You are on page 1of 6

Chemosphere 169 (2017) 437e442

Contents lists available at ScienceDirect

Chemosphere
journal homepage: www.elsevier.com/locate/chemosphere

Amine-functionalized, multi-arm star polymers: A novel platform for


removing glyphosate from aqueous media
Lianna Samuel*, Ran Wang, Geraud Dubois, Robert Allen, Rudy Wojtecki, Young-Hye La**
IBM Almaden Research Center, 650 Harry Road, San Jose, CA 95120, USA

h i g h l i g h t s g r a p h i c a l a b s t r a c t

 Bioaccumulation of glyphosate
ingested in even trace amounts can
negatively impact health.
 Rapid, efficient glyphosate adsorp-
tion by mesoporous-like star-poly-
mers with high amine functional
group aerial density.
 Ionic interactions which promote
adsorption are affected by pH.
 Glyphosate removal efficiency is
greater than 85% for neutral and
slightly basic conditions.
 Maximum glyphosate adsorption ca-
pacity is 229.017 mg glyphosate/g
star-polymer.

a r t i c l e i n f o a b s t r a c t :

Article history: We describe a novel method for efficiently removing glyphosate from aqueous media via adsorption onto
Received 12 August 2016 highly functionalized star-shaped polymeric particles. These particles have a polystyrene core with more
Received in revised form than 35 attached methacrylate polymer arms, each containing a plurality of pendant amines (poly(-
8 November 2016
dimethylamino ethyl methacrylate): PDMAEMA) that are partially protonated in water. Kinetic studies
Accepted 10 November 2016
demonstrate that these star-polymers successfully remove up to 93% of glyphosate present in aqueous
solution (feed concentration: 5 ppm), within 10 min contact time, outperforming activated carbon,
Handling Editor: Shane Snyder which removed 33% after 20 min. On these star-polymers, glyphosate adsorption closely follows the
Langmuir model indicating monolayer coverage at most. Ionic interaction between the protonated
Keywords: amines and glyphosate's dissociated carboxylic and phosphoric acid groups lead to effective glyphosate
Glyphosate removal capture even at feed concentrations below 1 ppm. Surface charge of these star polymers and dissociation
Star polymer of glyphosate are both influenced by pH, thus glyphosate removal efficiency increases from 63% to 93%
Adsorption when pH increases from 4.2 to 7.7. NMR studies conducted with butylamine as a proxy for these poly-
Functionalized nanomaterials
meric particles confirm that the amine group binds with both glyphosate's carboxylic and phosphoric
Water purification
acid groups when its concentrations are in a 2:1 or higher molar ratio with glyphosate.
© 2016 Elsevier Ltd. All rights reserved.

1. Introduction

Population growth has driven the increased use of pesticides


* Corresponding author.
** Corresponding author.
and herbicides in agriculture globally, to ensure adequate food
E-mail addresses: samuell@us.ibm.com (L. Samuel), yna@us.ibm.com (Y.-H. La). supply. Glyphosate [Ne(phosphonomethyl) glycine] is the active

http://dx.doi.org/10.1016/j.chemosphere.2016.11.049
0045-6535/© 2016 Elsevier Ltd. All rights reserved.
438 L. Samuel et al. / Chemosphere 169 (2017) 437e442

ingredient of a broad spectrum, nonselective, post-emergence structure for successfully removing low concentrations of glypho-
herbicide. This herbicide is one of the most widely used in the sate from aqueous media. To better understand the relationship of
world, controlling the growth of various weeds and grasses (Guo material structure and glyphosate removal efficacy, comparative
et al., 2005), which can stifle the development of desirable crops, adsorption studies were also performed with polyamidoamine
reducing both the quantity and quality of harvests. Glyphosate (PAMAM) dendrimer with numerous amine-functional groups in
usage has been largely enhanced by the use of genetically modified its terminal positions along with activated carbon (control). Addi-
crops that are glyphosate-resistant and promote its direct appli- tionally, the adsorption mechanism was also investigated together
cation to these crops as well as targeted weeds and grasses. This with the influence of pH, and feed concentration on the overall
widespread use of glyphosate has led to its presence and one of its glyphosate removal efficiency of the star polymer.
major metabolite aminomethylphosphonic acid (AMPA) in the
environment due to run-off (both during and after its application 2. Experimental (materials and methods)
due to leaching).
Additionally, any applied glyphosate that remains on the har- Batch glyphosate adsorption tests were carried out with the
vested crops can be transferred to consumer products by contam- amine-functionalized star polymer (SP-PDMAEMA), PAMAM den-
ination of processing machinery used during their preparation. drimers (PAMAM6) with ethylenediamine core (generation 6.0,
Thus, traces of glyphosate and AMPA can be found not only in Sigma-Aldrich, WI, USA), and activated carbon (100 mesh, pow-
natural waterways but also on fruits and produce available for der) (Sigma-Aldrich, WI, USA) to investigate the importance of
human consumption (Solomon and Thompson, 2003; Relyea, 2005; adsorbent structure, chemistry and amine-functional density on
Buffin and Jewell, 2014). The presence of glyphosate (in various glyphosate adsorption. SP-PDMAEMA consists of a hydrophobic
herbicide formulations) and its metabolite in both our food and core made up of cross-linked polystyrene and hydrophilic meth-
drinking water supplies has raised many concerns (Freuze et al., ylacrylate polymer arms with a plurality of tertiary amines (ex. Poly
2007; Ho and Cherry, 2009; Qian et al., 2009; Paganelli et al., (N, N-dimethylaminoethyl methacrylate): PDMAEMA) (Fig. 1(a)).
2010). As an herbicide, glyphosate disrupts the shikimate The polymer was synthesized via the procedure described in an
pathway from amino acid biosynthesis in plants (Scho € nbrunn et al., earlier literature (Lee et al., 2011).
2001). However, it has also been found to negatively affect enzyme A 10 ppm stock solution of glyphosate was made by dissolving
activity (Freuze et al., 2007), cause cytotoxicity and DNA damage in 0.01 g of analytical grade glyphosate powder (Sigma-Aldrich, Pes-
human cells (Gasnier et al., 2009; Paganelli et al., 2010) and damage tanal®) in 1 L deionized water (made in house, via reverse osmosis
to the mammalian endocrine system with prolonged exposure with resistivity of 18.4 million ohm-cm and TDS: ~7 ppm). 250 ml of
(Richard et al., 2005). the glyphosate stock solution was added to 250 ml of deionized
Globally, the allowable concentration of glyphosate in drinking water to create a desired initial concentration of 5 ppm. 10 ml ali-
water ranges from 0.1 mgL1 (European Union Drinking Water quots of the 5 ppm glyphosate solution were then placed into a
Regulations, for all pesticides) (European Commission, 1991) to series of 25 ml vials containing 0.005 g of activated carbon, 0.125 ml
0.7 mgL1 (Environmental Protection Agency drinking water PAMAM6 dendrimers (5% weight dendrimers in methanol) and
regulatory guidelines) (EPA webpage, accessed 2015). More 5 ml SP-PDMAEMA solution (0.1% weight SP-PDMAEMA by vol-
recently, the World Health Organization announced that glypho- ume), respectively, to ensure that the same mass of adsorbent was
sate is a likely carcinogen to humans, as well (Fritschi et al., 2014). used for each trial. Each trial was repeated four times and the
Additionally, prolonged exposure to and ingestion of glyphosate glyphosate concentrations measured were averaged. For each trial,
even in trace amounts (10 mg/L) has been shown to have devas- blank deionized water specimens were collected. The adsorption of
tating effects on kidney health when combined environmental glyphosate onto SP-PDMAEMA, PAMAM6 and activated carbon was
factors such as water hardness and nephrotoxic metals then determined as a function of time and adsorption kinetics
(Jayasumana et al., 2014). Therefore, there is a need to minimize the determined.
amount of glyphosate present in drinking water and other food Duplicate studies to determine the effect of pH and initial
products. However, glyphosate has small molecular weight, high glyphosate concentration were also conducted using SP-
polarity and solubility in water although it is insoluble in organic PDMAEMA, to better understand its efficiency. PAMAM6 was not
solvents (Stalikas and Konidari, 2001), making its removal from used in the pH study due to its poor performance in capturing
aqueous solution difficult. Currently, numerous techniques (chlo- glyphosate. To determine the effect of pH on glyphosate adsorption,
rination, ozonation, photolysis and heterogeneous photocatalysis) the initial pH of 5 ppm glyphosate solution was maintained (pH: 7.7,
are used to reduce glyphosate concentrations in drinking water, control) and adjusted to 4.2, 8.1 and 11.6 respectively, by adding
either individually or in different combinations (Speth, 1993; 0.1 mol/L hydrochloric acid or 0.1 mol/L sodium hydroxide to the
Assalin et al., 2009). Obtaining an efficient, affordable method for glyphosate solution after the addition of SP-PDMAEMA. To deter-
glyphosate removal is still needed (Hu et al., 2011). Adsorption can mine the effect of varying initial glyphosate concentrations, the
be an affordable removal method but most activated carbon, stock solution was diluted with deionized water to obtain solutions
although used in generic water treatment because of their high with glyphosate concentrations ranging from 10 ppm to500 ppb
micro-porosity and degree of surface reactivity are less effective on (10, 7.5, 5, 2.5 and 1 ppm and 500 ppb). Two adsorption-isotherm
acidic, hydrophilic and highly soluble compounds like glyphosate models were selected to aid in describing the adsorption equilib-
(Hamerlinck et al., 1994; Mohan and Pittman, 2006; Bozkaya- rium for the adsorbents (Langmuir and Freundlich isotherms).
Schrotter et al., 2008). After the glyphosate adsorption, the complexes of glyphosate
Stareshaped polymers provide a potential solution, they are and SP-PDMAEMA and glyphosate and PAMAM6 were filtered us-
structurally designed to exhibit high adsorption capacity, with a ing PS20 filters (polysulfone membrane with 20 kDa molecular
large number of arms that can be functionalized and fine-tuned to weight cut-off, Sepro Membranes Inc.) under a pressure of 25 psi
capture specific molecules and form highly microporous structures while the complexes of glyphosate and activated carbon were
(Mishra and Kobayashi, 1999). In this work, batch adsorption filtered using Acrodisc® 25 mm syringe filters with 1 mm PTFE
equilibrium study has been conducted to examine the potential use membranes (Pall Laboratories) attached to 12 ml plastic syringes.
of in-house prepared, tailored star-polymer particles with high These filters were used as tests on the filtrates obtained from pure
aerial density of amine functional groups and mesoporous-like deionized water passing through these filters did not show any
L. Samuel et al. / Chemosphere 169 (2017) 437e442 439

Fig. 1. Comparative look at the amine functionalized star polymer and PAMAM dendrimer. (a) Schematic drawing and chemical structure of the amine-functionalized star polymer:
grey region: crosslinked polystyrene (PS) core with PS linear arms (Mw of PS ¼ 4 kDa), blue region: polydimethylaminoethyl methacrylate arms (PDMAEMA, Mw ¼ 7 kDa), Average
arm numbers ¼ 37, green circles: amine group (approximately 45 amine moieties on each arm giving a total of 1665 amine groups per starpolymer) (b) Schematic drawing of the
PAMAM, generation 6 dendrimer, green circles amine functional groups on the terminal ends (about 256 amine groups).

leaching or capture of phosphorous. Glyphosate concentration in


the filtrate was determined by measuring total phosphorous con-
tent with inductively coupled plasma spectrometer (Thermo Sci-
entific ICAP 6300 Duo View Spectrometer), which has a detection
limit of 2 ppb for phosphorous in deionized water. The molecular
formula of glyphosate is C3H8NO5P and there is a theoretical linear
relationship between total phosphorous and glyphosate. The con-
centration of glyphosate in filtrate was finally obtained from the
calibration curve, which is generated by measuring phosphorous
contents in 15 different glyphosate solutions with known concen-
trations (10 ppme500 ppb), in the same deionized water matrix.
Once the glyphosate concentration was determined for each sam-
ple for a given initial concentration tested and the average con-
centration was then determined. Variation in the sample
glyphosate concentrations for each initial concentration was below
±10%. Additionally, ICP analysis of blank deionized water samples
revealed detection of no phosphorous.
Fig. 2. Glyphosate adsorption kinetics relative to 5 ppm feed concnetration onto
activated carbon (blue), star polymer (red) and PAMAM dendrimers (green).
3. Results and discussion

3.1. Effect of adsorbent size and density of amine functional (z ¼ 20.0 mV). Also, under neutral conditions both the first
moieties phosphoric acid (pKa e 0.8) and the carboxylic acid (pKa e 2.3)
protons of glyphosate are fully dissociated giving the glyphosate
The performance of various adsorbents in capturing glyphosate molecules a negative surface charge. The negative surface charge
is shown in Fig. 2. Under the given testing conditions, the efficiency on the activated carbon causes relatively low adsorption via
of SP-PDMAEMA for glyphosate removal is approximately three repulsion of negative groups on the glyphosate, while the positive
times higher than that of activated carbon and eighteen times surface charge on the star polymer encourages ionic interaction
higher than that of PAMAM6. For SP-PDMAEMA and PAMAM6, between the positively charged amine groups and the negatively
adsorption equilibrium was achieved after 10 min of contact with charged carboxylic and phosphoric groups on the glyphosate.
glyphosate, while 20 min of contact was need to achieve adsorption Both PAMAM6 and the SP-PDMAEMA have amine functional
equilibrium with the activated carbon. This indicates that SP- groups. However, PAMAM6 dendrimers are approximately 5 times
PDMAEMA has both a faster removal rate and higher removal ca- smaller in diameter (6.0 nm) and only possess 256 amine functional
pacity than the other adsorbents tested. groups on their terminal surface with relatively lower zeta poten-
Under neutral conditions, while both SP-PDMAEMA and acti- tial (z ¼ þ 14.0 mV), whereas each SP-PDMAEMA star polymer has
vated carbon particles possess large diameters compared to more than 1500 amine functional groups (about 45 amines per
PAMAM6 dendrimer, (28.05 ± 0.38 nm and 149 mm, respectively) each hydrophilic arms) with a higher positive potential
and hence larger surface areas, they have opposite surface charges. (z ¼ þ24.6 mV). The amine groups attached to more than 35-
For SP-PDMAEMA, under neutral conditions a large number of the hydrophilic arms of star polymers could also generate inner
amine groups on its arms are positively charged (zeta potential, porous environment with higher local density of amines, which
z ¼ þ24.6 mV), while activated carbon has a negative zeta-potential might be advantageous for an efficient glyphosate binding. From
440 L. Samuel et al. / Chemosphere 169 (2017) 437e442

these results, not only the presence of amine groups but also the 3.3. Effect of pH
number, density and arrangement of these functional groups on
polymeric particles seem to be critical in determining its glypho- While the adsorption efficiency of SP-PDMAEMA varies with
sate adsorption efficiency. solution pH, its glyphosate removal efficiency does not fall below
50% except under very basic conditions (pH  11), Fig. 4. With
glyphosate's remaining pKa values ranging from 6.0 to 11
3.2. NMR study
(pKa3 ¼ 6.0 by 2nd phosphoric, and pKa4 ¼ 11 by secondary
amine), it follows that as pH decreases (pH < 7e8) the net negative
H-NMR studies conducted using chloroform (CDCl3) as a solvent
charge of the glyphosate also decreases, reducing the strength of
and butylamine as a proxy for the amine functional groups on both
ionic interactions between the glyphosate molecules and SP-
these polymeric particles, illustrated the mechanism by which the
PDMAEMA, lowering the adsorption efficiency. At higher pH
adsorption process occurs as well as the importance of amine-
(>10), glyphosate has a trivalent negative charge, however the
density on the adsorption process (Fig. 3). Under these condi-
protonation tendency of the amine groups on the star polymer
tions, when the amine groups and the glyphosate concentrations
significantly decreases, losing their positive charge, as the pKa of
are present in 1:1 M ratio, there is competition between carboxylic
protonated amines are around 10 (Homola and James, 1977).
and phosphoric acid moieties to protonate the amine group
(Fig. 3(b)). However, when the amine groups are present in a larger
concentration than glyphosate (in this case a 2:1 M ratio, Fig. 3(c)),
both the carboxylic and phosphoric acid moieties are deprotonated,
implying interaction between each glyphosate molecule and two
amine groups.
Although acid dissociation behavior and amine's protonation
tendency in the NMR solvent (chloroform) are somewhat different
from those in water media as the acidic groups in glyphosate would
be immediately deprotonated in water even without any base
(pKa1 ¼ 0.8 by 1st phosphoric, and pKa2 ¼ 2.3 by carboxylate), the
NMR data still supports the origin of ionic interactions between
glyphosate and an amine-compound depending on their relative
concentrations. The numerous amine groups on the structure of SP-
PDMAEMA provide a local environment within which the amine
molar concentration is more than 2 times larger than that of
glyphosate. Therefore, the ionic interactions between SP-
PDMAEMA and glyphosate follow that seen in Fig. 3(b) and lead
to more complete capture of glyphosate particles (one glyphosate
molecule binds with two amine groups). Fig. 4. Glyphosate adsorption as a function of pH, relative to 5 ppm feed concentration
onto SP-PDMAEMA after 10 min contact time.

Fig. 3. H-NMR spectra of glyphosate mixed with butylamine (serving as a proxy for the amine functional arms of the star polymer) with two different molar ratios under CDCl3. (a)
glyphosate only, (b) 1:1 M ratio of glyphosate and butylamine and (c) 1:2 M ratio of glyphosate and butylamine.
L. Samuel et al. / Chemosphere 169 (2017) 437e442 441

Table 1
Kinetic adsorption study of glyphosate remval by star polymer (SP-PDMAEMA) and activaed carbon. Feed concnetration (Co) 5 ppm, qe test e the amount of glyphosate
adsorbed at equilibrium during experiment, k1 e the first order adsorption constant, k2 e second order adsorption constant, qe1 - the amount of glyphosate adsorbed at
equilibrium following 1st order adsorption kinetics, qe2 - the amount of glyphosate adsorbed at equilibrium following 2nd order adsorption kinetics and R (Solomon and
Thompson, 2003) e correlation coefficient.

Adsorbent C0 (ppm) qe test (mg/g) First order kinetics Second order kinetics
1 2
k1 (min ) qe1 (mg/g) R k2 (g/mg/min) qe2 (mg/g) R2

SP-PDMAEMA 5 9.6 0.130 1.50 0.9437 0.086 9.63 0.9994


Activated Carbon 5 3.2 0.015 1.04 0.9263 0.042 3.18 0.9932

Therefore, there is less possible electrostatic interaction between 35-linear polymer chains, each containing a plurality of pendent
glyphosate and SP-PDMAEMA and so adsorption efficiency declines amines, outperformed both the tested activated carbon and
even further. It should be expected that dramatic reduction in SP- PAMAM6, which has low amine group density on its surface. The
PDMAEMA adsorption efficiency should be noticeable around pH high adsorption efficiency of the SP-PDMAEMA within the rela-
9, therefore to maintain its high performance it should only be used tively short contact time of 10 min makes the material a potential
under slightly acidic to neutral conditions. adsorbent for use in a scalable process requiring fast and effective
removal of glyphosate and possibly other organophosphorous
compounds. Moreover, since the functional group of the star
3.4. Sorption isotherms
polymers can be tailored depending on a target compound, this
could be a basic material platform to deal with a challenging im-
Due to the low adsorption efficiency of PAMAM6, adsorption
purity in water by coupling with activated carbon and/or other
kinetics and Langmuir and Freundlich isotherms were only fitted to
conventional water treatment methods. Hence, we demonstrated
data from SP-PDMAEMA and activated carbon. Table 1 shows that
glyphosate-removal efficacy using a simplified aqueous media
the adsorption of glyphosate onto SP-PDMAEMA and activated
formulated with DI water without using real drinking and/or sur-
carbon is well described by pseudo-second order adsorption ki-
face water. Future work to improve our understanding of SP-
netics (special boundary solution to Langmuir isotherm), similar to
PDMAEMA's glyphosate adsorption efficiency will include compu-
the well-documented results of adsorption studies conducted to
€ tational studies to clarify the SP-PDMAEMA's geometric structure
capture phosphorous (Ozacar, 2003). Further examination of
and its complexation behavior with glyphosate in water.
glyphosate adsorption using the Langmuir and Freundlich
adsorption isotherm models, illustrate that for both adsorbents,
adsorption of glyphosate more closely follows the Langmuir Acknowledgements
isotherm with maximum glyphosate adsorption capacity of
229.017 mg/g for SP-PDMAEMA, approximately four times the The authors thank Dr. Reinhold Dauskardt and Stanford Uni-
maximum adsorption capacity of activated carbon (58.4 mg/g) versity's Environmental Measurement Facility for their assistance
tested in this study. Comparison of SP-PDMAEMA's performance to in conducting ICP analysis on the samples for this work.
that of previously examined, functionalized activated carbons such
as KOH-impregnated activated carbon (AC-KOH, Salman et al., References
2012) and the one coated with manganese dioxide (AC-MnO2, Cui
Assalin, M.R., De Moraes, S.G., Queiroz, S.C., Ferracini, V.L., Duran, N., 2009. Studies
et al., 2012), SP-PDMAEMA has approximately two times and on degradation of glyphosate by several oxidative chemical processes: ozona-
almost the maximum adsorption capacity of the functionalized tion, photolysis and heterogeneous photocatalysis. J. Environ. Sci. Health, Part B
activated carbons (104.2 mg/g for AC-KOH and 283 mg/g for AC- 45 (1), 89e94. http://dx.doi.org/10.1080/03601230903404598.
Bozkaya-Schrotter, B., Daines, C., Lescourret, A.S., Bignon, A., Breant, P.,
MnO2, respectively). Further comparison of glyphosate adsorption
Schrotter, J.C., 2008. Treatment of trace organics in membrane concentrates I:
by these activated carbons and SP-PDMAEMA, show that the pesticide elimination. Water Sci. Technol. 8 (2), 223e230. http://dx.doi.org/
adsorption efficiency of SP-PDMAEMA follows that seen for AC- 10.2166/ws.2008.056.
Buffin, D., Jewell, T., 2014. Health and Environmental Impacts of Glyphosate: The
KOH, with pH having the greatest impact under very basic condi-
Implications of Increased Use of Glyphosate in Association with Genetically
tions. However, the adsorption efficiency of AC-MnO2 did exhibit a Modified Crops. Friends of the Earth. http://www.foe.co.uk/resource/reports/
dependence on pH with optimal adsorption occurring at a pH of 3, impacts_glyphosate.pdf (accessed 07.16.14).
where as for SP-PDMAEMA there is a slight decrease in adsorption Cui, H., Li, Q., Qian, Y., Zhang, Q., Zhai, J., 2012. Preparation and adsorption perfor-
mance of MnO2/PAC composite towards aqueous glyphosate. Environ. Technol.
efficiency at low pH. Given the effectiveness of the SP-PDMAEMA in 33 (17), pp.2049e2056.
removing glyphosate, it should be expected that SP-PDMAEMA European Union, 1991. EU Council Directive 90/414/EEC. European Commission,
would efficiently remove other organophosphorous pesticides Brussels.
Freuze, I., Jadas-Hecart, A., Royer, A., Communal, P.Y., 2007. Influence of complex-
(ex. glufosinate) would through acid-base interactions. ation phenomena with multivalent cations on the analysis of glyphosate and
aminomethyl phosphonic acid in water. J. Chromatogr. A 1175 (2), 197e206.
http://dx.doi.org/10.1016/j.chroma.2007.10.092.
4. Conclusion Fritschi, L., McLaughlin, J., Sergi, C.M., Calaf, G.M., Le Curieux, F., Forastiere, F.,
Kromhout, H., Egeghy, P., Jahnke, G.D., Jameson, C.W., Martin, M.T., 2014. Car-
cinogenicity of tetrachlorvinphos, parathion, malathion, diazinon, and glypho-
Given glyphosate's widespread use, its presence in the envi- sate. Lancet Oncol. http://dx.doi.org/10.1016/S1470-2045(15)70134-8.
ronment and the health concerns associated with its ingestion, it is Gasnier, C., Dumont, C., Benachour, N., Clair, E., Chagnon, M.C., Se ralini, G.E., 2009.
important to find an effective method for its removal from water Glyphosate-based herbicides are toxic and endocrine disruptors in human cell
lines. Toxicology 262 (3), 184e191. http://dx.doi.org/10.1016/j.tox.2009.06.006.
and aqueous solutions. Glyphosate is also a good model compound
Groundwater and Drinking Water Consumer Factsheet on: GLYPHOSATE, 2015. U.S.
representing organophosphate/phosphoric acid-base herbicides to Environmental Protection Agency, Washington, D.C. https://safewater.zendesk.
demonstrate the concept that uniquely designed amine- com/hc/en-us/articles/212076457-4-What-are-EPA-s-drinking-water-regula-
functionalized star polymer can be used to effectively facilitate tions-for-glyphosate- (accessed on 08.12.15).
Guo, Z.X., Cai, Q., Yang, Z., 2005. Determination of glyphosate and phosphate in
their capture. Among all of the adsorbents examined in this study, water by ion chromatographydinductively coupled plasma mass spectrometry
the tailored star polymer (SP-PDMAEMA) comprised of more than detection. J. Chromatogr. A 1100, 160e167. http://dx.doi.org/10.1016/
442 L. Samuel et al. / Chemosphere 169 (2017) 437e442

j.chroma.2005.09.034. based herbicides produce teratogenic effects on vertebrates by impairing reti-


Hamerlinck, Y., Mertens, D.H., Vansant, E.F., 1994. Activated Carbon Principles in noic acid signaling. Chem. Res. Toxicol. 23 (10), 1586e1595. http://dx.doi.org/
Separation Technology. Elsevier, New York. 10.1021/tx1001749.
Ho, M.W., Cherry, B., 2009. Death by multiple poisoning, glyphosate and Roundup. Qian, K., Tang, T., Shi, T., Li, P., Li, J., Cao, Y., 2009. Solid-phase extraction and residue
Sci. Soc. 42, 14. determination of glyphosate in apple by ion-pairing reverse-phase liquid
Homola, A., James, R.O., 1977. Preparation and characterization of amphoteric chromatography with pre-column derivatization. J. Sep. Sci. 32 (14),
polystyrene lattices. J. Colloid Interface Sci. 59 (1), 123e134. http://dx.doi.org/ 2394e2400. http://dx.doi.org/10.1002/jssc.200900118.
10.1016/0021-9797(77)90346-0. Relyea, R.A., 2005. The lethal impact of Roundup on aquatic and terrestrial am-
Hu, Y.S., Zhao, Y.Q., Sorohan, B., 2011. Removal of glyphosate from aqueous envi- phibians. Ecol. Appl. 15 (4), 1118e1124. http://dx.doi.org/10.1890/04-1291.
ronment by adsorption using water industrial residual. Desalination 271 (1), Richard, S., Moslemi, S., Sipahutar, H., Benachour, N., Seralini, G.E., 2005. Differential
150e156. http://dx.doi.org/10.1016/j.desal.2010.12.014. effects of glyphosate and roundup on human placental cells and aromatase.
Jayasumana, C., Gunatilake, S., Senanayake, P., 2014. Glyphosate, hard water and Environ. Health Perspect. 716e720. http://dx.doi.org/10.1289/ehp.7728.
nephrotoxic metals: are they the culprits behind the epidemic of chronic kidney Salman, J.M., Abid, F.M., Muhammed, A.A., 2012. Batch study for pesticide glypho-
disease of unknown etiology in Sri Lanka? Int. J. Environ. Res. Public Health 11 sate adsorption onto palm oil fronds activated carbon. Asian J. Chem. 24 (12),
(2), 2125e2147. 5646.
Lee, V.Y., Havenstrite, K., Tjio, M., McNeil, M., Blau, H.M., Miller, R.D., Sly, J., 2011. Scho€nbrunn, E., Eschenburg, S., Shuttleworth, W.A., Schloss, J.V., Amrhein, N.,
Nanogel star polymer architectures: a nanoparticle platform for modular pro- Evans, J.N., Kabsch, W., 2001. Interaction of the herbicide glyphosate with its
grammable macromolecular self-assembly, intercellular transport, and dual- target enzyme 5-enolpyruvylshikimate 3-phosphate synthase in atomic detail.
mode cargo delivery. Adv. Mat. 23 (39), 4509e4515 doi:10.1002/ Proc. Natl. Acad. Sci. 98 (4), 1376e1380. http://dx.doi.org/10.1073/
adma.201102371 or 10.1002/adma.201190158. pnas.98.4.1376.
Mishra, M., Kobayashi, S., 1999. Star and Hyperbranched Polymers, vol. 53. CRC Solomon, K., Thompson, D., 2003. Ecological risk assessment for aquatic organisms
Press, New York, pp. 201e239. from over-water uses of glyphosate. J. Toxicol. Environ. Health, Part B 6 (3),
Mohan, D., Pittman, C.U., 2006. Activated carbons and low cost adsorbents for 289e324.
remediation of tri-and hexavalent chromium from water. J. Hazard. Mater 137 Speth, T.F., 1993. Glyphosate removal from drinking water. J. Environ. Eng. 119 (6),
(2), 762e811. 1139e1157. http://dx.doi.org/10.1061/(ASCE)0733-9372(1993)119:6(1139).

Ozacar, M., 2003. Equilibrium and kinetic modelling of adsorption of phosphorus on Stalikas, C.D., Konidari, C.N., 2001. Analytical methods to determine phosphonic and
calcined alunite. Adsorption 9 (2), 125e132. http://dx.doi.org/10.1023/A: amino acid group-containing pesticides. J. Chromatogr. A 907 (1), 1e19. http://
1024289209583. dx.doi.org/10.1016/S0021-9673(00)01009-8.
Paganelli, A., Gnazzo, V., Acosta, H., Lo pez, S.L., Carrasco, A.E., 2010. Glyphosate-

You might also like