Open access peer-reviewed chapter

Biological Remediation of Phenoxy Herbicide-Contaminated Environments

Written By

Magdalena Urbaniak and Elżbieta Mierzejewska

Submitted: 30 April 2019 Reviewed: 27 June 2019 Published: 07 August 2019

DOI: 10.5772/intechopen.88256

From the Edited Volume

Environmental Chemistry and Recent Pollution Control Approaches

Edited by Hugo Saldarriaga-Noreña, Mario Alfonso Murillo-Tovar, Robina Farooq, Rajendra Dongre and Sara Riaz

Chapter metrics overview

1,287 Chapter Downloads

View Full Metrics

Abstract

Phenoxy herbicides such as 2,4-dichlorophenoxyacetic acid (2,4-D) and 2-methyl-4-chlorophenoxyacetic acid (MCPA) are widely used in agriculture to control broadleaf weeds. Although their application has helped to increase the yield and value of crops, they are also recognized as a source of emerging environmental contamination. Their extensive use may promote contamination of soil, surface, and groundwater and lead to increased inhibition of plant development and soil toxicity. Hence, there is an urgent need to identify nature-based methods based on appropriate biological remediation techniques, such as bio-, phyto-, and rhizoremediation, that enable the effective elimination of phenoxy herbicides from the environment. Bioremediation typically harnesses microorganisms and their ability to utilize recalcitrant contaminants in complete degradation processes, while phytoremediation is a cost-effective, environmentally friendly strategy that uses plants to transform or mineralize xenobiotics to less or nontoxic compounds. Rhizoremediation (microbe-assisted phytoremediation), in turn, is based on the interactions between plant roots, root exudates enriched in plant secondary metabolites, soil, and microorganisms. Based on the above, this chapter presents current knowledge on the properties of phenoxy herbicides, as well as the concentrations detected in the environment, their toxicity, and the biological remediation techniques used for safe removal of the compounds of interest from the environment.

Keywords

  • 2
  • 4-D
  • MCPA
  • bioremediation
  • phytoremediation
  • rhizoremediation
  • toxicity
  • degradative genes

1. Phenoxy herbicides: general information

2,4-Dichlorophenoxyacetic acid (2,4-D) and 2-methyl-4-chlorophenoxyacetic acid (MCPA) are the most commonly used phenoxy acid herbicides in agriculture, and 2,4-D is now the fifth most extensively used active ingredient (a.i.) in the US agricultural and home/garden market sector [1]. In addition, in 2016, 6.5 mln kg of herbicides based on phenoxy-phytohormones (2,4-D and MCPA) were sold in in the EU, including ~2 mln kg sold in Poland [2].

Phenoxy herbicides are typically used to protect wheat, one of the most extensively cultivated crops, because they selectively control the growth of dicotyledonous weeds [3]. They are applied as post emergence agents and taken up by broad-leaved plants. 2,4-D has also been extensively used as an anti-stalling agent for the postharvest fresh fruit industry [4].

These herbicides are based on ring-like structures and have at least one chlorine atom attached to the ring at different positions [5]. Their action is similar to that of phytohormones (auxins) insofar that they can redirect the regulation of plant growth/physiological processes, resulting in nutrition deficiency and subsequent plant death [6].

They are typically released to the environment in the form of commercial products containing phenoxy acids salts or esters; however, they immediately hydrolyze to their corresponding anionic or neutral form [7]. The dosage of phenoxy herbicides lies in the range of 0.8–1.8 kg of a.i. per ha. Their transport through the environment is governed by soil and climate factors (e.g., distribution of soil particles, soil permeability, soil depth, soil pH, soil organic matter content, land slope) [8], and their retention and translocation in the soil profile also depend on their chemical and physical properties, which are described by several parameters (Table 1), particularly pKa (acid dissociation constant), logP (octanol-water partition coefficient), and Koc (organic carbon distribution coefficient). The degree of adsorption and desorption depends on time and the physicochemical properties of soil; however, 2,4-D and MCPA are rather poorly adsorbed on the soil particles in comparison to their derivatives, which have different sorption characteristics [7].

Table 1.

Physical and chemical properties of 2,4-D and MCPA.

Although phenoxy herbicides are described as nonpersistent and weakly adsorbed (Koc < 50) in soil, they can be transported with runoff and in soil profile and reach terrestrial and water ecosystems (surface and groundwater). Figure 1 summarizes the transport and transfer processes of phenoxy herbicides in the environment. After they are applied to land, they are spread through several processes, including sorption/desorption, leaching, runoff, and plant uptake [8]. Phenoxy herbicide molecules are negatively charged and are therefore highly mobile at neutral pH. In groundwater, they are nonvolatile and persistent to hydrolysis, but they can be degraded biologically under both aerobic and anaerobic conditions. These herbicides demonstrate significantly greater persistence in temperate climates characterized by low winter temperatures and, in many regions, by depleted soil organic carbon content and acidic pH [7].

Figure 1.

Transport and transfer processes of phenoxy herbicides in the environment.

Advertisement

2. Phenoxy herbicides: potential contaminants of soil and water environments

Extensive use of phenoxy herbicides can threaten surface and groundwater ecosystems by promoting the contamination of soil matrices. The International Agency for Research on Cancer classifies phenoxy acids as “possibly carcinogenic to humans.” Gupta et al. [11] report that 0.5 kg/ha is the optimal concentration of 2,4-D which avoids contamination of environmental matrices, with the effect of higher concentrations of 2,4-D on the environment being dependent on irrigation treatment. Hence, little is known of the distribution of phenoxy herbicides in the environment. Data from several sources have identified increased levels of 2,4-D and MCPA in the soil, ground-, surface, and drinking water (Table 2); for example, Ignatowicz and Struk-Sokołowska [12] note that the concentration of phenoxy herbicides in the Narew River (Poland) fluctuated seasonally from 0 to even 150 μg/L. The concentration of MCPA in the Parramatta River (Sydney Estuary, Australia) was 0.061 μg/L; however, its presence in river water was caused by increased runoff of storm water [13]. MCPA concentration has been found to be as high as 42.40 μg/L in the Rhone River (France) [14] and to be as little as 0.58 μg/L in Brejo of Cagarrão Stream (Portugal) [15]. The 2,4-D concentration has been found to vary from 1.678 μg/L in the water of McGregor Creek (Canada) [16] to 329.42 μg/L in water from a rice field [17]. By contrast, the maximum permissible concentration of pesticide residues in drinking water is 0.50 μg/L (Directive E98/83/EC). The data presented in Table 2 and described above indicate that phenoxy herbicides should be considered as emerging contaminant especially in water resources.

CompoundConcentrationEnvironmental matricesSource
2,4-D1.678 μg/LWater from McGregor Creek (Canada)[16]
2,4-D103.99–329.42 μg/LWater from rice field (Malaysia)[14]
2,4-D0.0052 mg/kgSoil from cereal plantations (Poland)[15]
2,4-D0.513 μg/LLebo drain[16]
MCPA, MCPP, 2,4-D0–150 μg/LWater from Narew River (Poland)[12]
MCPA0.0046 mg/kgSoil from cereal plantations (Poland)[15]
MCPA0.08–42.40 μg/LWater from Rhône River Delta (France)[18]
MCPA0.58 μg/LWater from Brejo of Cagarrão Stream (Portugal)[17]
MCPA0.061 μg/LWater from Parramatta River—Sydney Estuary (Australia)[13]
MCPA82.75–354.28 μg/LWater from rice field (Malaysia)[14]
MCPA0.002–0.010 mg/kgSoil from potato plantation (Poland)[19]

Table 2.

Concentration of phenoxy acids observed in various environments.

Despite the diversified levels of phenoxy herbicides noted in worldwide environments (Table 2), it has to be underlined that these compounds can exert serious toxic effects on the sustainability of ecosystems, even at lower concentrations (e.g., 0.275 μg/L) (Table 3). According to recent research, predicted no effect concentration (PNEC) for aquatic organisms is 500 μg/L for 2,4-D and 0.022 μg/L for MCPA [20]; however, PNEC has not yet been determined for terrestrial organisms.

Dose of phenoxy herbicideExposure timeTest organismEffect on organismSource
220.04 μg/L 2,4-D2 daysSinapis arvensis (wild mustard)Inhibition of root and hypocotyl elongation[22]
5.06 mg/L 2,4-D72 hoursPisum sativum (pea)Severe disturbances in mesophyll cell structure and proliferation of vascular tissue in young leaves[23]
10, 100, 500, 1000 μg/L MCPA7 daysHydrilla verticillata (waterthyme)Disturbance of growth, anatomy, and physiology[36]
IC50 1353.80 mg/L 2,4-D96 hoursAnkistrodesmus falcatus (green microalgae)External morphological alterations[24]
IC50 71.20 mg/L 2,4-D96 hoursMicrocystis aeruginosa (toxigenic cyanobacteria)Stimulation of the production of cyanotoxins
LC50 66 mg/L 2,4-D96 hoursCyprinus carpio (common carp)Behavioral changes[26]
LC50 9.06 and 7.76 mg/L 2,4-D96 and 168 hoursRhinella arenarum (species of toad)Reduced body size, delayed development, microcephaly, agenesis of gills, abnormal cellular proliferation processes[27]
10–500 mg/L 2,4-D1 hourHuman erythrocytesHemolysis[29]
0.275, 2.75, and 27.5 μg/L 2,4-D and MCPA30 minutesHepatic cells of Metynnis roosevelti (species of serrasalmid fish)Damage of cellular metabolism and homeostasis; increased oxidative stress[21]

Table 3.

The results of toxicological tests and effects of 2,4-D and MCPA on selected organisms.

Because the mode of action of phenoxy herbicides mimics that of plant growth hormones, their application causes disturbances among a range of physiological processes [21]. 2,4-D inhibits root/hypocotyl elongation in Sinapis arvensis (wild mustard) and disrupts mesophyll cell structure in Pisum sativum (pea) [22, 23]. There is increasing concern that 2,4-D has negative influence on water ecosystems, leading to cellular deformation of green algae, such as Ankistrodesmus falcatus [24]; malformations and behavioral changes to various fish, including Cyprinus carpio (common carp) and Danio rerio (zebrafish) [25, 26]; abnormal cellular proliferation in amphibians such as Rhinella arenarum (species of toad) [27]; and the development of nonviable embryos in invertebrates, such as Biomphalaria glabrata (species of freshwater snail) [28]. Sarikaya and Yilmaz [26] report that 2,4-D (66,000 μg/L) causes internal hemorrhage and behavioral changes in C. carpio.

Among animals, phenoxy herbicide application results in the inhibition of crucial enzymes in cell metabolism, including mitochondrial enzymes and those associated with DNA synthesis (Table 3). 2,4-D has also been found to induce erythrocyte lysis under laboratory conditions [29]. It is interesting to note that the intermediates formed during the degradation processes of 2,4-D, such as 2,4-dichlorophenol (2,4-DCP) and 3,5-dichlorocatechol (3,5-DCC), exhibit a strong ecotoxic effect on various organisms, including N. tabacum cells [30]; however, Taylor et al. [31] report that 2,4-DCP toxicity was found to be less phytotoxic than 2,4-D, under both in vitro and in vivo conditions, while 3,5-DCC exhibits higher toxicity than its parent compound [32].

Several studies have revealed that MCPA also can have a negative impact on the environment: MCPA application caused up to a 56% reduction in dehydrogenase, urease, and phosphatase activities and ergosterol content in soil [32]. In addition, this application leads to increased soil phytotoxicity to Fagopyrum esculentum var. Kora (buckwheat) and promoted stem deformation and leaf discoloration [33]. Mierzejewska et al. [34] note that a commercial product containing MCPA was highly toxic to the monocotyledon Sorghum saccharatum (sorghum) and dicotyledons Lepidium sativum (garden cress) and Sinapis alba (white mustard), inducing nearly 100% root growth inhibition. The authors also note that after 3 weeks of incubation at an ambient temperature, the high initial phytotoxicity was reduced to 3% for L. sativum and 34% for S. alba and that S. saccharatum demonstrated a 12% stimulation of root growth in comparison to uncontaminated control soil. The negative influence of MCPA on L. sativum, S. alba, and S. saccharatum growth was also confirmed by Urbaniak et al. [35]. Similarly to 2,4-D, MCPA causes also negative effects on freshwater organisms such as the freshwater crustaceans Daphnia magna, Thamnocephalus platyurus, and Artemia franciscana and alga Selenastrum capricornutum [36]. Both herbicides were found to induce the action of hepatic enzymes involved in detoxification and lipid peroxidation [21].

These studies emphasize the important role played by ecotoxicological approaches in evaluating the effect of chemical stressors observed in the ecosystem communities. Despite the relatively short half-life (Table 1) of 2,4-D and MCPA, their remnants can be transported and deposited extensively in the environment, and this can present a potential threat to the soil and water ecosystems as well as to human health. Therefore, there is a need to identify nature-based solutions such as bio-, phyto-, and rhizoremediation that can enhance the process of phenoxy herbicide elimination from the environment.

Advertisement

3. Phenoxy herbicides: removal using biological methods

One approach to removing phenoxy herbicides (2,4-D and MCPA) from soil is via degradation by the soil microbiota (biodegradation). This is achieved most effectively by bacteria harboring the appropriate functional genes, which are involved in the phenoxy herbicide degradation pathways (Figure 2). Alternatively, plants can be used to decontaminate sites, a process known as phytoremediation (Figure 2). Another promising approach, rhizoremediation, enhances the removal of such recalcitrant xenobiotics from the environment by exploiting the interactions between selected plants (able to grow under the presence of given xenobiotics such as phenoxy herbicides), root exudates (including plant secondary metabolites, PSMs), and microorganisms (Figure 2). The purpose of this section is to review the literature on established and potential biological methods of phenoxy herbicide removal from environmental matrices.

Figure 2.

The biological processes of phenoxy herbicide biodegradation mediated by soil, rhizospheric, and endophytic bacteria.

3.1 Bioremediation

Bioremediation is a method that uses microbiological processes to degrade or transform contaminants to less toxic or nontoxic forms. Biodegradation of organic contaminants occurs very slowly in bulk soil; therefore biostimulation and bioaugmentation methods are used to enhance the biologically driven removal of toxic compounds from environmental matrices. The effectiveness of biodegradation is dependent on several factors, among them the characteristics of the soil, the bioavailability of the contaminants, and their chemical properties.

An important way of phenoxy herbicide removal from soil is by the use of indigenous soil bacteria harboring desirable catabolic genes. The first step in the phenoxy herbicide biodegradation pathway is initiated by α-ketoglutarate-dependent dioxygenase, an enzyme encoded by tfdA or tfdA-like genes [37] located in the tfdABCDEF gene cluster [38].

In recent decades, increasingly rapid advances in the application of molecular analysis in environmental studies have helped identify the bacterial communities involved in phenoxy herbicide biodegradation (Table 4). The bacteria able to metabolize phenoxy herbicides have been classified into three groups as follows: according to their physiology, employed degrading enzymes, and evolutionary origin [39, 40, 41] (Table 4).

  1. The first group consists of fast-growing copiotrophic bacteria belonging to β- and γ-proteobacteria harboring the tfdA gene (e.g., Cupriavidus necator JMP134, Burkholderia sp. strain RASC, and Rhodoferax sp. strain P230). This first group has been subdivided into four subclasses according to tfdA sequence: tfdA Class I, II, and III [42, 43] and tfdA α [38]. Class I is found in Cupriavidus pinatubonensis; Class II is less widely distributed, being found only in Burkholderia spp.; and Class III is found in Comamonas acidovorans [38]. TfdAα was first identified in Bradyrhizobium sp. [40]. However, tfdAα-encoded protein has been described as α-ketoglutarate-dependent 2,4-D dioxygenase with lower activity than JMP134 dioxygenase.

  2. The second group consists of slow-growing oligotrophic bacteria belonging to α-proteobacteria, phylogenetically closely related to Bradyrhizobium sp. [41], which were isolated from pristine environments. In this group, the phenoxy herbicide degradative gene was also identified and classified as tfdAα. Its gene sequence shows 50–60% similarity to the Group I degrader Cupriavidus necator JMP134.

  3. The third group consists of bacteria belonging to the α-proteobacteria harboring the tfdAα gene, with Sphingomonas being the key member [41]. The wide diversity displayed by tfdA-like genes can partly be attributed to the wide range of bacteria (α-, β-, and γ-proteobacteria) capable of degrading phenoxy acids in the environment. Due to the high degree of homology between strains, the tfdA genes have been selected as biomarkers of the capability of bacteria to metabolize 2,4-D and MCPA [37, 38], and they are frequently used in studies of phenoxy acid biodegradation.

ClassStrainOriginStudied compoundsIdentified functional genesSource
α-ProteobacteriaSphingomonas paucimobilisSoil from Michigan (USA)2,4-D[44]
Sphingomonas agrestis 58–1Soil from Fukuoka Prefecture (Japan)2,4-D, MCPAcadA, cadB[45]
Bradyrhizobium sp.; Sphingomonas sp.Root nodules; pristine environments (Hawaii, central California, USA; southwestern Australia, southwestern Africa; central Chile; northern Saskatchewan, Canada; northwestern Russia); volcanic soil (National Park (Kipuka Keana Bihopa, Hawaii, USA)2,4-DtfdAα, cadA, and cadB[38, 39, 41]
Sphingomonas sp.Sediment from an aquifer in Fladerne Creek (Denmark)MCPAcadA and cadB[46, 47]
β-ProteobacteriaComamonas acidovorans strain MCIHerbicide-contaminated building rubble (Germany)2,4-D and MCPAtfdB and tfdC genes[48]
Variovorax paradoxusSoil from the Dijon INRA experimental station (France)2,4-DtfdA, tfdB, and tfdR[49]
Delftia sp.Polluted river in Buenos Aires (Argentina)2,4-D[50]
Cupriavidus campinensis BJ712,4-D-enriched soils from wheat fields in Beijing exposed for 2,4-D for at least 10 years (China)2,4-DClass I tfdA gene[51]
Achromobacter sp. LZ35Soil in a disused pesticide factory in Suzhou (China)2,4-D and MCPAtfdA and tfdB[52]
Halomonadaceae sp.Alkali Lake site in Oregon (USA) contaminated with 2,4-D production wastes2,4-DtfdA[53]
γ-ProteobacteriaPseudomonas pickettiiAgricultural soil from Michigan (USA)2,4-D[54]
Pseudomonas maltophiliaWheat rhizosphere (laboratory experiment)2,4-D[55]

Table 4.

Bacteria degrading phenoxy herbicides isolated from pristine and contaminated environments.

Much of the current literature on phenoxy herbicide metabolic pathways pays particular attention to the degradation pathway of 2,4-D. One of the most extensively studied 2,4-D degraders is Cupriavidus necator JMP134, known to harbor the 80-kb pJP4 plasmid. pJP4 carries all of the structural and regulatory genes needed to convert phenoxy herbicides to 2-chloromaleylacetic acid [56]. The tfdA fragment is responsible for the conversion of 2,4-D to 2,4-DCP [57]. Subsequently, 2,4-DCP is hydrolyzed to 3,5-dichlorocatechol by 2,4-DCP hydroxylase, which is encoded by tfdB. 3,5-Dichlorocatechol is further degraded via a pathway encoded by tfdCDEF.

Far too little attention has been paid to the metabolism of MCPA. MCPA degradation takes place by the cleavage of an ether linkage, resulting in the formation of the major metabolite, 4-chloro-2-methylphenol (MCP), and acetic acid [47]. This process is preceded by the expression of the tfdA gene. Mierzejewska et al. [34] report that microorganisms demonstrating the presence of tfdAα and tfdA Class III genes in soil contaminated with a commercial product containing MCPA displayed biodegradation potential.

The bacteria carrying cad genes, which encode the non-heme iron oxygenase, also have the potential to degrade both herbicides. The cadRABKC gene cluster was first identified and characterized in strain Bradyrhizobium sp. HW3 which was isolated from pristine environment in Volcanoes National Park, Hawaii [39]. So far, however, there has been little research on the mode of action and exact function of cad genes. According to Kitagawa et al. [38], cadA, cadB, and cadC genes are responsible for multicomponent oxygenase production, whereas cadR is a transcriptional regulator gene, which regulates the transcription of cadABKC in the presence of 2,4-D or 4-chlorophenoxyacetic acid. The cadA gene products show structural and functional differences to the tfdA gene with regard to their substrate preferences. Both the cadA and cadB and the tfdA genes code for aromatic ring hydroxylation dioxygenases (RHDO), which are widely distributed in a number of microorganisms and might be transferred through horizontal gene transfer [38]; CadA- and cadB-encoded proteins are involved in the same initial step of 2,4-D degradation; however, the enzyme subunits have a different mode of action to the ketoglutarate-dependent dioxygenase encoded by tfdA. CadA and cadB were mostly identified in bacteria belonging to Groups I and II of phenoxy herbicide degraders. The products of cadA gene expression are able to initiate the degradation of both MCPA and 2,4-D. Furthermore, the abundance of cadA gene stimulates MCPA degradation [47]. The cadA gene is also essential for 2,4-D conversion in pure cultures of α-proteobacteria [38, 45, 58], and the cadB gene is also thought to play a sole role in the phenoxy acid degradation; however, the exact role of the cad genes remains not fully understood. The two genes share ~50% identity with tfdA, and it has been found that cadA, cadB, and tfdA are expressed simultaneously during MCPA degradation. Interestingly, some bacteria harbor all three cadA, cadB, and tfdA genes, thereby displaying a dual system of degradative genes [47].

The microbial degradation metabolic pathway of phenoxy herbicides has been elaborated in recent years (Figure 3). The first step of this catabolic pathway is initiated by either the tfdA gene which encodes α-ketoglutarate-dependent dioxygenase or cadAB genes which encode subunits of non-heme iron oxygenase [47]. Although these enzymes use different modes of action, both catabolic proteins have been shown to perform the same initial step in phenoxy acid degradation, turning 2,4-D into 2,4-DCP and MCPA into MCP.

Figure 3.

Pathways of microbial degradation of 2,4-D and MCPA proposed by Pieper et al. [59]; in the picture there are indicated functional genes which encode catabolic enzymes as follows: tfdA, α-ketoglutarate-dependent dioxygenase; cadAB, subunits of non-heme iron oxygenase; tfdB, chlorophenol hydroxylase; tfdC, catechol 1,2-dioxygenase; tfdD, dichloromuconate cycloisomerase; tfdE, carboxymethylene butenolidase; tfdF, maleylacetate reductase.

Until recently, there has been little interest in the stereospecific Fe-(II) α-ketoglutarate-dependent dioxygenases which are encoded by rdpA and sdpA genes. These enzymes are described in literature as the ones which can also initiate the first step of the MCPA and 2,4-D degradation pathway. They were identified in Delftia acidovorans, Rhodoferax sp., and Sphingobium sp. Although the proteins encoded by the rdpA and sdpA genes possess the highly conserved amino acid sequence motif of tfdA-encoded proteins, they share only 37% identity with the tfdA genes of C. necator JMP134 [60, 61].

In addition to the soil bacteria, soil microfauna can also profoundly affect the biodegradation of organic contaminants. An important example of this relationship is the activity of earthworms, which move through the soil, causing better aeration and increasing soil moisture. Hence, insofar as their activity can influence the profile of the microorganism communities in the soil, they can indirectly enhance the process of phenoxy herbicide aerobic bacterial degradation [61].

3.2 Phytoremediation

A steadily developing strategy for the in situ treatment of contaminated soils is phytoremediation. It is a cost-effective and environmentally friendly strategy that uses plants to transform or mineralize xenobiotics to less toxic or environmentally neutral compounds [62]. Plants play a crucial role in the development of soil structure and stabilization of fundamental soil ecosystem functions such as water flow [63]. They produce also an array of catabolic enzymes, which operate to protect the host organisms and detoxify xenobiotic compounds [64]. Therefore, phytoremediation not only contributes to the detoxification of the environmental matrices but also has a positive influence on the functioning of the entire ecosystem.

The process of contaminant absorption by plants depends on several factors, including regional climate, soil type, and the nature of the pollutant [65]. The selection of an appropriate plant species and cultivar is critical for effective removal of a given contaminant from soil [66, 67]. This choice of phytoremediation candidate should particularly take into account plant growth rate, high biomass production, capacity for pollutant accumulation, and tolerance to higher xenobiotic concentrations [67].

In terms of phenoxy herbicide removal, there has been little investigation of the plant-mediated removal of 2,4-D and/or MCPA. For example, Ramborger et al. [68] evaluated the phytoremediation potential of Plectranthus neochilus (tea) exposed to the commercial pesticide containing 2,4-D (Aminol) in soil and water. The removal rate for 2,4-D reached 49% during 60 days, and the herbicide was not detected in plant leaves. Despite the fact that the phytoremediation potential of P. neochilus in soil was not sufficient, the plant exhibited satisfactory resistance to herbicide application. Moreover, the presence of phenolic compounds (e.g., ferulic and coumaric acid) in tea tissues indicated the ability of these plants to provide defense mechanisms against 2,4-D. The mechanism of the herbicide in plant begins by affecting the plasma membrane properties, subsequently leading to poor performance of mitochondria and peroxisomes [69]. In consequence, it stimulates the overexpression of abscisic acid (ABA) and ethylene biosynthesis genes, leading to significant changes of cellular redox potential by the production of reactive oxygen species (ROS) [70]. The occurrence of ROS leads to the production of phenolic compounds, i.e., ferulic acid and coumaric acid, which are responsible for the antioxidant self-defense mechanism of the plant against the herbicide. These phenolic compounds were found in higher concentrations only in plants that were exposed to 2,4-D and not in the controls.

3.3 Rhizoremediation

As mentioned above, plants play a key role in soil ecosystems by stabilizing the soil structure and by serving as primary sources of organic matter and energy which stimulate soil microbial activity [63]. Despite this, they are not the only contributors in the efficient phytoremediation of organic contaminants. Due to existing interactions between plant roots, root exudates, soil, and microorganisms, it has been proposed that the most effective method for the remediation of contaminated soil may be microbe-assisted phytoremediation (rhizoremediation).

Rhizoremediation is a naturally occurring process within the plant root zone (rhizosphere), where the growth of microorganisms and their degradative activity are stimulated by root exudates enriched by plant secondary metabolites (PSMs). Plant-derived compounds can [1] serve as primary substrates in cometabolism and provide energy for microbial growth [2], act as inducers of degradative enzymes due to their structural similarities to xenobiotics, and [3] enhance the degree of contamination removal by increasing pollutant bioavailability in soil [71].

The effectiveness of rhizospheral biodegradation depends also on the potential of the microorganisms inhabiting the rhizosphere to adapt to pollutant concentrations [72]. For effective degradation of contaminants to take place, a wide range of plants and bacterial traits is needed, involving the orchestrated interaction of a multitude of genes and enzymes. Rhizoremediation can therefore be optimized by selecting suitable plant-microbe sets, which can be achieved by combining plant and plant growth-promoting rhizobacteria (PGPR) [73] and/or microbes capable of contaminant degradation [74]. PGPR can improve phytoremediation efficiency by enhancing plant tolerance to various environmental stresses, promoting root growth and improving plant growth and health. In turn, some rhizospheral microorganisms can directly use their own degradative capabilities to metabolize organic pollutants [74, 75]. A study of rhizosphere-enhanced biodegradation of 2,4-D by Boyle et al. [76] found a significant difference in the mineralization of 2,4-D between monocot rhizosphere soils, dicot rhizosphere soils, and non-rhizosphere soils, with greater microbial activity being observed in monocot rhizosphere soil than in dicot rhizosphere soil or bulk soil. Therefore, both the soil and plant species determine the mineralization of tested contaminant. According to Shaw and Burns [77], the amendment of soil with 2,4-D increased the number of rhizospheric bacteria degrading 2,4-D in Trifolium pratense (red clover). Germaine et al. [78] also note the abundance of 2,4-D degraders in the stem and leaves of pea plant and that, under exposure to phenoxy herbicide, pea plants developed a stubby root system.

Furthermore, it has been hypothesized that PSMs may have a profound impact on the biodegradation of xenobiotics by providing the energy for microorganisms to carry out cometabolism; in this case, the xenobiotic is degraded as a secondary substrate [45, 71, 72, 73]. PSMs can be used as a primary source of carbon for bacterial communities to support their growth and stimulate the expression of desirable genes involved in the catabolic pathway of given xenobiotic. This is evident in the case of biphenyl, naringin, coumarin, myricetin, and l-carvone, which stimulate the activity of polychlorinated biphenyl (PCB)-degrading bacteria such as A. eutrophus, Corynebacterium sp., P. putida [79], and Arthrobacter sp. strain B1B [80]. Another example of PSM-stimulated PCB biodegradation was identified in mulberry (Morus sp.). In this case, the PSMs morusin, morusinol, and kuwanon C have been found to support the growth of the PCB-degrading bacterium Burkholderia sp. LB400 [81]. Likewise, the PSM (cumene) stimulates the activity of TCE-degrading R. gordonia bacteria [85]. According to Yi et al. [82], salicylic and linoleic acids, excreted by Raphanus sativus, enhanced the bioavailability of polycyclic aromatic hydrocarbons (PAHs) and increased the effectivity of their removal form soil. According to Ely and Smets [83], PAH biodegradation is stimulated by the presence of phenolic compounds, flavonoids, and gibberellic acid. Compounds such as acetophenone, phenethyl alcohol, p-hydroxybenzoic acid, and trans-cinnamic acid enhance the biotransformation of cis-1,2-dichloroethylene [84].

In addition, it has been hypothesized that PSMs may also induce the detoxification mechanisms taking place in bacterial cells [85, 86]. The expression of functional genes in bacteria is essential for the successful bioremediation of xenobiotics and can be stimulated by PSMs in different ways. However, very little information is given in the literature regarding the influence of PSMs on the induction of genes involved in catabolic pathways. Siciliano et al. [87] report greater induction of catabolic genes (ndoB, alkB, xylE) involved in the degradation of naphthalene in the rhizosphere soil of Festuca arundinacea (tall fescue) than in unplanted soil. Salicylate has been reported to have an upregulating effect on the expression of bphA, which encodes biphenyl dioxygenase in the PCB degrader Pseudomonas sp. Cam-1 [88]. The presence of salicylic acid was found to enhance the expression of the bphA gene in R. eutropha H850 and P. fluorescens P2W [89].

In addition, it has been hypothesized that the structural similarity between selected xenobiotics and PSMs may have a profound impact on the biodegradation of given, structurally related xenobiotic [71, 85]. For example, Urbaniak et al. [35] demonstrated the effect of a PSM, syringic acid, on the enhanced removal of structurally similar herbicide, MCPA, by indigenous soil bacteria, with greater MCPA depletion being achieved in samples enriched with PSM. The molecular analysis revealed ubiquitous enrichment of the samples with Rhodoferax spp., Achromobacter spp., Burkholderia spp., and Cupriavidus spp., which are commonly known as MCPA degraders. Also, a study by McLoughlin et al. [89] found the PSMs limonene and α-pinene to enhance 2,4-DCP degradation, but only following pre-exposure to both 2,4-DCP and monoterpene, with total 2,4-DCP mineralization extents of up to 71%.

Taking into account the abovementioned aspects, rhizoremediation can serve as a potential tool for phenoxy herbicide removal from soil ecosystems. However, to date, most studies have focused solely on the phyto- or biodegradation properties of plants or bacteria [71]. Consequently only limited data is available in terms of the impact of rhizoremediation on phenoxy herbicide removal from soil.

3.4 Endophyte-enhanced phytoremediation

Endophytic bacteria that reside inside plant tissues are also known to play a crucial role in the remediation of organic compounds. Plant-associated bacteria can enhance plant growth and degrade organic contaminants such as trichloroethylene and hydrocarbons [90]. The activity of endophytic bacteria can mitigate and improve plant conditions in stressful environments (such as contaminated soils). Field studies by Eevers et al. [91] showed that zucchini (Cucurbita pepo) plants inoculated with a consortium of three plant growth-promoting endophytic strains demonstrated an increased concentration of dichloro-bis(p-chlorophenyl)ethylene (DDE) in the aerial parts. The amount of DDE accumulated in C. pepo per growing season was significantly higher for inoculated plants. Thus such an approach might be promising for phytoremediation applications.

It has also been found that application of 2,4-D (1.42, 2.84, and 5.68 mg a.i./g soil) had a negative effect on the physio-morphological parameters of aerobic rice (Oryza sativa) and reduced the number of plant endophytes [92]; however, inoculation of seeds with the endophytic bacteria strain Stenotrophomonas maltophilia improved plant characteristics under herbicide-stressed soils. S. maltophilia has previously been described as a plant growth-promoting endophytic strain with the ability to produce auxins and siderophores [92].

Bacterial endophyte-enhanced phytoremediation was also studied by Germaine et al. [78] on the example of P. sativum: plants were inoculated with genetically tagged endophytic bacteria, which naturally possess the ability to biodegrade 2,4-D. The inoculated plants not only displayed more efficient herbicide removal but also demonstrated a lack of 2,4-D accumulation in their aerial parts. Additionally the endophytic strain protected the pea plant from the toxic effects of 2,4-D, resulting in a greater increase of plant biomass and thus greater 2,4-D transportation to the aboveground parts of the plant from the soil.

Table 5 compares the presented biological methods of remediation of soils contaminated with phenoxy herbicides. It illustrates the differences of the removal of phenoxy herbicides from soil. It is apparent from this table that the most efficient method of contaminant removal is endophyte-assisted phytoremediation; however, more research on this topic needs to be undertaken before the association between role of symbiotic microorganisms and plants in removal of contaminants from environmental matrices is more clearly understood.

MethodCompoundInitial concentration of compound used in an experimentDuration of an experiment (days)Removal of phenoxyacetic acid (%)CommentsSource
Bioremediation2,4-D and MCPA0.09 mmol/kg of soil11860Activity of bulk soil microbial population from various soil samples[93]
2,4-D1.8 kg/ha1045–48Activity of bulk soil microbial population from clay and loamy soil samples[94]
Phytoremediation and rhizoremediation2,4-D11.42 kg/ha2049Use of P. neochilus for phytoremediation[76]
2,4-D
2,4-DCP
1.22 10−3 μm
1.19 10−3 μmol
66~60Activity of rhizospheric soil bacteria derived from monocots[68]
Endophyte-enhanced phytoremediation2,4-D47–360 mg/kg of soil5393–100The inoculation of P. sativum by endophytic bacteria P. putida VM1450

Table 5.

Biological remediation methods and % average removal of phenoxy herbicides from soil matrices.

Advertisement

4. Conclusions

Uncontrolled use of phenoxy herbicides (2,4-D and MCPA) in the agricultural and gardening sector can result in their dispersal in soil and water ecosystems, which can significantly disturb the sustainability of the environment and increase its ecotoxicity level. Although their persistence in soil is limited due to their chemical characteristics, they can be transported and accumulated in water ecosystems through runoff and leaching. According to recent reports, phenoxy herbicides are especially toxic for plants, freshwater crustaceans, and amphibians; hence there is a growing need to limit the release of phenoxy acids in natural environments.

Taking into account the abovementioned aspects, the integration of bio-, phyto-, and rhizoremediation can serve as a potential tool for phenoxy herbicide removal from soil ecosystems. The ability of bacteria to metabolize phenoxy herbicides has been extensively studied over the last decades. However, to date, only limited data is available in terms of the impact of phyto- and rhizoremediation on phenoxy herbicide removal from soil. What is not yet clear is the impact of PSMs on the degradation of phenoxy herbicides. The similarity of the chemical structure of chosen PSMs and xenobiotics can be reflected in the xenobiotic degradation rates, e.g., the presence and induction of degradative genes and production of degradative enzymes, and the composition of microbial populations. To date, little evidence has been found associating the removal of phenoxy herbicides using both plants and microorganisms. However, the abovementioned research serves as a base for future studies on their application for the improvement of soil quality.

Considering the above, the chapter describes an interdisciplinary approach to tackling the problem of environmental phenoxy acid herbicide contamination through integrating available literature data on the physicochemical properties of 2,4-D and MCPA, as well as their levels in the environment and toxicity to the organisms from different trophic levels. It also outlines possible methods for their removal using nature-based techniques such as bio-, phyto-, and rhizoremediation.

Advertisement

Acknowledgments

This work was supported by the European Structural and Investment Funds, OP RDE-funded project “CHEMFELLS4UCTP” (No. CZ.02.2.69/0.0/0.0/17_050/0008485).

Advertisement

Conflict of interest

There is no conflict of interest.

References

  1. 1. Atwood D, Paisley-Jones C. 2008-2012 Market Estimates. Pestic Ind Sales Usage; 2017
  2. 2. Eurostat. Sales of pesticides by type of pesticide
  3. 3. Smith AE, Mortensen K, Aubin AJ, Molloy MM. Degradation of MCPA, 2,4-D, and other phenoxyalkanoic acid herbicides using an isolated soil bacterium. Journal of Agricultural and Food Chemistry. 1994;42(2):401-405
  4. 4. Ma Q , Ding Y, Chang J, Sun X, Zhang L, Wei Q , et al. Comprehensive insights on how 2,4-dichlorophenoxyacetic acid retards senescence in post-harvest citrus fruits using transcriptomic and proteomic approaches. Journal of Experimental Botany. 2014;65(1):61-74
  5. 5. Kamrin MA, editor. Phenoxy and benzoic acid herbicides. In: Pesticide Profiles. New York: CRC Press; 1997. pp. 299-224
  6. 6. Skiba E, Wolf WM. Commercial phenoxyacetic herbicides control heavy metal uptake by wheat in a divergent way than pure active substances alone. Environmental Sciences Europe. 2017;29(1):26
  7. 7. Paszko T, Muszyński P, Materska M, Bojanowska M, Kostecka M, Jackowska I. Adsorption and degradation of phenoxyalkanoic acid herbicides in soils: A review. Environmental Toxicology and Chemistry. 2016;35(2):271-286
  8. 8. Gavrilescu M. Fate of pesticides in the environment and its bioremediation. Engineering in Life Sciences. 2005;5(6):497-526
  9. 9. Parajulee A, Lei YD, Cao X, McLagan DS, Yeung LWY, Mitchell CPJ, et al. Comparing winter-time herbicide behavior and exports in urban, rural, and mixed-use watersheds. Environmental Science: Processes & Impacts. 2018;20(5):767-779
  10. 10. Agency USEP. Reregistration Eligibility Decision (RED) 2,4-D. Washington, DC: EPA 738-R-05-002; 2005
  11. 11. Gupta M, Garg NK, Joshi H, Sharma MP. Persistence and mobility of 2,4-D in unsaturated soil zone under winter wheat crop in sub-tropical region of India. Agriculture, Ecosystems and Environment. 2012;146(1):60-72
  12. 12. Ignatowicz K, Struk-Sokołowska J. Sezonowe wahania zanieczyszczeń agrotechnicznych w rzece Narwi ze szczególnym uwzględnieniem herbicydów fenoksyoctowych. Środkowo-Pomorskie Tow Nauk Ochr Środowiska. 2004;4:189-205
  13. 13. Birch GF, Drage DS, Thompson K, Eaglesham G, Mueller JF. Emerging contaminants (pharmaceuticals, personal care products, a food additive and pesticides) in waters of Sydney estuary, Australia. Marine Pollution Bulletin. 2015;97(1-2):56-66
  14. 14. Ismail BS, Prayitno S, Tayeb MA. Contamination of rice field water with sulfonylurea and phenoxy herbicides in the Muda Irrigation Scheme, Kedah, Malaysia. Environmental Monitoring and Assessment. 2015;187(7):406
  15. 15. Kucharski M, Domaradzki K. Changes in soil contamination by selected herbicides used in protection of cereals. Polish Journal of Soil Science. 2014;47(2):81-82
  16. 16. Metcalfe CD, Helm P, Paterson G, Kaltenecker G, Murray C, Nowierski M, et al. Pesticides related to land use in watersheds of the Great Lakes basin. Science of the Total Environment. 2019;648:681-692
  17. 17. Palma P, Matos C, Alvarenga P, Köck-Schulmeyer M, Simões I, Barceló D, et al. Ecological and ecotoxicological responses in the assessment of the ecological status of freshwater systems: A case-study of the temporary stream Brejo of Cagarrão (South of Portugal). Science of the Total Environment. 2018;634:394-406
  18. 18. Chiron S, Comoretto L, Rinaldi E, Maurino V, Minero C, Vione D. Pesticide by-products in the Rhône delta (Southern France). The case of 4-chloro-2-methylphenol and of its nitroderivative. Chemosphere. 2009;74(4):599-604
  19. 19. Kucharski M, Urbanowicz J. Badanie pozostałości linuronu i MCPA w glebie i roślinach ziemniaka. Biuletyn Instytutu Hodowli i Aklimatyzacji Roślin. 2008;248:61-66
  20. 20. López-Roldán R, Jubany I, Martí V, González S, Cortina JL. Ecological screening indicators of stress and risk for the Llobregat river water. Journal of Hazardous Materials. 2013;263:239-247
  21. 21. Salvo LM, Malucelli MIC, da Silva JRMC, Alberton GC, Silva De Assis HC. Toxicity assessment of 2,4-D and MCPA herbicides in primary culture of fish hepatic cells. Journal of Environmental Science and Health, Part B. Pesticides, Food Contaminants, and Agricultural Wastes. 2015;50(7):449-455
  22. 22. Wei YD, Zheng HG, Hall JC. Role of auxinic herbicide-induced ethylene on hypocotyl elongation and root/hypocotyl radial expansion. Pest Management Science. 2000;56(5):377-387
  23. 23. Pazmiño DM, Rodríguez-Serrano M, Romero-Puertas MC, Archilla-Ruiz A, del Río LA, Sandalio LM. Differential response of young and adult leaves to herbicide 2,4-dichlorophenoxyacetic acid in pea plants: Role of reactive oxygen species. Plant, Cell and Environment. 2011;34(11):1874-1889
  24. 24. Martínez-Ruiz EB, Martínez-Jerónimo F. Exposure to the herbicide 2,4-D produces different toxic effects in two different phytoplankters: A green microalga (Ankistrodesmus falcatus) and a toxigenic cyanobacterium (Microcystis aeruginosa). Science of the Total Environment. 2018;619(620):1566-1578
  25. 25. Li K, Wu JQ , Jiang LL, Shen LZ, Li JY, He ZH, et al. Developmental toxicity of 2,4-dichlorophenoxyacetic acid in zebrafish embryos. Chemosphere. 2017;171:40-48
  26. 26. Sarikaya R, Yilmaz M. Investigation of acute toxicity and the effect of 2,4-D (2,4-dichlorophenoxyacetic acid) herbicide on the behavior of the common carp (Cyprinus carpio L., 1758; Pisces, Cyprinidae). Chemosphere. 2003;52(1):195-201
  27. 27. Aronzon CM, Sandoval MT, Herkovits J, Pérez-Coll CS. Stage-dependent toxicity of 2,4-dichlorophenoxyacetic on the embryonic development of a south American toad, Rhinella arenarum. Environmental Toxicology. 2011;26(4):373-381
  28. 28. Estevam EC, Nakano E, Kawano T, de Bragança Pereira CA, Amancio FF, de Albuquerque Melo AMM. Dominant lethal effects of 2,4-D in Biomphalaria glabrata. Mutation Research, Genetic Toxicology and Environmental Mutagenesis. 2006;611(1-2):83-88
  29. 29. Bukowska B. Effects of 2,4-D and its metabolite 2,4-dichlorophenol on antioxidant enzymes and level of glutathione in human erythrocytes. Comparative Biochemistry and Physiology, Part C: Toxicology & Pharmacology. 2003;135(4):435-441
  30. 30. Perkins EJ, Stiff CM, Lurquin PF. Use of Alcaligenes eutrophus as a source of genes for 2,4-D resistance in plants. Weed Science. 1987;35(S1):12-18
  31. 31. Taylor SG, Shilling DG, Quesenberry KH, Chaudhry GR. Phytotoxicity of 2,4-D and 2,4-dichlorophenol to red clover (Trifolium pratense). Weed Science. 1989;37(6):825-829
  32. 32. Schweigert N, Hunziker R, Escher B, Eggen R. Acute toxicity of (chloro-)catechol-copper combinations in Escherichia coli corresponds to their membrane toxicity in vitro. Environmental Toxicology and Chemistry. 2001;20(2):239-247
  33. 33. Podolska G. The effectiveness and phytotoxicity of herbicide in buckwheat cv. Kora. Polish Journal of Agronomy. 2014;19:17-24
  34. 34. Mierzejewska E, Baran A, Urbaniak M. The influence of MCPA on soil phytotoxicity and the presence of genes involved in its biodegradation. Archives of Environmental Protection. 2017;44(4):58-64
  35. 35. Urbaniak M, Mierzejewska E, Tankiewicz M. The stimulating role of syringic acid, a plant secondary metabolite, in the microbial degradation of structurally-related herbicide, MCPA. Peer J. 2019;7:e6745
  36. 36. Weerakoon HPAT, Atapaththu KSS, Asanthi HB. Toxicity evaluation and environmental risk assessment of 2-methyl-4-chlorophenoxy acetic acid (MCPA) on non-target aquatic macrophyte Hydrilla verticillata. Environmental Science and Pollution Research. 2018;25(30):30463-30474
  37. 37. Kitagawa W, Kamagata Y. Diversity of 2,4-dichlorophenoxyacetic acid (2,4-D)-degradative genes and degrading bacteria. In: Nojiri H, Fukuda M, Tsuda M, Kamagata Y, editors. Biodegradative Bacteria: How Bacteria Degrade, Survive, Adapt, and Evolve. Japan: Springer; 2014. pp. 43-57
  38. 38. Kitagawa W, Takami S, Miyauchi K, Masai E, Kamagata Y, Tiedje JM, et al. Novel 2,4-dichlorophenoxyacetic acid degradation genes from oligotrophic bradyrhizobium sp. strain HW13 isolated from a pristine environment. Journal of Bacteriology. 2002;184(2):509-518
  39. 39. Kamagata Y, Fulthorpe RR, Tamura K, Takami H, Forney LJ, Tiedje JM. Pristine environments harbor a new group of oligotrophic 2, 4-dichlorophenoxyacetic acid-degrading bacteria. Applied Environmental Microbiology. 1997;63(6):2266-2272
  40. 40. Itoh K, Kanda R, Sumita Y, Kim H, Kamagata Y, Suyama K, et al. tfdA-like genes in 2,4-dichlorophenoxyacetic acid-degrading bacteria belonging to the Bradyrhizobium-Agromonas-Nitrobacter-Afipia cluster in α-proteobacteria. Applied and Environmental Microbiology. 2002;68(7):3449-3454
  41. 41. Itoh K, Tashiro Y, Uobe K, Suyama K, Yamamoto H. Root nodule Bradyrhizobium spp. harbor acid-degrading proteins homologous with genes encoding 2, 4-dichlorophenoxyacetic acid-degrading proteins. Applied and Environmental Microbiology. 2004;70:2110-2118
  42. 42. Mcgowan C, Fulthorpe R, Wright A, Tiedje JM. Evidence for interspecies gene transfer in the evolution of 2,4-dichlorophenoxyacetic acid degraders. Applied and Environmental Microbiology. 1998;64(10):4089-4092
  43. 43. Poll C, Pagel H, Devers-Lamrani M, Martin-Laurent F, Ingwersen J, Streck T, et al. Regulation of bacterial and fungal MCPA degradation at the soil-litter interface. Soil Biology and Biochemistry. 2010;42(10):1879-1887
  44. 44. Ka JO, Holben WE, Tiedje JM. Analysis of competition in soil among 2,4-dichlorophenoxyacetic acid-degrading bacteria. Applied and Environmental Microbiology. 1994;60(4):1121-1128
  45. 45. Shimojo M, Kawakami M, Amada K. Analysis of genes encoding the 2,4-dichlorophenoxyacetic acid-degrading enzyme from Sphingomonas agrestis 58-1. Journal of Bioscience and Bioengineering. 2009;108(1):56-59
  46. 46. Gözdereliler E, Boon N, Aamand J, De Roy K, Granitsiotis MS, Albrechtsen HJ, et al. Comparing metabolic functionalities, community structures, and dynamics of herbicide-degrading communities cultivated with different substrate concentrations. Applied and Environmental Microbiology. 2013;79(1):367-375
  47. 47. Nielsen TK, Xu Z, Gözdereliler E, Aamand J, Hansen LH, Sørensen SR. Novel insight into the genetic context of the cadAB genes from a 4-chloro-2-methylphenoxyacetic acid-degrading Sphingomonas. Stevenson B, editor. PLoS ONE. 2013;8(12):e83346
  48. 48. Müller RH, Jorks S, Kleinsteuber S, Babel W. Comamonas acidovorans strain MC1: A new isolate capable of degrading the chiral herbicides dichlorprop and mecoprop and the herbicides 2,4-D and MCPA. Microbiological Research. 1999;154(3):241-246
  49. 49. Vallaeys T, Albino L, Soulas G, Wright AD, Weightman AJ. Isolation and characterization of a stable 2,4-dichlorophenoxyacetic acid degrading bacterium, Variovorax paradoxus, using chemostat culture. Biotechnology Letters. 1998;20(11):1073-1076
  50. 50. González AJ, Gallego A, Gemini VL, Papalia M, Radice M, Gutkind G, et al. Degradation and detoxification of the herbicide 2,4-dichlorophenoxyacetic acid (2,4-D) by an indigenous Delftia sp. strain in batch and continuous systems. International Biodeterioration and Biodegradation. 2012;66(1):8-13
  51. 51. Han L, Zhao D, Li C. Isolation and 2,4-D-degrading characteristics of Cupriavidus campinensis BJ71. Brazilian Journal of Microbiology. 2015;46(2):433-441
  52. 52. Long ZX, Yan Z, Xin Z, Li YS. Biodegradation of the herbicide 2, 4-dichlorophenoxyacetic acid by a new isolated strain of Achromobacter sp. LZ35. Current Microbiology. 2017;74(2):193-202
  53. 53. Maltseva O, Mcgowan C, Fulthorpet R, Oriel P. Degradation of 2,4-dichlorophenoxyacetic acid by haloalkaliphilic bacteria. Microbiology. 1 May 1996;142(5):1115-1122
  54. 54. Ka JO, Holben WE, Tiedje JM. Use of gene probes to aid in recovery and identification of functionally dominant 2,4-dichlorophenoxyacetic acid-degrading populations in soil. Applied and Environmental Microbiology. 1994;60(4):1116-1120
  55. 55. Lappin HM, Greaves MP, Slatert JH. Degradation of the herbicide mecoprop [2-(2-methyl-4-chlorophenoxy) propionic acid] by a synergistic microbial community. Applied and Environmental Microbiology. 1985;49(2):429-433
  56. 56. Bælum J, Henriksen T, Christian H, Hansen B, Jacobsen CS. Degradation of 4-chloro-2-methylphenoxyacetic acid in top- and subsoil is quantitatively linked to the class III tfdA gene. Applied and Environmental Microbiology. 2006;72(2):1476-1486
  57. 57. Fukumori F, Hausinger RP. Alcaligenes eutrophus JMP134 “2,4-dichlorophenoxyacetate monooxygenase” is an alpha-ketoglutarate-dependent dioxygenase. Journal of Bacteriology. 1993;175(7):2083-2086
  58. 58. Itoh K, Tashiro Y, Uobe K, Kamagata Y, Suyama K, Yamamoto H. Root nodule bradyrhizobium spp. Harbor tfdA and cadA, homologous with genes encoding 2,4-dichlorophenoxyacetic acid-degrading proteins. Applied and Environmental Microbiology. 2004;70(4):2110-2118
  59. 59. Pieper DH, Reineke W, Engesser K-H, Knackmuss H-J. Metabolism of 2,4-dichlorophenoxyacetic acid, 4-chloro-2-methylphenoxyacetic acid and 2-methylphenoxyacetic acid by Alcaligenes eutrophus JMP 134. Archives of Microbiology. 1988;150(1):95-102
  60. 60. Schleinitz KM, Kleinsteuber S, Vallaeys T, Babel W. Localization and characterization of two novel genes encoding stereospecific dioxygenases catalyzing 2(2,4-dichlorophenoxy)propionate cleavage in Delftia acidovorans MC1. Applied and Environmental Microbiology. 2004;70(9):5357-5365
  61. 61. Liu Y, Liu S, Drake HL, Horn MA. Consumers of 4-chloro-2-methylphenoxyacetic acid from agricultural soil and drilosphere harbor cadA, r/sdpA, and tfdA-like gene encoding oxygenases. FEMS Microbiology Ecology. 2013;86:114-129
  62. 62. Gerhardt KE, Huang X, Glick BR, Greenberg BM. Plant science phytoremediation and rhizoremediation of organic soil contaminants: Potential and challenges. Plant Science. 2009;176:20-30
  63. 63. Machado F, Anderson C, Meenken E, Gillespie R, Peterson M, Harold M. The importance of plants to development and maintenance of soil structure, microbial communities and ecosystem functions. Soil and Tillage Research. 2018;175:139-149
  64. 64. Singer AC, Crowley DE, Thompson IP. Secondary plant metabolites in phytoremediation and biotransformation. Trends in Biotechnology. 2003;21(3):123-130
  65. 65. Reshma AC, Krishna RR. Plant species identification for phytoremediation of mixed contaminated soils. Journal of Hazardous, Toxic, and Radioactive Waste. 2012;19:218-229
  66. 66. Siwek M. Biologiczne sposoby oczyszczania srodowiska—fitoremediacja. Wiadomości Botaniczne. 2008;52(1/2):23-28
  67. 67. Posmyk K, Urbaniak M. Fitoremediacja jako alternatywna metoda oczyszczania środowiska. Aura. 2014;7:10-12
  68. 68. Ramborger BP, Ortis Gularte CA, Rodrigues DT, Gayer MC, Sigal Carriço MR, Bianchini MC, et al. The phytoremediation potential of Plectranthus neochilus on 2,4-dichlorophenoxyacetic acid and the role of antioxidant capacity in herbicide tolerance. Chemosphere. 2017;188:231-240
  69. 69. Rodríguez-Serrano M, Pazmiño DM, Sparkes I, Rochetti A, Hawes C, Romero-Puertas MC, et al. 2,4-Dichlorophenoxyacetic acid promotes S-nitrosylation and oxidation of actin affecting cytoskeleton and peroxisomal dynamics. Journal of Experimental Botany. 2014;65(17):4783-4793
  70. 70. Grossmann K. Auxin herbicides: Current status of mechanism and mode of action. Pest Management Science. 2009;66(2):113-120
  71. 71. Musilova L, Ridl J, Polivkova M, Macek T, Uhlik O. Effects of secondary plant metabolites on microbial populations: Changes in community structure and metabolic activity in contaminated environments. Iriti M, editor. International Journal of Molecular Sciences. 2016;17(8):1205
  72. 72. Lugtenberg BJ, Dekkers L, Bloemberg GV. Molecular determinants of rhizosphere colonization by pseudomonas. Annual Review of Phytopathology. 2001;39(1):461-490
  73. 73. Ahemad M, Kibret M. Mechanisms and applications of plant growth promoting rhizobacteria: Current perspective. Journal of King Saud University—Science. 2014;26(1):1-20
  74. 74. Kuiper I, Lagendijk EL, Bloemberg GV, Lugtenberg BJJ. Rhizoremediation: A beneficial plant-microbe interaction bioremediation: A natural method. Molecular Plant-Microbe Interactions. 2004;17(1):6-15
  75. 75. Glick BR. Using soil bacteria to facilitate phytoremediation. Biotechnology Advances. 2010;28(3):367-374
  76. 76. Boyle JJ, Shann JR. Biodegradation of phenol, 2,4-DCP, 2,4-D, and 2,4,5-T in field-collected rhizosphere and nonrhizosphere soils. Journal of Environmental Quality. 1995;24(4):782
  77. 77. Shaw LJ, Burns RG. Enhanced mineralization of [U-14C]2,4-dichloropheeoxyacetic acid in soil from the rhizosphere of Trifolium pratense. Applied and Environmental Microbiology. 2004;70(8):4766-4774
  78. 78. Germaine KJ, Liu X, Cabellos GG, Hogan JP, Ryan D, Dowling DN. Bacterial endophyte-enhanced phytoremediation of the organochlorine herbicide 2,4-dichlorophenoxyacetic acid. FEMS Microbiology Ecology. 2006;57(2):302-310
  79. 79. Donnelly PK, Hegde RS, Fletcher JS. Growth of PCB-degrading bacteria on compounds from photosynthetic plants. Chemosphere. 1994;28(5):981-988
  80. 80. Gilbert ES, Crowley DE. Plant compounds that induce polychlorinated biphenyl biodegradation by Arthrobacter sp. strain B1B. Applied and Environmental Microbiology. 1997;63(5):1933-1938
  81. 81. Leigh MB, Fletcher JS, Fu X, Schmitz FJ. Root turnover: An important source of microbial substrates in rhizosphere remediation of recalcitrant contaminants. Environmental Science & Technology. 2002;36(7):1579-1583
  82. 82. Yi H, Crowley DE. Biostimulation of PAH degradation with plants containing high concentrations of linoleic acid. Environmental Science & Technology. 2007;41(12):4382-4388
  83. 83. Ely CS, Smets BF. Bacteria from wheat and cucurbit plant roots metabolize PAHs and aromatic root exudates: Implications for rhizodegradation. International Journal of Phytoremediation. 3 Oct 2017;19(10):877-883
  84. 84. Fraraccio S, Strejcek M, Dolinova I, Macek T, Uhlik O. Secondary compound hypothesis revisited: Selected plant secondary metabolites promote bacterial degradation of cis-1,2-dichloroethylene (cDCE). Scientific Reports. 2017;7(1):8406
  85. 85. Hu C, Zhang Y, Tang X, Luo W. PCB biodegradation and bphA1 gene expression induced by salicylic acid and biphenyl with Pseudomonas fluorescence P2W and Ralstonia eutropha H850. Polish Journal of Environmental Studies. 2014;23(5):1591-1598
  86. 86. Uhlik O, Musilova L, Ridl J, Hroudova M, Vlcek C, Koubek J, et al. Plant secondary metabolite-induced shifts in bacterial community structure and degradative ability in contaminated soil. Applied Microbiology and Biotechnology. 2013;97(20):9245-9256
  87. 87. Siciliano SD, Germida JJ, Banks K, Greer CW, Lafayette W. Changes in microbial community composition and function during a polyaromatic hydrocarbon phytoremediation field trial. Applied and Environmental Microbiology. 2003;69(1):483-489
  88. 88. Master ER, Mohn WW. Induction of bphA, encoding biphenyl dioxygenase, in two polychlorinated biphenyl-degrading bacteria, Psychrotolerant Pseudomonas strain Cam-1 and Mesophilic Burkholderia strain LB400. Applied and Environmental Microbiology. 2001;67(6):2669-2676
  89. 89. McLoughlin E, Rhodes AH, Owen SM, Semple KT. Biogenic volatile organic compounds as a potential stimulator for organic contaminant degradation by soil microorganisms. Environmental Pollution. 2009;157(1):86-94
  90. 90. Eevers N, Hawthorne JR, White JC, Vangronsveld J, Weyens N. Exposure of Cucurbita pepo to DDE-contamination alters the endophytic community: A cultivation dependent vs a cultivation independent approach. Environmental Pollution. 2016;209:147-154
  91. 91. Eevers N, Hawthorne JR, White JC, Vangronsveld J, Weyens N, Hawthorne JR, et al. Endophyte-enhanced phytoremediation of DDE-contaminated using Cucurbita pepo: A field trial. International Journal of Phytoremediation. 21 Mar 2018;20(4):301-310
  92. 92. Nahi A, Othman R, Omar D. Effects of Sb16 bacterial strain and herbicides on endophytic bacterial populations and growth of aerobic rice. Plant, Soil and Environment. 2016;62(10):453-459
  93. 93. Bælum J, Prestat E, David MM, Strobel BW, Jacobsen CS, Park K. Modeling of phenoxy acid herbicide mineralization and growth of microbial degraders in 15 soils monitored by quantitative real-time PCR of the functional tfdA gene. Applied and Environmental Microbiology. 2012;78(15):5305-5312
  94. 94. Boivin A, Amellal S, Schiavon M, van Genuchten MT. 2,4-Dichlorophenoxyacetic acid (2,4-D) sorption and degradation dynamics in three agricultural soils. Environmental Pollution. 2005;138(1):92-99

Written By

Magdalena Urbaniak and Elżbieta Mierzejewska

Submitted: 30 April 2019 Reviewed: 27 June 2019 Published: 07 August 2019