Environment International 75 (2015) 11–20

Contents lists available at ScienceDirect

Environment International journal homepage: www.elsevier.com/locate/envint

Review

The use of constructed wetlands for removal of pesticides from agricultural runoff and drainage: A review Jan Vymazal ⁎, Tereza Březinová Czech University of Life Sciences Prague, Faculty of Environmental Sciences, Department of Applied Ecology, Kamýcká 129, 165 21 Praha 6, Czech Republic

a r t i c l e

i n f o

Article history: Received 11 August 2014 Received in revised form 28 October 2014 Accepted 30 October 2014 Available online xxxx Keywords: Pesticides Constructed wetlands Soils Drainage Runoff Plants

a b s t r a c t Pesticides are used in modern agriculture to increase crop yields, but they may pose a serious threat to aquatic ecosystems. Pesticides may enter water bodies through diffuse and point sources, but diffuse sources are probably the most important. Among diffuse pollution, surface runoff and erosion, leaching and drainage represent the major pathways. The most commonly used mitigation techniques to prevent pesticide input into water bodies include edge-of-field and riparian buffer strips, vegetated ditches and constructed wetlands. The first attempts to use wetland macrophytes for pesticide removal were carried out as early as the 1970s, but only in the last decade have constructed wetlands for pesticide mitigation become widespread. The paper summarizes 47 studies in which removal of 87 pesticides was monitored. The survey revealed that constructed wetlands with free water surface are the most commonly used type. Also, it has been identified that removal of pesticides is highly variable. The results of the survey revealed that the highest pesticide removal was achieved for pesticides of the organochlorine, strobilurin/strobin, organosphosphate and pyrethroid groups while the lowest removals were observed for pesticides of the triazinone, aryloxyalkanoic acid and urea groups. The removal of pesticides generally increases with increasing value of KOC but the relationship is not strong. © 2014 Elsevier Ltd. All rights reserved.

Contents 1.

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.1. Pathways of pesticide input in aquatic ecosystems . . . . . . . . . . . . . 1.2. Prevention of pesticide input in water bodies . . . . . . . . . . . . . . . 2. The use of constructed wetlands for pesticide mitigation in runoff and drainage waters 3. Efficiency of constructed wetlands in pesticide removal . . . . . . . . . . . . . . 4. Effect of plants on pesticide removal in constructed wetlands . . . . . . . . . . . 5. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1. Introduction 1.1. Pathways of pesticide input in aquatic ecosystems The use of agricultural pesticides results in increased crop yields, but their effects are less than desirable when they leave agricultural ecosystems, especially by entering waterways (Olivier et al., 2012). Widespread ⁎ Corresponding author. E-mail addresses: [email protected] (J. Vymazal), [email protected] (T. Březinová).

http://dx.doi.org/10.1016/j.envint.2014.10.026 0160-4120/© 2014 Elsevier Ltd. All rights reserved.

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

11 11 12 13 13 16 17 18 18

use of pesticides in modern agriculture contributes to agricultural nonpoint source pollution in rivers and streams across the world threating drinking water resources and aquatic ecosystems (Kimbrough and Litke, 1996; Kreuger, 1998; Leu et al., 2004; Schulz, 2004; Jargentz et al., 2005; Probst et al., 2005; Zhang and Zhang, 2011). It has been found that pesticide concentrations in surface waters are related to crop and soil management practices in the catchment (Dabrowski et al., 2002; Zablotowicz et al., 2006; Anderson et al., 2013). Moore et al. (2009) pointed out that despite some efforts to entirely eliminate pesticides from agricultural practices, numerous studies have revealed severe consequences of such action. For example, Oerke et al.

12

J. Vymazal, T. Březinová / Environment International 75 (2015) 11–20

(1994), Oerke (2006) estimated crop yield reduction up to 45–50% without the use of pesticides. However, the increasing use of pesticides for crop production protection may pose a serious environmental concern. Pesticides may enter water bodies through diffuse and point sources, but diffuse sources are probably the most important. Among diffuse pollution, Reichenberger et al. (2007) noted 1) surface runoff and erosion, 2) spray-drift, 3) leaching, 4) drainflows (preferential flow), and 5) other sources such as atmospheric deposition and atmospheric transport of wind eroded soil. Surface runoff is a process by which pesticides are transported in dissolved or particulate forms along the surface of sloping agricultural land (Tang et al., 2012). Surface runoff can, in principle, occur on almost all arable land, even in nearly flat terrain (Wauchope, 1978). Under average conditions, the amount of herbicides lost by movement from soil is typically b 0.1% to 1% of the applied mass (Carter, 2000; Riise et al., 2004) but under certain local conditions, loss can reach up to 5% or greater (Flury, 1996; Carter, 2000). Only for strongly-sorbing pesticides with KOC (Freundlich sorption coefficient normalized to soil organic carbon content) N1000 mg l−1 is erosion considered the main loss pathway (Kenaga, 1980; Wu et al., 2004). Pesticides are typically applied as sprays which are formed when the liquid is atomized through a hydraulic nozzle. Therefore, the fine fraction moved by wind beyond the intended area of application is defined as spray drift (Stephenson et al., 2006). Spray drift is important especially in the areas of permanent crops (Bereswill et al., 2012; Schulz et al., 2001a), and within the last decades, it has been established that aerial pesticide drift is the major contributor to environmental pollution by pesticides besides runoff and erosion (FOCUS, 2007). Pesticide spraydrift phenomenon has been reviewed by Ucar and Hall (2001), Gill and Sinfort (2005), Reichenberger et al. (2007) and Felsot et al. (2010). Leaching is vertical downward displacement of substances through the soil profile and the unsaturated zone, finally reaching groundwater (Reichenberger et al., 2007). Rainfall is a significant factor on pesticide leaching (Flury, 1996; Kladivko et al., 2001; Tiktak et al., 2004) — pesticide leaching generally increases with increasing annual precipitation. Steffens et al. (2013) stressed the importance of temperature-dependent processes on pesticide leaching. Leaching is also affected by organic content of the soil (Larsbo et al., 2013). Transport of pesticides through preferential water flow through macropores to tile drainage plays an important role in the rapid transport of pesticides to surface waters (Kladivko et al., 2001; Brown and van Beinum, 2009; Tang et al., 2012). The main factors affecting the presence of pesticide in drainage waters are soil texture and structure, depth of ground water table, drainage system, pesticide physicochemical properties, rainfall distribution, pesticide application rate and application season (Reichenberger et al., 2007). Other pesticide transport sources include atmospheric deposition after volatilization (Grover et al., 1985; Messing et al., 2013), atmospheric transport of wind eroded soil (Larney et al., 1999) and point-sources such as farmyards, storage facilities or roads (Reichenberger et al., 2007).

1.2. Prevention of pesticide input in water bodies The most commonly used mitigation techniques to prevent pesticide input into water bodies include edge-of-field and riparian buffer strips, vegetated ditches, and constructed wetlands (CWs). These Best Management Practices have been shown to be effective in pesticide removal. Filter strips are narrow strips of permanent vegetation widely prescribed to reduce contaminants in surface runoff from adjacent agricultural fields (Schmitt et al., 1999; Otto et al., 2008). A vegetative barrier is a strip of dense, tall, stiff grass that functions like a porous dam to temporarily pond runoff water, settle its sediment load, and gradually release water downslope (Dosskey, 2001). Grassed edge-of-field buffer strips have extensively been reviewed by Dosskey (2001), Lacas et al.

(2005), Krutz et al. (2005), Reichenberger et al. (2007) and Liu et al. (2008). Grassed waterways are strips of buffer installed primarily to convey excess surface runoff from fields to the field margin without causing gully erosion. Dense, low-growing grasses are used to stabilize and protect the soil surface against erosion by concentrated runoff (Dosskey, 2001). To serve this function effectively, there is usually a selection of fast growing grasses which are mowed frequently in order to prevent sward-damaging sedimentation and to reduce hydraulic roughness (Ree, 1949; Asmussen et al., 1977; Fiener and Auerswald, 2003). Also, grassed waterways reduce the velocity of overland flow and reduce peak discharge rate (Chow et al., 1999). Bennett et al. (2005) pointed out that historically, the values and function of agricultural ditches have been ignored except for occasional dredging to remove built-up sediments and plants impeding efficient drainage. Recently, it has been shown in several studies that agricultural ditches with wetland vegetation may be useful traps for commonly used pesticides (Moore et al., 2001a, 2008; Bennett et al., 2005; Roessink et al., 2005; Dabrowski et al., 2006; Gill et al., 2008; Rogers and Stringfellow, 2009; Anderson et al., 2011). Riparian wetlands are ecosystems in which soils and soil moisture are influenced by the adjacent stream or river. Riparian ecosystems act as a nutrient sink for lateral runoff from uplands and therefore, these wetlands are important buffer zones between agricultural lands and streams (Mitsch and Gosselink, 2000). Riparian buffer zones have long been known for effective reduction of nitrate loads (Peterjohn and Correll, 1984; Lowrance, 1992; Gilliam, 1994; Blicher-Mathiesen and Hoffmann, 1999), but these areas have also been evaluated in terms of pesticide removal (Pavel et al., 1999; Dosskey, 2001; Liu et al., 2008; Kidmose et al., 2010; Ohliger and Schulz, 2010; Bereswill et al., 2012; Lizotte et al., 2012; Karpuczu et al., 2013). CW treatment systems are engineered systems that have been designed and constructed to utilize the natural processes involving wetland vegetation, soils, and their associated microbial assemblages to assist in treating wastewater. They are designed to take advantage of many of the processes that occur in natural wetlands, but do so within a more controlled environment (Vymazal and Kröpfelová, 2008). Recently, focus has been paid to CWs which are very effective in pesticide removal. Pesticides are removed in CWs by physical (sedimentation, flocculation, absorption, co-precipitation, precipitation), chemical (oxidation, reduction, cation exchange, hydrolysis, photolysis), biological (plant absorption and metabolism) or biochemical processes (microbial degradation) (Vink and Van der Zee, 1997; Stangroom et al., 2000; Runes et al., 2003; Imfeld et al., 2009). CWs have been used for treatment of various types of wastewater for more than five decades (Vymazal, 2011) and there are many types of CWs that could be categorized according to free water surface absence or presence, direction of flow, or types of macrophytes used. Constructed wetlands with surface flow or free water constructed wetlands (FWS CWs) consist of basins or channels, with soil or another suitable medium to support the rooted vegetation and water at a relatively shallow depth flowing through the unit. The shallow water depth, low flow velocity, and the presence of the plant stalks and litter regulate water flow and, especially in long, narrow channels, ensure plug-flow conditions (Reed et al., 1988). Most treatment processes occur in the water column or in the litter layer on the bottom. The FWS CWs can use all types of macrophytes, i.e. free floating, submerged, floating leaved and emergent (Vymazal and Kröpfelová, 2008; Kadlec and Wallace, 2009). CWs with subsurface flow may be classified according to the direction of flow to horizontal flow (HF CW) and vertical (VF CW). In horizontal flow CWs, wastewater is continuously fed in at the inlet and flows slowly through the porous medium under the surface of the bed in a more or less horizontal path until it reaches the outlet zone where it is collected before leaving via level control arrangement at the outlet. During this passage wastewater will come into contact with a network of aerobic, anoxic and anaerobic zones. Aerobic zones occur around

J. Vymazal, T. Březinová / Environment International 75 (2015) 11–20

roots and rhizomes that leak oxygen into the substrate but the filtration bed is mostly anoxic or even anaerobic (Brix, 1987). Vertical flow systems usually consist of a filtration bed filled with graded gravel or sand planted with macrophytes. VF CWs are fed intermittently with a large batch of wastewater which then gradually percolates down through the bed and is collected by a drainage network at the base. The bed drains completely free and it allows air to refill the bed. This kind of dosing leads to good oxygen transfer that supports nitrification (Cooper, 2005). VF CWs may also be fed from the bottom with the treated water being discharged from the filtration bed surface. However, such filters are water-saturated and the filtration bed is mostly anoxic. The combination of various types of CWs in a stage manner is called a hybrid CW (Vymazal, 2013). The most common system is the combination of VF and HF, where each stage provides different redox conditions. Also, the inclusion of FWS CWs in a hybrid CW has recently been used more frequently (Vymazal, 2013). As part of an integrated management system, CWs can be used in conjunction with other edgeof-field measures such as vegetative buffer strips (Dabney et al., 2006). The use of constructed wetlands was reviewed by Stehle et al. (2011). In this review, 22 constructed wetlands and nine vegetated ditches in six countries were included with 34 pesticides being assessed. However, since then the number of studies on the use of constructed wetlands for pesticide mitigation has grown substantially and therefore, the objective of this paper is to provide updated survey of the use of CWs. The survey summarizes 47 studies from 35 constructed wetlands on the use of various types of CWs in Australia, Brazil, Canada, China, Colombia, France, Norway, Portugal, Spain, South Africa, Suriname, United Kingdom and USA. In the survey, 87 pesticides, including 35 herbicides, 27 fungicides and 25 insecticides are presented (Table 1). 2. The use of constructed wetlands for pesticide mitigation in runoff and drainage waters CWs have been used for treatment of various types of wastewater since the 1950s (Vymazal, 2009, 2011), however, pesticide removal in CWs was evaluated much later. During the early 1970s, microcosm experiments were carried out to evaluate the effect of plants on mevinphos removal from water. Results revealed that in microcosms planted with Nymphaea odorata and Paspalum distichum, the pesticide was undetected after 12 days, while in microcosms with only fresh water and those with freshwater and soil (no vegetation), 46% and 14% of the initial concentrations were recorded after 12 days, respectively (Wolverton, 1975; Wolverton and Harrison, 1974). In 1988, a large series of experimental wetlands were built on the Biological Field Station of the University of Mississippi near Oxford, Mississippi (Rodgers and Dunn, 1992) where numerous experiments

Table 1 List of pesticides monitored in constructed wetland used for pesticide mitigation included in the survey. Herbicides: Acetochlor, aclonifen, alachlor, ametryn, atrazine, bentazone, chlorotoluron, dicamba, dichlorprop, diflufenican, diuron (and its degradation product 3,4-DCA), ethofumesate, fluometuron, fluroxypyr, gluphosinate, glyphosate (and its degradation product AMPA), isoproturon, isoxaben, linuron, MCPA, mecoprop, mefenpyr-diethyl, metamitron, metazachlor, metolachlor, metribuzin, napropamide, pendimethalin, pentachlorophenol, propachlor, prosulfocarb, simazine, terbuthylazine, 2,4-D Insecticides: Aldicarb, azinphos-methyl, bifenthrin, chlorpyrifos, cyfluthrin, λ-cyhalothrin, cypermethrin, diazinon, dimethoate, endosulfan, esfenvalerate, fipronil, flufenoxuron, imidacloprid, indoxacarb, lindane, methyl parathion, mevinphos, omethoate, parathion, pentachlorophenol, permethrin, prothiofos, thiacloprid, triflumuron Fungicides: Azoxystrobin, carbendazim, chlorothalonil, cyazofamid, cymoxanil, cyproconazole, cyprodinil, difenoconazole, dimethomorph, epoxiconazole, fenpropidine, fenpropimorph, fludioxamine iprodione, kresoxim methyl, metalaxyl, penconazole, pencycuron, penflufen, pentachlorobenzene, pentachlorophenol, propiconazole, pyrimethanil, spiroxamine, tebuconazole, tetraconazole, trifloxystrobin

13

were carried out during the late 1990s and the early 2000s (e.g. Moore et al., 2000, 2001b; Milam et al., 2004). During the early 1990s, attempts were made to evaluate atrazine removal and fate in the Des Plaines River Wetlands Demonstration Project where four CWs were built to treat about 40% of the average stream flow in Wadsworth, Illinois, USA (Kadlec and Hey, 1994; Alvord and Kadlec, 1996). Atrazine removal was also studied in a vertical flow wetland mesocosms in the United Kingdom (McKinlay and Kasperek, 1999). Since these “pioneer” experiments, numerous studies on pesticide mitigation in CWs have been carried out (Table 2). The literature survey revealed that free water surface CWs have been mostly used thus far (Table 2). Subsurface flow CWs have been used less frequently. Unfortunately, there are no direct comparative results available from different types of CWs and, therefore, it is not possible to determine the most effective CW type.

3. Efficiency of constructed wetlands in pesticide removal In the following section, constructed wetland efficiency for pesticide removal is discussed. Physicochemical properties of pesticides reported in the literature on the use of constructed wetlands for pesticide removal are presented in Table S1. The detailed evaluation of the removal of pesticides which have been reported at least in two studies is presented in Supplementary material S2. Pesticide distribution among different environmental compartments (water, soil, and plants) is a complex process affected by pesticide physico-chemical characteristics such as water solubility, soil/ water partition coefficient (KOC), octanol/water partition coefficient (KOW), pesticide half-life on soil (T50) and water (T50), soil and water photolysis (Stangroom et al., 2000; Carter, 2000; Poissant et al., 2008; Gregoire et al., 2009; Tournebize et al., 2011). Pesticides are removed/ retained in CWs through many processes, but the most important processes are probably sedimentation, photolysis, hydrolysis, adsorption, microbial degradation and plant uptake. The extent of processes involved in pesticide removal in CWs depends on many factors such as organic matter content, clay content, filtration material quality, pH, redox conditions, the presence and/or absence of water, retention time, inflow pesticide mass, the presence and type of macrophytes, or type of CW. Maillard et al. (2011) pointed out that intermittent flow conditions and hydrochemical characteristics are key variables that control removal mechanisms such as sedimentation and degradation in CWs. However, it is not easy to single out individual processes as the removal processes are closely connected. Also, Bouldin et al. (2006) pointed out that investigations failed to distinguish between macrophytespecific degradation of pesticides and remediation through associated pathways where sediment, organic matter, and microbial action combine to form a complex dynamic of pesticide remediation. Wetland hydrology plays an important role in pesticide removal in constructed wetlands (Gregoire et al., 2009), especially hydraulic retention time is an important factor in pesticide retention in the wetland. Sherrard et al. (2004) described a steady decrease in chlorothalonil concentration with time over the period of 72 h in a mesocosm experiment. Moore et al. (2002) observed a sharp decrease in chlorpyrifos concentration in water within the first week with simultaneous increase in chlorpyrifos concentration in the sediment. Moore et al. (2009) found that during the 55 days of experiment the concentration of λ-cyhalothrin steadily decreased during the experiment and was not detected after 27 days. Locke et al. (2011) observed a steady decrease in atrazine concentration over the period of 24 h while fluometuron concentration sharply dropped after 1 h of exposure. On the other hand, Stearman et al. (2003) observed only slight increase in simazine removal with increasing hydraulic retention time in a subsurface constructed wetland. For metolachlor, the influence of hydraulic retention time was observed between 2.3 and 5.1 days but further increase in retention time did not result in higher removal of metolachlor. The positive effect on pesticide

14

Table 2 The use of constructed wetlands for pesticide removal from agricultural runoff and drainage. Location

Type

USA, Mississippi

FWS

Alvord and Kadlec (1996)

USA, Illinois

FWS

McKinlay and Kasperek (1999) Moore et al. (2000) Moore et al. (2001b)

U.K.

VF

3 wetlands 1.86 ha, 2.22 ha, 2.33 ha 3.4 × 0.7

USA, Mississippi

FWS

14 × 59–73

South Africa

FWS

36 × 134

Moore et al. (2002) Schulz et al. (2001b)

Size (m/ha)

Sediment/filter material

Plants

Pesticide

Type of runoff

Soil

Nymphaea odorata, Juncus repens, Paspalum distichum Typha latifolia, Polygonum amphibium, Nymphaea tuberosa

Mevinphos

Microcosm

Atrazine

River water

Gravel, 20 mm and 6–11 mm

Schoenoplectus lacustris, Iris pseudacorus, Typha latifolia, Phragmites australis

Atrazine

Mesocosm study

Typha latifolia, Scirpus cyperinus, Zizania aquatica Juncus effusus, Leersia sp., Ludwigia sp. Typha capensis (60%), Juncus kraussii (10%), Cyperus dives (5%)

Atrazine Metolachlor

Simulated runoff

84% sand, 16% silt

Schulz and Peall (2001) Moore et al. (2002) Schulz et al. (2003) Cheng et al. (2002)

Germany

VF

1.0 m2

Gravelsand 0–8 mm)

George et al. (2003) Stearman et al. (2003) Braskerud and Haarstad (2003)

USA, Tennessee

HF

4.9 × 1.2 (2.4)

Norway

FWS

840 m2 (100 m long)

Quartz gravel 1.1 cm and 1.9 cm fractions 18–37% LOI

Runes et al. (2003)

USA, Oregon

FWS

3 × 200

Kohler et al. (2004) Sherrard et al. (2004)

USA, Indiana USA,

FWS FWS (batch)

1.95 ha 1.85 × 0.63

Bouldin et al. (2005) Moore et al. (2006) Rose et al. (2006)

USA, Akansas USA, Mississippi Australia

FWS FWS FWS

0.41 × 0.51 50 × 5.5 20 × 10 vegetated 10 × 10 unvegetated

Matamoros et al. (2007)

Spain

HF

55 m2

Blankenberg et al. (2007)

Norway

FWS

40 × 3

Moore et al. (2007)

USA, Mississippi

FWS

30 × 180

Moore et al. (2009) Locke et al. (2011)

Typha capensis (80%), Juncus kraussii (15%), Cyperus dives (5%) Colocasia esculenta, Ischaemum aristatum var. glaucum Scirpus validus

Chlorpyrifos Azinphos-methyl

Orchard runoff

Azinphos-methyl chlorpyrifos, prothiofos Chlorpyrifos Azinphos-methyl Parathion, omethoate, MCPA, dicamba

Mesocosm study

Simazine, metolachlor

Container nursery runoff

Sparganium erectum, Phragmites australis, Propachlor, metribuzin, linuron, Phalaris arundinacea, Myosotis metamitron, metalaxyl, propiconazole, scorpioides, Urtica dioica fenpropimorph, mecoprop, dicamba, MCPA, dichlorprop, bentazone, fluroxypyr 10% sand, 70% silt, 20% clay, Typha latifolia Atrazine 2.13% TOC Planted with 10,800 plants of 18 species Mixture Sand amended with Scirpus cyperinus Chlorothalonil, chlorpyrifos compost Ditch sediment Juncus effusus, Ludwigia peploides Atrazine, λ-cyhalothrin Juncus effusus, Ludwigia peploides Methyl parathion Persicaria spp., Ludwigia peploides, Fluorometuron, diuron, aldicarb, Myriophyllum papillosum, Juncus usitatus, endosulfan Bolboschoenus medianus, Typha domingensis Gravel (3.5 mm) Phragmites australis Simazine, alachlor, chlorpyrifos, pentachlorobenzene, pentachlorophenol, endosulfan, lindane, diuron, mecoprop Phalaris arundinacea, Glyceria fluitans, Metalaxyl, metamitron, metribuzin, Typha latifolia, Sparganium erectum, propachlor, linuron, fenpropimorph Phragmites australis Cyperus iria, Sorghum halepense, Digitaria Lambda-cyhalothrin, cyfluthrin ischaemum, Polygonum lapathifolium, Alternanthera philoxeroides Diazinon Alternanthera philoxeroides, Cephalanthus Atrazine, fluorometuron occidentalis, Echinochloa colona, Echinochloa crus-galli (major species)

Agricultural watershed runoff

Simulated runoff in container nursery Golfcourse runoff Laboratory experiment Microcosm study Simulated runoff Cotton field tailwater

Pilot plant (pesticide mixture injected to sewage) Simulated runoff

Simulated runoff

Simulated runoff

J. Vymazal, T. Březinová / Environment International 75 (2015) 11–20

Reference Wolverton (1975)

Rose et al. (2008) Beketov and Liess (2008) Lin et al. (2008) Lizotte et al. (2009)

Australia Germany China USA, Mississippi

FWS FWS VF (batch) FWS

20 × 0.32 0.3 × 0.3 700 × 25

Gravel/sand 0.2–3.7 mm Gravel 8–15 mm Modified natural wetland

Nasturtium officinale Typha latifolia

Borges et al. (2009) Gregoire et al. (2009) Budd et al. (2009, 2011)

Brazil France USA, California

HF FWS FWS

24 × 1.0 990 m2 2.3 ha and 2.5 ha

Gravel (D60 = 7 mm)

Typha latifolia

Agudelo et al. (2010) Page et al. (2010)

Colombia Australia

HF FWS

1.0 × 0.6 11 ha

Igneous rock, 3.9–6.4 mm

Elsaesser et al. (2011)

Norway

FWS

40 × 3

Maillard et al. (2011)

France

FWS/HF

234/104

Paspalum distichum, Polygonum lapathifolium, Echinochloa crus-galli

Clay 44%, fine silt 33%, coarse silt 10%, fine sand 5%, coarse sand 8%/gravel

Moore et al. (2013) Dordio and Carvalho (2013) Tournebize et al. (2013)

USA, Misissippi Portugal France

FWS Not specified

1.3 × 0.7 0.62 m

2

Silt loam, sand Gravel, LECA

2

FWS, in-stream

4165 m

FWS off-stream

1280 m2

Phragmites australis (90%), Juncus effuses, Typha latifolia Typha latifolia, Leersia oryzoides, Sparganium americanum Phragmites australis Agrostis stolonifera, Nasturtium officinale, Veronica baccabunga, Lysimachia nummularia

Dicamba, dimethoate, trifloxystrobin, metamitron, tebuconazole AMPA, azoxystrobin, cymoxanil, cyprodinil, carbendazim, dimethomorph, diuron, flufenoxuron, gluphosinate, glyphosate, isoxaben, kresoxim methyl, metalaxyl, pyrimethanil, simazine, terbuthylazine, tetraconazole Glyphosate, AMPA

Microcosm study Stormwater runoff

Simulated runoff Vineyard runoff

Atrazine, diazinon, permethrin

Mesocosm study

MCPA

Mesocosm study

Isoproturon, metazachlor, S-metolachlor, chlorotoluron, iprodione, azoxystrobin, tebuconazole, napropamide, epoxiconazole, prosulfocarb, pendimethalin, diflufenican, aclonifen

Agricultural runoff

Glyceria maxima (53%), Festuca arundinacea (12%), Phragmites australis (10%), Phalaris arundinacea (9%)

Passeport et al. (2013)

Bois et al. (2013) Yang et al. (2013)

France USA, Ohio

FWS FWS/VF

0.39 × 0.24 m 6.8 m2/7.1 m2

Sand (0–4 mm), sediment Gravel/soil

Elsaesser et al. (2013)

Germany

FWS

45 × 0.4

Not specified

Stang et al. (2014) Elsayed et al. (2014)

France

VF upflow

15 cm diameter

Suriname

FWS

0.6 × 0.3

Gravel (0.1–2 mm), sand (0.4–0.63 mm) Sandy loam sediment and potting soil

Mahabali and Spanoghe (2014)

Ametryn Glyphosate, penconazole Bifenthrin, λ-cyhalothrin, cypermethrin, esfenvalerate, diazinon, permethrin, chlorpyrifos Chlorpyrifos Simazine, diuron, atrazine

Field runoff Mesocosm system Laboratory experiment Simulated runoff from 16 ha agriculture field Mesocosm study Demonstration system Agricultural tailwater

Aclonifen, atrazine, chlorothalonil, chlorotoluron, cyproconazole, diflufenican, epoxiconazole, ethofumesate, fenpropidine, isoproturon, mefenpyr-diethyl, metazachlor, napropamide, prosulfocarb, S-metazachlor, tebuconazole, Diuron, 3,4-DCA, glyphosate Laboratory experiment Atrazine, glyphosate, dicamba, 2,4-D Mesocosm experiment

Phragmites australis Eupatorium perfoliatum, Tradescantia ohiensis, Veronicastrum virginicum, Eragrostis spectabilis, Sorghastrum nutans, Echinacea purpurea Elodea nuttallii Indoxacarb, tebuconazole, thiacloprid, trifloxystrobin Pencycuron, penflufen, triflumuron Phragmites australis Metolachlor, alachlor, acetochlor Nymphaea amazonum, Eleocharis mutata

λ-Cyhalothrin, imidacloprid

J. Vymazal, T. Březinová / Environment International 75 (2015) 11–20

Imfeld et al. (2013)

Phragmites australis Phragmites australis, Eleocharis sphacelata, Schoenoplectus validus, Baumea articulata, Typha orientalis Phalaris arundinacea, Typha latifolia, Phragmites australis Phragmites australis, Schoenoplectus lacustris, Typha latifolia/Lolium perenne

Endosulfan, fluorometuron Thiacloprid Atrazine Atrazine, S-metolachlor, fipronil

Mesocosm facility

Laboratory experiment Mesocosm

15

16

J. Vymazal, T. Březinová / Environment International 75 (2015) 11–20

removal was also observed by Blankenberg et al. (2007) or Haarstad and Braskerud (2005) in Norway. The hydrological conditions differ very much in various types of constructed wetlands. Free water surface CWs are mostly aerobic with anaerobic conditions occurring only at the bottom in the layer of decomposing plant material. The pesticides may be degraded by aerobic hydrolysis and photolysis as well. The contact with soil is limited but FWS CWs use soil as a substratum so adsorption of pesticides in the sediment is likely. If plants absorb pesticides directly from the water, FWS CWs provide a suitable condition for this process. On the other hand in HF CWs, the conditions are mostly anoxic/anaerobic due to continuous submergence of the filtration bed. The water level is kept below the surface so photolysis and aerobic hydrolysis are limited. Also, HF CWs use washed gravel of crushed rock, i.e. filtration media which do not contain organic matter and therefore adsorption to organics is limited and could occur only in older mature systems where organic matter concentration increases due to sedimentation of suspended solids and formation on biofilms. The exposure of pesticides in HF CWs to plant roots is high and therefore, the potential for pesticides to be taken up is high. The results of the survey revealed that removal of pesticides is highly variable for individual pesticides, however, some trends could be seen when pesticides are grouped according to the chemical basis of the compounds or removal is related to physicochemical parameters of the pesticides. The results presented in Fig. 1 clearly indicate that pesticides of certain substance groups are removed more efficiently than the others. The highest average removals (97%) were achieved for pesticides from organochlorine group (endosulfan, pentachlorophenol), strobilurin/ strobin group (96%, kresoxim methyl, trifloxystrobin and azoxystrobin) followed by organophosphate pesticides (94%, azinophos methyl, diazinon, dimethoate, glufosinate, chlorpyrifos, methyl parathion, mevinphos, omethoate, parathion, prothiofos) and pyrethroids (84%, bifenthin, cyhalothrin, cypermethrin, esfenvalerate, permethrin). Most of these pesticides have very low solubility, very high KOW and KOC coefficients, namely pyrethroids and strobilurins (Table S1). On the other hand, the lowest removals were achieved for triazinone pesticides (24%, metamitron, metribuzin), aryloxyalkanoic acid group (35%, dichlorprop, MCPA, mecoprop) and urea-based pesticide (50%, diuron, fluorometuron, chlorotoluron, isoproturon, linuron). The poorly removed pesticides do not show any clear relationship between removal and solubility, KOW and KOC and these parameters are highly variable (Table S1). It has been suggested that pesticides with high K OC value (N 1000 mg l−1, i.e. log KOC = 3.0) may strongly adsorb to soil particles (Wu et al., 2004; Poissant et al., 2008) and therefore their removal in 120

80

4. Effect of plants on pesticide removal in constructed wetlands It has been shown that dense vegetation increases the effectivity of pesticide removal (Wolverton, 1975; Braskerud and Haarstad, 2003; Rogers and Stringfellow, 2009; Vallée et al., 2014). In general, removal of pesticides through plant uptake and sorption through the root system could be expected for systemic herbicides. Plants may increase pesticide removal either directly through uptake or indirectly through associated periphyton and humic contribution (Brock et al., 1992; Stomp et al., 1994; Karen et al., 1998; Schulz and Peall, 2001; Olette et al., 2008). The ability of wetland plants to take up and cumulate pesticides was shown in several early studies under laboratory conditions. Hinman and Klaine (1992) observed rapid uptake of atrazine, lindane and chlordane by macrophyte Hydrilla verticillata but the equilibrium between shoots and water was reached much faster for atrazine (2 h) then for lindane (24 h) and chlordane (144 h). Feurtel-Mazel et al. (1996) found rapid accumulation of isoproturon by Elodea densa and Ludwigia natans with bioaccumulation being concentration dependent. Crum et al. (1999) studied sorption of nine pesticides by macrophytes Elodea nuttallii and Lemna gibba and macroalga Chara globularis and found a strong relationship between sorption coefficient and pesticide solubility in water. Guo et al. (2014) have observed that less hydrophobic (low KOW) organochlorine pesticides were more easily accumulated and transported in the tissues of Phragmites australis, Typha sp. and Ceratophyllum demersum. In one of the earliest CW studies, Wolverton (1975) observed that microcosms planted with N. odorata and P. distichum removed 100% of the insecticide mevinphos after two weeks of exposure, while microcosms without plants and soil removed only 26%, and microcosms with soil only removed 94% of mevinphos. The HF CW planted with Scirpus validus removed more metolachlor (62%) as compared to an identical filter without plants (34%) from a container nursery runoff in

60 40 20 0

Removal (%)

Removal (%)

100

CWs could be high. However, the results shown in Fig. 2 indicate that the relationship between log KOC and removal efficiency is not strong. However, there is a clear tendency for more efficient removal of pesticides with higher KOC. Also, Vallée et al. (2014) observed in their laboratory studies that the pesticide retention was greater for pesticides with hydrophobic properties (low solubility and high KOC). The KOC values are inversely related to pesticide solubility, and, therefore it is not surprising that the relationship between pesticide solubility and pesticide removal exhibits an opposite trend (Fig. 3). The relationship, however, is not strong either. In Fig. 4, relationship between pesticide aerobic half-time and removal pesticide removal is shown. However, no clear relationship could be drawn. Elsaesser et al. (2013) proposed that plant density and pesticide solubility had the highest explanatory power for the pesticide concentration reduction. Stehle et al. (2011) concluded based on the analysis of literature data that the most powerful predictors of pesticide retention were KOC and plant coverage followed by water-phase dissipation time.

100 90 80 70 60 50 40 30 20 10 0

y = 31.98ln(x) + 32.43 R² = 0.162

1

2

3

4

5

Log KOC Fig. 1. Removal of pesticides according to the pesticide chemical groups. Numbers in parentheses indicate number of pesticides in each group.

Fig. 2. Relationship between KOC and removal of pesticides.

6

Rdemoval (%)

J. Vymazal, T. Březinová / Environment International 75 (2015) 11–20

100 90 80 70 60 50 40 30 20 10 0

y = -2.48ln(x) + 74.88 R² = 0.162

0

100

200

300

400

500

600

Solubility in water (mg/L) Fig. 3. Relationship between pesticide solubility in water (at 20 °C) and pesticide removal.

Removal (%)

Tennessee, USA (George et al., 2003). Dordio and Carvalho (2013) noted that wetland mesocosms planted with P. australis removed up to 28% more MCPA than unvegetated units. Also, Leersia oryzoides (45%, 88%) and Typha latifolia (35%, 88%) reduced significantly more overall atrazine and trans-permethrin loads as compared to unvegetated controls (13%, 68%), respectively (Moore et al., 2013). Beketov and Liess (2008) observed that concentration of the insecticide Thiacloprid decreased by 20–60% more in a mesocosm section planted with Nasturtium officinale as compared to open water part of the mesocosm. The amount of pesticides retained in plants is highly variable. Moore et al. (2001b, 2002, 2009) found 10% of measured metolachlor, 25% of applied chlorpyrifos, 49% of λ-cyhalothrin and 76% of cyfluthrin mass were associated with vegetation in a CW in Mississippi, USA, respectively. The major process responsible for endosulfan dissipation in a CW was alkaline hydrolysis and sedimentation with hydrolysis being reduced and sedimentation enhanced by the presence of plants (Rose et al., 2008). On the other hand, the same study demonstrated that fluorometuron removal occurred through biofilm reaction and photolysis, both being enhanced by plants. Elsaesser et al. (2011) reported that wetland cells with vegetation reduced peak pesticide concentration more effectively (up to 91%) than cells without vegetation (72%) in a field study carried out in Norway. However, uptake by plants was low (up to 4%) with cells with Phalaris arundinacea being more effective than cells with T. latifolia as dominant species. Schulz et al. (2003) observed more effective removal of methyl parathion in vegetated wetlands as compared to cells without vegetation. Runes et al. (2003) reported that the presence of plants did not affect the treatment of pesticides from irrigation runoff in Oregon, USA. Recently, Mahabali and Spanoghe (2014) observed very high absorption of pesticide imidacloprid (79%) by Nymphaea amazonum with majority of pesticide being found in leaves and shoots. On the other hand, Eleocharis mutata absorbed only about 15%. For another pesticide, λ-cyhalothrin, both

100 90 80 70 60 50 40 30 20 10 0 0

20

40

60

80

100

Aerobic soil half-me (d) Fig. 4. Relationship between pesticide aerobic soil half-time and pesticide removal.

17

plants absorbed less than 1% of the pesticide in a mesocosm experiment in Suriname. Concentration of pesticides in plants increases shortly after the contact with pesticide and remains stable or decreases over the time Locke et al. (2011). The authors observed five-fold increase in atrazine concentration in plant material in the first 24 h and then the concentration remained stable for the rest of the experiment (19 days). On the other hand fluorometuron concentration in plants was the highest after 1 h and then sharply decreased and remained stable for the rest of the experiment. Friesen-Pankratz et al. (2003) found in a laboratory study that unicellular algae decreased the aqueous persistence of atrazine and lindane probably due to either pesticide sorption or pesticide degradation facilitation. The authors pointed out that algae should be taken into account in management of CWs designed for pesticide removal. Rose et al. (2006) observed that aldicarb and endosulfan were no longer quantifiable following the occurrence of algal bloom in an open water of the CW. Based on their research Rose et al. (2006) concluded that CWs designed for pesticide retention should comprise of both open water and vegetated zones, to increase the potential for complementary chemical, photolytic, microbial and plant-mediated pesticide breakdown. However, Maillard and Imfeld (2014) pointed out that the wetland vegetation enhanced the pesticide degradation during the growing season but the pesticides were released during the plant decay. The authors also concluded that the prevailing storage compartment and dissipation processes vary during the year. While during the spring plant uptake was the major dissipation process, during the summer and late summer, biodegradation and sorption on bed sediment prevailed, respectively. The respective storage compartments were bed sediment and plants, water and suspended solids and bed sediment. Overall, bed sediment was the major storage pool and biodegradation was the major dissipation process. Also, Stang et al. (2014) observed that the sorption to macrophyte E. nuttallii was, in addition to dispersion, the second most important process causing the reduction of the peak pesticide concentrations. The adsorption of pesticides to E. nuttallii was assessed as a reversible process with desorption occurring after the peak concentrations passed the wetland.

5. Summary CWs have become the best management practice for pesticide mitigation from non-point source agricultural runoff and drainage in many countries. So far, CWs with free water surface have been primarily used. The subsurface flow CWs, both vertical and horizontal, have recently been used as well. However, there is no side by side experiment which would compare various types of CWs at one location. As both aerobic and anaerobic processes are involved in pesticide removal hybrid constructed wetlands may offer efficient solution. Current survey indicated that removal of pesticides is generally effective, but the efficiency varies widely among pesticides and also among systems for a particular pesticide. There are many processes which are responsible for pesticide mitigation such as hydrolysis, photolysis, sedimentation, adsorption, microbial degradation or plant uptake, however, the extent of these processes depends on local conditions, and it is difficult to single out the most important ones. There is strong evidence to suggest that the presence of vegetation enhances pesticide retention. The results of the survey revealed that the highest pesticide removal was achieved for pesticides of the organochlorine, strobilurin/strobin, organosphosphate and pyrethroid groups while the lowest removals were observed for pesticides of the triazinone, aryloxyalkanoic acid and urea groups. The removal of pesticides generally increases with increasing value of KOC but the relationship is not strong. Supplementary data to this article can be found online at http://dx. doi.org/10.1016/j.envint.2014.10.026.

18

J. Vymazal, T. Březinová / Environment International 75 (2015) 11–20

Acknowledgments The survey was supported by grant no. TA 04020512 from the Technology Agency of the Czech Republic.

References Agudelo, R.M., Peňuela, G., Aguirre, N.J., Morató, J., Jaramillo, M.L., 2010. Simultaneous removal of chlorpyrifos and dissolved organic carbon using horizontal sub-surface flow pilot wetlands. Ecol. Eng. 36, 1401–1408. Alvord, H.H., Kadlec, R.H., 1996. Atrazine fate and transport in the Des Plaines Wetlands. Ecol. Model. 90, 97–107. Anderson, B., Phillips, B., Hunt, J., Largay, B., Shihadeh, R., Tjeerdema, R., 2011. Pesticide and toxicity reduction using an integrated vegetated treatment system. Environ. Toxicol. Chem. 30, 1036–1043. Anderson, T.A., Salice, C.J., Erickson, R.A., McMurry, S.T., Cox, S.B., Smith, L.M., 2013. Effects of landuse and precipitation on pesticides and water quality in playa lakes of the southern high plains. Chemosphere 92, 84–90. Asmussen, L.E., White Jr., A.W., Hauser, E.W., Sheridan, J.M., 1977. Reduction of 2,4D load in surface runoff down a grassed waterway. J. Environ. Qual. 6, 159–162. Beketov, M.A., Liess, M., 2008. Variability of pesticide exposure in a stream mesocosm system: macrophyte-dominated vs. non-vegetated sections. Environ. Pollut. 156, 1364–1367. Bennett, E.R., Moore, M.T., Cooper, C.M., Smith Jr., S., Shields Jr., F.D., Drouillard, K.G., Schulz, R., 2005. Vegetated agricultural drainage ditches for the mitigation of pyrethroid-associated runoff. Environ. Toxicol. Chem. 24, 2121–2127. Bereswill, R., Golla, B., Streloke, M., Schulz, R., 2012. Entry and toxicity of organic pesticides and copper in vineyard streams: erosion rills jeopardise the efficiency of riparian buffer strips. Agric. Ecosyst. Environ. 146, 81–92. Blankenberg, A.-G.B., Haarstad, K., Braskerud, B.C., 2007. Pesticide retention in an experimental wetland treating non-point source pollution from agricultural runoff. Water Sci. Technol. 55, 37–44. Blicher-Mathiesen, G., Hoffmann, C.C., 1999. Denitrification as a sink for dissolved nitrouoxide in a freshwater riparian fen. J. Environ. Qual. 28, 257–262. Bois, P., Huguenot, D., Jézéquel, K., Lollier, M., Cornu, J.Y., Lebeau, T., 2013. Herbicide mitigation in microcosms simulating stormwater basins subject to polluted water inputs. Water Res. 47, 1123–1135. Borges, A.C., do Carmo Calijuri, M., Teixeira de Matos, A., Ribeiro de Queiroz, M.E.L., 2009. Horizontal subsurface flow constructed wetlands for mitigation of ametryncontaminated water. Water SA 35, 441–445. Bouldin, J.L., Farris, J.L., Moore, M.T., Smith Jr., S., Stephens, W.W., Cooper, C.M., 2005. Evaluated fate and effects of atrazine and lambda-cyhalothrin in vegetated and unvegetated microcosms. Environ. Toxicol. 20, 487–498. Bouldin, J.L., Farris, J.L., Moore, M.T., Smith Jr., S., Cooper, C.M., 2006. Hydroponic uptake of atrazine and lambda-cyhalothrin in Juncus effusus and Ludwigia peploides. Chemosphere 65, 1049–1057. Braskerud, B.C., Haarstad, K., 2003. Screening the retention of thirteen pesticides in a small constructed wetland. Water Sci. Technol. 48, 267–274. Brix, H., 1987. Treatment of wastewater in the rhizosphere of wetland plants — the root zone method. Water Sci. Technol. 19, 107–118. Brock, T.C.M., Crum, S.J.H., Van Wijngaarden, R., Budde, B.J., Tijink, J., Zuppelli, A., Leeuwangh, P., 1992. Fate and effects of the insecticide Dursban 4e in indoor elodea-dominated and macrophyte-fee freshwater model ecosystem. I. Fate and primary effects of the active ingredient chlorpyrifos. Arch. Environ. Contam. Toxicol. 2, 69–84. Brown, C.D., van Beinum, W., 2009. Pesticide transport via sub-surface drains in Europe. Environ. Pollut. 157, 3314–3324. Budd, R., O'Geen, A., Goh, K.S., Bondarenko, S., Gan, J., 2009. Efficacy of constructed wetlands in pesticide removal from tailwaters in the Central Valley, California. Environ. Sci. Technol. 43, 2925–2930. Budd, R., O'Geen, A., Goh, K.S., Bondarenko, S., Gan, J., 2011. Removal mechanisms and fate of insecticides in constructed wetlands. Chemosphere 83, 1581–1587. Carter, A.D., 2000. Herbicide movement in soils: principles, pathways and processes. Weed Res. 40, 113–122. Cheng, S., Vidakovic-Cifrek, Ž., Grosse, W., Karrenbrock, F., 2002. Xenobiotics removal from polluted water by a multifunctional constructed wetland. Chemosphere 48, 415–418. Chow, T.L., Rees, H.W., Daigle, J.L., 1999. Effectiveness of terraces/grassed waterway systems for soil and water conservation: a filed evaluation. J. Soil Water Conserv. 54, 577–583. Cooper, P.F., 2005. The performance of vertical flow constructed wetland systems with special reference to the significance of oxygen transfer and hydraulic loading rates. Water Sci. Technol. 51, 81–90. Crum, S.J.H., van Kammen-Polma, A.M.M., Leistra, M., 1999. Sorption of nine pesticides to three aquatic macrophytes. Arch. Environ. Contam. Toxicol. 37, 310–316. Dabney, S.M., Moore, M.T., Locke, M.A., 2006. Integrated management of in-field edge-offield and after-field buffers. J. Am. Water Resour. Assoc. 42, 15–24. Dabrowski, J.M., Peall, S.K.C., Reinecke, A.J., Liess, M., Schulz, R., 2002. Runoff-related pesticide input into the Lourens River, South Africa: basic data for exposure assessment and risk mitigation at the catchment scale. Water Air Soil Pollut. 135, 265–283. Dabrowski, J.M., Bennett, E.R., Bollen, A., Schulz, R., 2006. Mitigation of azinophos-methyl in a vegetated stream: comparison of runoff- and spray-drift. Chemosphere 62, 204–212.

Dordio, A., Carvalho, A.J.P., 2013. Constructed wetlands with light expanded clay aggregates for agricultural wastewater treatment. Sci. Total Environ. 463–464, 454–461. Dosskey, M.G., 2001. Toward quantifying water pollution abatement in response to installing buffers on crop land. Environ. Manag. 28, 577–598. Elsaesser, D., Buseth Blankenberg, A.-G., Geist, A., Mæhlum, T., Schulz, R., 2011. Assessing the influence of vegetation on reduction of pesticide concentration in experimental surface flow constructed wetlands: application of the toxic units. Ecol. Eng. 37, 955–962. Elsaesser, D., Stang, C., Bakanov, N., Schulz, R., 2013. The Landau Stream Mesocosm Facility: pesticide mitigation in vegetated flow-through streams. Bull. Environ. Contam. Toxicol. 90, 640–645. Elsayed, O.F., Maillard, E., Vuilleumier, S., Nijenhuis, I., Richnow, H.H., Imfeld, G., 2014. Using compound-specific isotope analysis to assess the degradation of chloroacetanilide herbicides in lab-scale wetlands. Chemosphere 99, 89–95. Felsot, A.S., Unsworth, J.B., Linders, J.B.H.J., Roberts, G., Rautman, D., Harris, C., Carazo, E., 2010. Agrochemical spray drift; assessment and mitigation — a review. J. Environ. Sci. Health B 46, 1–23. Feurtel-Mazel, A., Grollier, T., Grouselle, M., Ribeyre, F., Boudou, A., 1996. Experimental study of bioaccumulation and effects of the herbicide isoproturon on freshwater rooted macrophytes (Elodea densa and Ludwigia natans). Chemosphere 32, 1499–1512. Fiener, P., Auerswald, K., 2003. Effectiveness of grassed waterways in reducing runoff and sediment delivery from agricultural watersheds. J. Environ. Qual. 32, 929–936. Flury, M., 1996. Experimental evidence of transport of pesticides through field soils — a review. J. Environ. Qual. 25, 25–45. FOCUS, 2007. Landscape and mitigation factor. Report of the FOCUS Working Group on Landscape and Mitigation Factors in Ecological Risk Assessment. EC Document Reference SANCO/10422/2005 V.2.0. Aquatic Risk Assessment. Extended Summary and Recommendations vol. 1, pp. 1–169. Friesen-Pankratz, B., Doebel, C., Farenhorst, A., Goldsborough, L.G., 2003. Interactions between algae (Selenastrum capricornutum) and pesticides: implications for managing constructed wetlands for pesticide removal. J. Environ. Sci. Health B 38, 147–155. George, D., Stearman, G.K., Carlson, K., Lansford, S., 2003. Simazine and metolachlor removal by subsurface flow constructed wetland. Water Environ. Res. 75, 101–112. Gill, Y., Sinfort, C., 2005. Emission of pesticides to the air during sprayer application: a bibliographic review. Atmos. Environ. 39, 5183–5193. Gill, S.L., Spurlock, F.C., Goh, K.S., Ganapathy, C., 2008. Vegetated ditches as a management practice in irrigated alfalfa. Environ. Monit. Assess. 144, 261–267. Gilliam, J.W., 1994. Riparian wetlands and water quality. J. Environ. Qual. 23, 896–900. Gregoire, C., Elsaesser, D., Huguenot, D., Lange, J., Lebeau, T., Merli, A., Mose, R., Passeport, E., Payraudeau, S., Schütz, T., Schulz, R., Tapia-Padilla, G., Tournebize, J., Trevisan, M., Wanko, A., 2009. Mitigation of agricultural nonpoint-source pesticide pollution in artificial wetland ecosystems. Environ. Chem. Lett. 7, 205–231. Grover, R., Shewchuk, S.R., Cessna, A.J., Smith, A.E., Hunter, J.H., 1985. Fate of 2,4-D isooctyl ester after application to a wheat field. J. Environ. Qual. 14, 203–210. Guo, W., Zhang, H., Huo, S., 2014. Organochlorine pesticides in aquatic hydrophyte tissues and surrounding sediments in Baiyangdian wetland, China. Ecol. Eng. 67, 150–155. Haarstad, K., Braskerud, B.C., 2005. Pesticide retention in the watershed and in a small constructed wetland treating diffuse pollution. Water Sci. Technol. 51, 143–150. Hinman, M.L., Klaine, S.J., 1992. Uptake and translocation of selected organic pesticides by rooted aquatic plant Hydrilla verticillata Royle. Environ. Sci. Technol. 26, 609–613. Imfeld, G., Braeckevelt, M., Kuschk, P., Richnow, H.H., 2009. Monitoring and assessing processes of organic chemicals removal in constructed wetlands. Chemosphere 74, 349–362. Imfeld, G., Lefranq, M., Maillard, E., Payraudeau, S., 2013. Transport and attenuation of dissolved glyphosate and AMPA in a stormwater wetland. Chemosphere 90, 1333–1339. Jargentz, S., Mugni, H., Bonetto, C., Schulz, R., 2005. Assessment of insecticide contamination in runoff and stream water of small agricultural streams in the main soybean area of Argentina. Chemosphere 61, 817–826. Kadlec, R.H., Hey, D.L., 1994. Constructed wetlands for river water quality improvement. Water Sci. Technol. 29, 159–168. Kadlec, R.H., Wallace, S.D., 2009. Treatment Wetlands, 2nd edition. CRC Press, Boca Raton, Florida. Karen, D.J., Joab, B.M., Wallin, J.M., Johnson, K.A., 1998. Partitioning of chlorpyrifos between water and an aquatic macrophyte (Elodea densa). Chemosphere 37, 1579–1586. Karpuczu, M.E., Sedlak, D.L., Stringfellow, W.T., 2013. Biotransformation of chlorpyrifos in riparian wetlands in agricultural watershed: implications for wetland management. J. Hazard. Mater. 244–245, 111–120. Kegley, S.E., Hill, B.R., Orme, S., Choi, A.H., 2014. PAN Pesticide Action Network, North America (Oakland, CA, 2014). , (http://www.pesticideinfo.org). Kenaga, F.E., 1980. Predicted bioconcentration factors and soils sorption coefficients of pesticides and other chemicals. Ecotoxicol. Environ. Saf. 4, 26–38. Kidmose, J., Dahl, M., Engesgaard, P., Nilsson, B., Christensen, B.S.B., Andersen, S., Hoffmann, C.C., 2010. Experimental and numerical study of the relation between flow paths and fate of a pesticide in a riparian wetland. J. Hydrol. 386, 67–79. Kimbrough, R.A., Litke, D.W., 1996. Pesticides in streams draining agricultural and urban areas in Colorado. Environ. Sci. Technol. 30, 908–918. Kladivko, E.J., Brown, L.C., Baker, J.L., 2001. Pesticide transport to subsurface tile drains in humid regions of North America. Crit. Rev. Environ. Sci. Technol. 31, 1–62. Kohler, E.A., Poole, V.L., Reicher, Z.J., Turco, R.F., 2004. Nutrient, metal, and pesticide removal during storm and nonstorm events by a constructed wetland on an urban golf course. Ecol. Eng. 23, 285–298. Kreuger, J., 1998. Pesticides in stream water within an agricultural catchment in southern Sweden, 1990–1996. Sci. Total Environ. 216, 227–251. Krutz, L.J., Senseman, S.A., Zablotowicz, R.M., Matocha, M.A., 2005. Reducing herbicide runoff from agricultural fields with vegetative filter strips: a review. Weed Sci. 53, 353–367.

J. Vymazal, T. Březinová / Environment International 75 (2015) 11–20 Lacas, J.G., Voltz, M., Gouy, V., Carluer, N., Grill, J.J., 2005. Using grassed strips to limit pesticide transfer to surface water: a review. Agron. Sustain. Dev. 25, 253–266. Larney, F.J., Cessna, A.J., Bullock, M.S., 1999. Herbicide transport on wind-eroded sediment. J. Environ. Qual. 28, 1412–1421. Larsbo, M., Löfstrand, E., van Alpen de Veer, D., Ulén, B., 2013. Pesticide leaching from two Swedish topsoils of contrasting texture amended with biochar. J. Contam. Hydrol. 147, 73–81. Leu, C., Singer, H., Stamm, C., Muller, S.R., Schwarzenbach, R.P., 2004. Variability of herbicide losses from 13 fields to surface water within a small catchment after a controlled herbicide application. Environ. Sci. Technol. 38, 3835–3841. Lin, T., Wen, Y., Jiang, L., Li, J., Yang, S., Zhou, Q., 2008. Study of atrazine degradation in subsurface flow constructed wetland under different salinity. Chemosphere 72, 122–128. Liu, X., Zhang, X., Zhang, M., 2008. Major factors influencing the efficacy of vegetated buffer strips on sediment trapping: a review and analysis. J. Environ. Qual. 37, 1667–1674. Lizotte Jr., R.E., Shields Jr., F.D., Knight, S.S., Bryant, C.T., 2009. Efficiency of a modified backwater wetland trapping a pesticide mixture. Ecohydrology 2, 287–293. Lizotte Jr., R.E., Shields Jr., F.D., Murdock, J.N., Kröger, R., Knight, S.S., 2012. Mitigation agrichemicals from an artificial runoff event using a managed riverine wetland. Sci. Total Environ. 427–428, 373–381. Locke, M.A., Weaver, M.A., Zablotowicz, R.M., Steinriede, R.W., Bryson, C.T., Cullum, R.F., 2011. Constructed wetlands as a component of the agricultural landscape: mitigation of herbicides in simulated runoff from upland drainage areas. Chemosphere 83, 1532–1538. Long, R., Gan, Y., Nett, M., 2005. Pesticide choice: best management practice (BMP for protecting surface water quality in agriculture. Publication 8161. University of California, Agriculture and Natural Resources Communication Services, Oakland, California. Lowrance, R.R., 1992. Groundwater nitrate and denitrification in a coastal plain riparian forest. J. Environ. Qual. 21, 401–405. Mahabali, S., Spanoghe, P., 2014. Mitigation of two insecticides by wetland plants: feasibility study for the treatment of agricultural runoff in Suriname (South America). Water Air Soil Pollut. 225, 1771. Maillard, E., Imfeld, G., 2014. Pesticide mass budget in a stormwater wetland. Environ. Sci. Technol. 48, 8603–8611. Maillard, E., Payraudeau, S., Faivre, E., Grégorie, C., Gangloff, S., Imfeld, G., 2011. Removal of pesticide mixtures in a stormwater wetland controlling runoff from a vineyard catchment. Sci. Total Environ. 409, 2317–2324. Matamoros, V., Puigagut, J., García, J., Bayona, J.M., 2007. Behavior of selected priority organic pollutants in horizontal subsurface flow constructed wetlands: a preliminary screening. Chemosphere 69, 1374–1380. McKinlay, R.G., Kasperek, K., 1999. Observations on decontamination of herbicidepolluted water by marsh plant system. Water Res. 33, 505–511. Messing, P., Farenhorst, A., Waite, D., Sproull, J., 2013. Influence of usage and chemicalphysical properties on the atmospheric transport and deposition of pesticides to agricultural regions of Manitoba, Canada. Chemosphere 90, 1997–2003. Milam, C.D., Bouldin, J.L., Farris, J.L., Schulz, R., Moore, M.T., Bennett, E.R., Cooper, C.M., Smith Jr., S., 2004. Evaluating acute toxicity of methyl parathion application in constructed wetland mesocosms. Environ. Toxicol. 19, 471–479. Mitsch, W.J., Gosselink, J.G., 2000. Wetlands, 3rd edition. John Wiley & Sons, New York. Moore, M.T., Rodgers Jr., J.H., Cooper, C.M., Smith Jr., S., 2000. Constructed wetlands for mitigation of atrazine-associated agricultural runoff. Environ. Pollut. 110, 393–399. Moore, M.T., Bennett, E.R., Cooper, C.M., Smith Jr., S., Shields Jr., F.D., Milam, C.D., Farris, J.L., 2001a. Transport and fate of atrazine and lambda-cyhalothrin in an agricultural drainage ditch in the Mississippi Delta, USA. Agric. Ecosyst. Environ. 87, 309–314. Moore, M.T., Rodgers Jr., J.H., Smith Jr., S., Cooper, C.M., 2001b. Mitigation of metolachlorassociated agricultural runoff using constructed wetlands in Mississippi, USA. Agric. Ecosyst. Environ. 84, 169–176. Moore, M.T., Schulz, R., Cooper, C.M., Smith Jr., S., Rodgers Jr., J.H., 2002. Mitigation of chlorpyrifos runoff using constructed wetlands. Chemosphere 46, 827–835. Moore, M.T., Bennett, E.R., Cooper, C.M., Smith Jr., S., Farris, J.L., Drouillard, K.G., Schulz, R., 2006. Influence of vegetation in mitigation of methyl parathion runoff. Environ. Pollut. 142, 288–294. Moore, M.T., Cooper, C.M., Smith Jr., S., Cullum, R.F., Knight, S.S., Locke, M.A., Bennett, E.R., 2007. Diazinon mitigation in constructed wetlands: influence of vegetation. Water Air Soil Pollut. 184, 313–321. Moore, M.T., Denton, D.L., Cooper, C.M., Wrysinski, J., Miller, J.L., Reece, K., Crane, D., Robins, P., 2008. Mitigation assessment of vegetated drainage ditches for collecting irrigation runoff in California. J. Environ. Qual. 37, 486–493. Moore, M.T., Cooper, C.M., Smith Jr., S., Cullum, R.F., Knight, S.S., Locke, M.A., Bennett, E.R., 2009. Mitigation of two pyrethroid insecticides in a Mississippi Delta constructed wetland. Environ. Pollut. 157, 250–256. Moore, M.T., Tyler, H.L., Locke, M.A., 2013. Aqueous pesticide mitigation efficiency of Typha latifolia (L.), Leersdia oryzoides (L.) Sw., and Sparganium americanum Nutt. Chemosphere 92, 1307–1313. Oerke, E.C., 2006. Crop losses to pests. J. Agric. Sci. 144, 31–43. Oerke, E.C., Dehne, H.W., Schonbeck, F., Weber, A., 1994. Crop Production and Crop Protection: Estimated Losses in Major Food and Cash Crops. Elsevier, Amsterdam. Ohliger, R., Schulz, R., 2010. Water body and riparian buffer strip characteristics in a vineyard area to support aquatic pesticide exposure assessment. Sci. Total Environ. 408, 5405–5413. Olette, R., Couderchet, M., Biagianti, S., Eullaffroy, P., 2008. Toxicity and removal of pesticides by selected aquatic plants. Chemosphere 70, 1414–1421. Olivier, D.P., Kookana, R.S., Anderson, J.S., Cox, J.W., Waller, N., Smith, L.H., 2012. Off-site transport of pesticides in dissolved and particulate forms from two land uses in the Mt. Lofty Ranges, South Australia. Agric. Water Manag. 106, 78–85.

19

Otto, S., Vianello, M., Infantino, A., Zanin, G., Di Guardo, A., 2008. Effect of a full-grown vegetative filter strip on herbicide runoff: maintaining of filter capacity over time. Chemosphere 71, 74–82. Page, D., Dillon, P., Mueller, J., Bartkow, M., 2010. Quantification of herbicide removal in a constructed wetland using passive samplers and composite water quality monitoring. Chemosphere 81, 394–399. Passeport, E., Tournebize, J., Chaumont, C., Guenne, A., Coquet, Y., 2013. Pesticide contamination interception strategy and removal efficiency in forest buffer and artificial wetland in a tile-drainage agricultural watershed. Chemosphere 91, 1289–1296. Pavel, E.W., Lopez, A.R., Berry, D., Smith, E.P., Reneau Jr., R.B., Mostaghimi, S., 1999. Anaerobic degradation of dicamba and metribuzin in riparian wetland soils. Water Res. 33, 87–94. Peterjohn, W.T., Correll, D.L., 1984. Nutrient dynamics in an agricultural watershed: observations on the role of a riparian forest. Ecology 65, 1466–1475. Poissant, L., Beauvais, C., Lafrance, P., Deblois, C., 2008. Pesticides in fluvial wetlands catchments under intensive agricultural activities. Sci. Total Environ. 404, 182–195. Probst, M., Berenzen, N., Lentzen-Godding, A., Schulz, R., Liess, M., 2005. Linking land use variables and invertebrate taxon richness in small and medium-sized agricultural streams on a landscape level. Ecotoxicol. Environ. Saf. 60, 140–146. Ree, W.O., 1949. Hydraulic characteristics of vegetation for vegetated waterways. Agric. Eng. 30, 184–187. Reed, S.C., Middlebrooks, E.J., Crites, R.W., 1988. Natural Systems for Waste Management and Treatment, 1st ed. McGraw-Hill Book Company, New York. Reichenberger, S., Bach, M., Skitschak, A., Frede, H.-G., 2007. Mitigation strategies to reduce pesticide inputs into ground- and surface water and their effectiveness: a review. Sci. Total Environ. 384, 1–35. Riise, G., Lundekvam, H., Wu, Q.L., Haugen, L.E., Mulder, J., 2004. Loss of pesticides from agricultural fields in SE Norway — runoff through surface and drainage water. Environ. Geochem. Health 26, 269–276. Rodgers Jr., J.H., Dunn, A., 1992. Developing design guidelines for constructed wetlands to remove pesticides from agricultural runoff. Ecol. Eng. 1, 83–95. Roessink, I., Arts, G.H.P., Belgers, J.D.M., Bransen, F., Maund, S.J., Brock, T.C.M., 2005. Effect of lambda-cyhalothrin in two ditch microcosms systems of different trophic status. Environ. Toxicol. Chem. 24, 1684–1696. Rogers, M.R., Stringfellow, W.T., 2009. Partitioning of chlorpyrifos to soil and plants in vegetated agricultural drainage ditches. Chemosphere 75, 109–114. Rose, M.T., Sanchez-Bayo, F., Crossan, A.N., Kennedy, I.R., 2006. Pesticide removal from cotton farm tailwater by a pilot-scale ponded wetland. Chemosphere 63, 1849–1858. Rose, M.T., Crossan, A.N., Kennedy, I.R., 2008. The effect of vegetation on pesticide dissipation from ponded treatment wetlands: quantification using a simple model. Chemosphere 72, 999–1005. Runes, H.B., Jenkins, J.J., Moore, J.A., Bottomley, P.J., Wilson, B.D., 2003. Treatment of atrazine in nursery irrigation runoff by a constructed wetland. Water Res. 37, 539–550. Schmitt, T.J., Dosskey, M.G., Hoagland, K.D., 1999. Filter strip performance and processes for different vegetation, widths, and contaminants. J. Environ. Qual. 28, 1479–1489. Schulz, R., 2004. Field studies on exposure, effects, and risk mitigation of aquatic nonpointsource insecticide pollution: a review. J. Environ. Qual. 33, 419–448. Schulz, R., Peall, S.K.C., 2001. Effectiveness of a constructed wetland for retention of nonpoint-source pesticide pollution in the Lourens River Catchment, South Africa. Environ. Sci. Technol. 35, 422–426. Schulz, R., Peall, S.K.C., Dabrowski, J.M., Reinecke, A.J., 2001a. Spray deposition of two insecticides into surface waters in a South African orchard area. J. Environ. Qual. 30, 814–822. Schulz, R., Peall, S.K.C., Hugo, C., Krause, V., 2001b. Concentration, load and toxicity of spraydrift-borne azinphos-methyl at the inlet and outlet of a constructed wetland. Ecol. Eng. 18, 239–245. Schulz, R., Hahn, C., Bennett, E.R., Dabrowski, J.M., Thiere, G., Peal, S.K.C., 2003. Fate end effects of azinphos-methyl in a flow-through wetland in South Africa. Environ. Sci. Technol. 37, 2139–2144. Sherrard, R.M., Bearr, J.S., Murray-Gulde, C.L., Rodgers Jr., J.H., Shah, Y.T., 2004. Feasibility of constructed wetlands for removing chlorothalonil and chlorpyrifos from aqueous mixtures. Environ. Pollut. 127, 385–394. Stang, C., Wieczorek, M.V., Noss, C., Lorke, A., Scherr, F., Goerlitz, G., Schulz, R., 2014. Role of submerged vegetation in the retention processes of three plant protection products in flow-through stream mesocosms. Chemosphere 107, 13–22. Stangroom, S.J., Collins, C.D., Lester, J.N., 2000. Abiotic behavior of organic micropollutants in soils and the aquatic environment. A review: II. Transformation. Environ. Technol. 21, 865–882. Stearman, G.K., George, D.B., Carlson, K., Lansford, S., 2003. Pesticide removal from container nursery runoff in constructed wetland cells. J. Environ. Qual. 32, 1548–1556. Steffens, K., Larsbo, M., Moeys, J., Jarvis, N., Lewan, E., 2013. Predicting pesticide leaching under climate change: importance of model structure and parameter uncertainty. Agric. Ecosyst. Environ. 172, 24–34. Stehle, S., Elsaesser, D., Gregoire, C., Imfeld, G., Niehaus, E., Passeport, E., Payrsudeau, S., Schäfer, R.B., Tournebize, J., Schulz, R., 2011. Pesticide risk mitigation by vegetated treatment systems: a meta-analysis. J. Environ. Qual. 40, 1068–1080. Stephenson, G.R., Ferris, I.G., Holland, P.T., Nordberg, M., 2006. Glossary of terms relating to pesticides (IUPAC recommendations 2006). Pure Appl. Chem. 78, 2075–2154. Stomp, A.M., Han, K.H., Wilbert, S., Gordon, M.P., Cunningham, S.D., 1994. Genetic strategies for enhancing phytoremediation. Ann. N. Y. Acad. Sci. 721, 481–491. Tang, X., Zhu, B., Katou, H., 2012. A review of rapid transport of pesticides from sloping farmland to surface waters: processes and mitigation strategies. J. Environ. Sci. 24, 351–361. Tiktak, A., de Nie, D.S., Piñeros Garcet, J., Jones, A., Vanclooster, M., 2004. Assessment of the pesticide leaching risk at the pan-European level: the EuroPEARL approach. J. Hydrol. 289, 222–238.

20

J. Vymazal, T. Březinová / Environment International 75 (2015) 11–20

Tournebize, J., Vincent, B., Chaumont, C., Gramaglia, C., Margoum, C., Molle, P., Carluer, N., Gril, J.J., 2011. Ecological services of artificial wetland for pesticide mitigation. Sociotechnical adaptation for watershed management through TRUSTEA project feedback. Procedia Environ. Sci. 9, 183–190. Tournebize, J., Passeport, E., Chaumont, C., Fesnau, C., Guenne, A., Vincent, B., 2013. Pesticide de-contamination of surface waters as a wetland ecosystem service in agricultural landscape. Ecol. Eng. 56, 51–59. Ucar, T., Hall, F.R., 2001. Windbreaks as a pesticide drift mitigation strategy: a review. Pest Manag. Sci. 57, 663–675. University of Hertfordshire, 2006–2013. The Pesticide Properties DataBase (PPDB) Developed by the Agriculture & Environment Research Unit (AERU). University of Hertfordshire, (http://sitem.herts.ac.uk/aeru/projects/ppdb/index.htm). Vallée, R., Dousset, S., Billet, D., Benoit, M., 2014. Sorption of selected pesticides on soils, sediment and straw from a constructed agricultural drainage ditch or pond. Environ. Sci. Pollut. Res. 21, 4895–4905. Vink, J.P.M., Van der Zee, S.E.A.T.M., 1997. Pesticide biotransformation in surface waters: multivariate analyses of environmental factors at field sites. Water Res. 31, 2858–2868. Vymazal, J., 2009. The use constructed wetlands with horizontal sub-surface flow for various types of wastewater. Ecol. Eng. 35, 1–17. Vymazal, J., 2011. Constructed wetlands for wastewater treatment: five decades of experience. Environ. Sci. Technol. 45, 61–69. Vymazal, J., 2013. The use of hybrid constructed wetlands for wastewater treatment with special attention to nitrogen removal: a review of a recent development. Water Res. 47, 4795–4811.

Vymazal, J., Kröpfelová, L., 2008. Wastewater treatment in constructed wetlands with horizontal subsurface flow. Springer, Dordrecht, The Netherlands. Wauchope, R.D., 1978. The pesticide content of surface water draining from agricultural fields — a review. J. Environ. Qual. 7, 459–472. Wolverton, B.C., 1975. Aquatic plants for removal of mevinphos from the aquatic environment. NASA Tech Mem TM-X-72720. Wolverton, B.C., Harrison, D.D., 1974. Aquatic plants for removal of mevinphos from the aquatic environment. J. Miss. Acad. Sci. 19, 84–88. Wu, Q.L., Riise, G., Lundekvam, H., Mulder, J., Haugen, L.E., 2004. Influences of suspended particles on the runoff of pesticides from an agricultural field at Askim, SE-Norway. Environ. Geochem. Health 26, 295–302. Yang, H., Dick, W.A., McCoy, E.L., Phelan, P.L., Grewal, P.S., 2013. Field evaluation of a new biphasic rain garden for stormwater flow management and pollutant removal. Ecol. Eng. 54, 22–31. Zablotowicz, R.M., Locke, M.A., Krutz, L.J., Lerch, R.N., Lizotte, R.E., Knight, S.S., Gordon, R.E., Steinriede, R.W., 2006. Influence of watershed system management on herbicide concentrations in Mississippi Delta oxbow lakes. Sci. Total Environ. 370, 552–560. Zhang, X., Zhang, M., 2011. Modeling effectiveness of agricultural BMPs to reduce sediment load and organophosphate pesticides in surface runoff. Sci. Total Environ. 409, 1949–1958.

The use of constructed wetlands for removal of pesticides from agricultural runoff and drainage: a review.

Pesticides are used in modern agriculture to increase crop yields, but they may pose a serious threat to aquatic ecosystems. Pesticides may enter wate...
614KB Sizes 7 Downloads 7 Views