solution.

(4) (Zhang and Huang, 2017). ΔG and ΔS can be calculated according to


Eq. (5) and Eq. (6) (Xiao et al., 2016), respectively. ΔH, ΔG and ΔS are
3.8. Desorption agents of glyphosate
presented in Table S3. The values of ΔH are all positive, illustrating that
the adsorption of glyphosate by R–Fe3+ is endothermic further (Zhang
25 mL of H2O, 2 mol/L NaOH, 2 mol/L H2SO4 and 2 mol/L
and Huang, 2017). In particular, the values of ΔH increase with the
Fe2(SO4)3 are used as desorption agents respectively, and the results are
increase of adsorption capacity. The values of ΔG are negative, con-
shown in Table S4. The desorption rate of glyphosate is only 5.35% as
firming that the adsorption of glyphosate by R–Fe3+ is spontaneous
water is used as the desorption agent. The coordination between gly-
(Lou et al., 2012). Moreover, the absolute values of ΔG increase with
phosate and Fe3+ is stronger than that of water, so the desorption rate
the increasing temperature, implying it is the more spontaneous ad-
of glyphosate by water is small. 2 mol/L NaOH can be used to re-
sorption of glyphosate for the higher temperature. The result is in
generate R–Fe3+. Fe3+ inside R–Fe3+ is easy to bind with OH−, so that
agreement with their capacities of glyphosate by R–Fe3+. The positive
Fe3+ no longer coordinates with glyphosate. 2 mol/L Fe2(SO4)3 is used
values of ΔS can be explained by the release of H2O molecules around
as the desorption agent, and the regeneration ratio of glyphosate is up
Fe3+ after the adsorption of glyphosate. This is the reason that gly-
to 98.92%. More Fe3+ in desorption agent Fe2(SO4)3 will compete with
phosate, as a stronger ligand than H2O molecule, coordinates with
Fe3+ inside R–Fe3+ for glyphosate. Glyphosate adsorbed by R–Fe3+
Fe3+. The positive values of ΔS show that the adsorption of glyphosate
will coordinate with Fe3+ of Fe2(SO4)3, and glyphosate can be desorbed
by R–Fe3+ is the ligand exchange of glyphosate and water.
from R–Fe3+. The desorption rate of glyphosate is up to 99.97% using
2 mol/L H2SO4 as the desorption agent. Under the strong acidic con-
3.7. Coordination ratio of glyphosate to Fe3+ inside R–Fe3+ dition of 2 mol/L H2SO4, R–Fe3+ is converted to R–H+, releasing of
Fe3+ and its ligand of glyphosate. 2 mol/L NaOH, 2 mol/L H2SO4 and
The mole ratio of glyphosate to Fe3+ inside R–Fe3+ (0.12:1, 0.24:1, 2 mol/L Fe2(SO4)3 can all be used for the desorption agent to achieve
0.36:1, 0.48:1, 0.72:1, 0.96:1, 1.2:1, 2.4:1, 3.6:1, 4.8:1, 6.0:1, 12:1) is over 97% regeneration of R–Fe3+ and the recovery of glyphosate.
adopted. The adsorption reaches equilibrium at 298 K, and the co-
ordination ratio of glyphosate to Fe3+ inside R–Fe3+ is calculated, and

Table 1
Comparison of qmax of glyphosate adsorbed by R–Fe3+ with the adsorbents in the literature.
NO Adsorbents qmax/(mg/g) Salt References

1 D201Cu 113.7 18% NaCl Zhou et al. (2017)


2 Cu-zeolite 4A 112.7 No Zavareh et al. (2018)
3 Biochar supported nano-zero-valent iron 80 No Jiang et al. (2018)
4 MnFe2O4-G 39 No Yamaguchi et al., 2016
5 Alum sludge in dewatered form 85.9 No Hu et al. (2011)
6 Alum sludge in liquid form 113.6 No Hu et al. (2011)
7 GQ-08 resin 66.44 16% NaCl Xiao and Wen (2016)
8 Steam activated RHBC 123.03 No Herath et al. (2016)
9 UiO-67/GO 482.69 No Yang et al. (2017)
10 Ni2AlNO3 layered double hydroxide 172.4 No Khenifi et al. (2010)
11 Potassium tungstophosphate on BEA zeolite 92.2 No Vasiljević et al. (2019)
12 Active carbon 58.4 No Shen et al. (2006)
13 Alkalescent fiber FFA-1 565.19 No Zhou et al. (2014)
14 PANI/ZSM-5 61.9 No Milojević -Rakić et al., 2013
15 PANI 59.9 No Milojević -Rakić et al., 2013
16 Woody biochar 44 No Mayakaduwa et al. (2016)
17 (D151-COO)3Fe 481.85 16% NaCl This work

5
G. Xiao and Q. Meng Ecotoxicology and Environmental Safety 190 (2020) 110140

4. Conclusions Lou, S., et al., 2012. Synthesis of functional adsorption resin and its adsorption properties
in purification of flavonoids from Hippophae rhamnoides L. leaves. Ind. Eng. Chem.
Res. 51, 2682–2696.
D151 resin preloaded with Fe3+ is simple to operate, low cost and Mata, K.D., et al., 2014. Synthesis and characterization of cross-linked molecularly im-
beneficial to industrial production. XPS spectra reveals the coordination printed polyacrylamide for the extraction/preconcentration of glyphosate and ami-
of Fe3+ inside R–Fe3+ with O atoms of P–O and N atoms in glyphosate nomethylphosphonic acid from water samples. React. Funct. Polym. 83, 76–83.
Mayakaduwa, S.S., et al., 2016. Equilibrium and kinetic mechanisms of woody biochar on
molecule. Fe3+ inside D151 resin exhibits the eximious adsorption aqueous glyphosate removal. Chemosphere 144, 2516–2521.
performance of glyphosate and excellent salt resistance in the presence Milojević -Rakić, M., et al., 2013. Polyaniline and its composites with zeolite ZSM-5 for
16% NaCl. The maximum coordination ratio of glyphosate to Fe3+ in- efficient removal of glyphosate from aqueous solution. Microporous Mesoporous
Mater. 180, 141–155.
side D151 resin is 1:1. The coordination adsorption of glyphosate by Peng, R.X., et al., 2019. Catalyst-free synthesis of triazine-based porous organic polymers
R–Fe3+ not only solves the serious environmental problems of gly- for Hg2+ adsorptive removal from aqueous solution. Chem. Eng. J. 371, 260–266.
phosate, but also has important economic benefits through the efficient Ramezanpour, B., et al., 2014. Seven and eight-coordinate Fe(III) complexes containing
pre-organized ligand 1,10-phenanthroline-2,9-dicarboxylic acid: solvent effects, su-
recovery of glyphosate.
pramolecular interactions and DFT calculations. Inorg. Chim. Acta 484, 264–275.
Rubí-Juárez, H., et al., 2016. Removal of herbicide glyphosate by conductive-diamond
CRediT authorship contribution statement electrochemical oxidation. Appl. Catal. B Environ. 188, 305–312.
Sheals, J., et al., 2007. Adsorption of glyphosate on goethite: molecular characterization
of surface complexes. Environ. Sci. Technol. 36, 3090–3095.
Guqing Xiao: Methodology, Data curation, Writing - original draft. Shen, J.N., et al., 2013. The use of BMED for glyphosate recovery from glyphosate neu-
Qiudong Meng: Writing - review & editing. tralization liquor in view of zero discharge. J. Hazard Mater. 260, 660–667.
Shen, L.J., et al., 2006. Static absorption of glyphosate by active carbon. Chem. Eng. 10
(2006), 46–49 (in Chinese).
Acknowledgments Skeff, W., et al., 2018. Adsorption behaviors of glyphosate, glufosinate, aminomethyl-
phosphonic acid, and 2-aminoethylphosphonic acid on three typical Baltic Sea se-
We gratefully acknowledge generous support provided by the diments. Mar. Chem. 198, 1–9.
Song, J.F., et al., 2013. Composite hollow fiber nanofiltration membranes for recovery of
National Natural Science Foundation of China (Grant No. 51678225). glyphosate from saline wastewater. Water Res. 47, 2065–2074.
Takjoo, R., et al., 2014. Synthesis, characterization, X-ray structure and DFT calculation
Appendix A. Supplementary data of two Mo(VI) and Ni(II) Schiff-base complexes. C. R. Chimie 17, 1144–1153.
Vasiljević, B.N., et al., 2019. In situ synthesis of potassium tungstophosphate supported
on BEA zeolite and perspective application for pesticide removal. J. Environ. Sci. 81,
Supplementary data to this article can be found online at https:// 136–147.
doi.org/10.1016/j.ecoenv.2019.110140. Wang, S.Z., et al., 2016. Bio)degradation of glyphosate in water-sediment microcosms -A
stable isotope co-labeling approach. Water Res. 99, 91–100.
Wang, X.M., et al., 2019. Amino-modified hyper-cross-linked polymers with hierarchical
References porosity for adsorption of salicylic acid from aqueous solution. J. Chem. Thermodyn.
131, 1–8.
Arroyave, J.M., et al., 2016. Effect of humic acid on the adsorption/desorption behavior Xiao, G.Q., et al., 2017. Adsorption performance of salicylic acid on a novel resin with
of glyphosate on goethite. Isotherms and kinetics. Chemosphere 145, 34–41. distinctive double pore structure. J. Hazard Mater. 329, 77–83.
Divisekara, T., et al., 2018. Impact of a commercial glyphosate formulation on adsorption Xiao, G.Q., et al., 2016. Effects of the hydrophobicity of adsorbate on the adsorption of
of Cd(II) and Pb(II) ions on paddy soil. Chemosphere 198, 334–341. salicylic acid and 5-sulfosalicylic acid onto the hydrophobic-hydrophilic macro-
Fiorilli, S., et al., 2017. Iron oxide inside SBA-15 modified with amino groups as reusable porous polydivinylbenzene/polymethylacrylethylenediamine IPN. Fluid Phase
adsorbent for highly efficient removal of glyphosate from water. Appl. Surf. Sci. 411, Equilib. 421, 33–38.
457–465. Xiao, G.Q., Wen, R.M., 2016. Comparative adsorption of glyphosate from aqueous solu-
Herath, I., et al., 2016. Mechanistic modeling of glyphosate interaction with rice husk tion by 2-aminopyridine modified polystyrene resin, D301 resin and 330 resin: infl-
derived engineered biochar. Microporous Mesoporous Mater. 225, 280–288. uencing factors, salinity resistance and mechanism. Fluid Phase Equilib. 411, 1–6.
Hu, Y.S., et al., 2011. Removal of glyphosate from aqueous environment by adsorption Xie, Q.Y., et al., 2017. Effective adsorption and removal of phosphate from aqueous so-
using water industrial residual. Desalination 271, 150–156. lutions and eutrophic water by Fe-based MOFs of MIL-101. Sci. reports 7, 3316–3331.
Huo, R., et al., 2017. Visible-light photocatalytic degradation of glyphosate over BiVO4 Yamaguchi, N.U., et al., 2016. Magnetic MnFe2O4-graphene hybrid composite for effi-
prepared by different co-precipitation methods. Mater. Res. Bull. 88, 56–61. cient removal of glyphosate from water. Chem. Eng. J. 295, 391–402.
Jiang, X.Y., et al., 2018. Mechanism of glyphosate removal by biochar supported nano- Yang, Q.F., et al., 2017. Interface engineering of metal organic framework on graphene
zero-valent iron in aqueous solutions. Colloids Surf., A 547, 64–72. oxide with enhanced adsorption capacity for organophosphorus pesticide. Chem.
Jóvári, P., et al., 2014. Fourfold coordinated Te atoms in amorphous GeCu2Te3 phase Eng. J. 313, 19–26.
change material. Scr. Mater. 68, 122–125. Yang, Y., et al., 2018. Comparative study of glyphosate removal on goethite and mag-
Khenifi, A., et al., 2010. Adsorption of Glyphosate and Glufosinate by Ni2AlNO3 layered netite: adsorption and photo-degradation. Chem. Eng. J. 352, 581–589.
double hydroxide. Appl. Clay Sci. 47, 362–371. Zavareh, S., et al., 2018. Modification of zeolite 4A for use as an adsorbent for glyphosate
Lan, H.C., et al., 2016. An activated carbon fiber cathode for the degradation of gly- and as an antibacterial agent for water. Ecotoxicol. Environ. Saf. 155, 1–8.
phosate in aqueous solutions by the Electro-Fenton mode: optimal operational con- Zhang, Q., et al., 2018. The combined toxicity effect of nanoplastics and glyphosate on
ditions and the deposition of iron on cathode on electrode reusability. Water Res. Microcystis aeruginosa growth. Environ. Pollut. 243, 1106–1112.
105, 575–582. Zhang, T., Huang, J.H., 2017. N-vinylimidazole modified hyper-cross-linked resins and
Lazarević, S., et al., 2007. Adsorption of Pb2+, Cd2+ and Sr2+ ions onto natural and acid- their adsorption toward Rhodamine B: effect of the cross-linking degree. J. Taiwan
activated sepiolites. Appl. Clay Sci. 37, 47–57. Inst Chem Eng. 80, 293–300.
Li, Y.J., et al., 2018. Adsorption performance and mechanism of magnetic reduced gra- Zhou, C.R., et al., 2014. Study on behavior of alkalescent fiber FFA-1 adsorbing gly-
phene oxide in glyphosate contaminated water. Environ. Sci. Pollut. Res. 25, phosate from production wastewater of glyphosate. Fluid Phase Equilib. 362, 69–73.
21036–21048. Zhou, C.Y., et al., 2017. Removal of glyphosate from aqueous solution using nanosized
Liu, Z.Y., et al., 2013. Pretreatment of membrane separation of glyphosate mother liquor copper hydroxide modified resin: equilibrium isotherms and kinetics. J. Chem. Eng.
using a precipitation method. Desalination 313, 140–144. Data 62, 3585–3592.

You might also like