Keywords

4.1 Introduction

1,4-Dioxane (C4H8O2) is an industrial solvent produced for diverse direct applications including raw material and as a solvent. Its chemical abstract service number (CAS no.) assigned by the American Chemical Society is 123-91-1 with synonyms including dioxan, diethylene dioxide, dioxane, p-dioxane, diethylene oxide, glycol ethylene ether, or diethylene ether (McFee et al. 1994). At laboratory scale, this colorless and flammable liquid is used as a stable reaction media and as an extraction liquid for vegetable and animal oils (ATSDR 2006). Dioxane is useful as a growth substrate in organisms (Barajas-Rodriguez et al. 2019) and solvent in paints, ink, cosmetics, detergent, varnishes, and cleaning and shining fluids. With its nonpolar and aprotic solvent as it is presented in Fig. 4.1, dioxane is in addition utilized in the processing of petrochemicals, pesticide, pharmaceuticals, plastics and rubber, explosives, polishing, and pulp and paper as a solvent (DES 2018).

Fig. 4.1
1,3 dioxane has a 6-carbon ring. C1 and C3 are replaced by O. C2, C4, C5, and C6 are replaced by C H 2. 1,4 dioxane has a 6-carbon ring. C1 and C4 are replaced by O. 1,4 dioxane has a chair confirmation ring. C1 and C4 are replaced by O. 1,2 dioxane has a 6-carbon ring. C1 and C2 are replaced by O.

Chemical structures of 1,4-dioxane (88.11 g/mol)

Among the chemical structures presented in Fig. 4.1, structure B containing two planes of symmetry is the most stable due to minimum atomic constraints. The laboratory-scale production is shown in Fig. 4.2, while the commercial production of dioxane follows reaction root shown in Fig. 4.3.

Fig. 4.2
An experimental reaction to produce 1,4 dioxane. The ethylene oxide reacts with H 2 O to yield diethylene glycol, followed by 1,4 dioxane by the elimination of H 2 O. The ethylene oxide has a 3-carbon ring. C1 is replaced by O. C 2 and C 3 are replaced by C H 2.

Laboratory production of 1,4-dioxane (van Buskirk 2014)

Fig. 4.3
A chemical reaction to produce 1,4 dioxane. 2 chloro 2 hydroxydiethyl ether reacts with 1,2 dibromoethane in the presence of 20% weight of N a O H and N a H S O 4 at the required temperature to yield 1,4 dioxane.

Industrial production of 1,4-dioxane (Wilbur and Jones 2012)

Through literature survey, there are no industries in the world that directly produce 1,4-dioxane, implying that its production might be sporadic. The produced dioxane above undergoes heating with acids, distillation, salting out with CaCl2, and distillation again in order to attain high level of purity. More than 90% of the produced dioxane is used for stabilization of chlorinated solvents (US EPA 2014a) such as trichloroethene and 1,1,1-trichloroethane (Milavec et al. 2020).

4.1.1 Properties of 1,4-Dioxane

Dioxane forms an adduct with 1,1,1-trichloroethane (TCE) that inhibits through poisoning the catalytic reaction between TCE and aluminum containers used for transportation of chlorinated solvents (Adamson et al. 2015). In most reactions and solvent requirements, dioxane has replaced tetrahydrofuran, which has high toxicity level than dioxane (Madhu 2018), high boiling point, hygroscopicity, diether chelating ligand (Tusher et al. 2021), and high water miscibility. According to the hard and soft Lewis acids and bases (HSAB) theory (Pinter et al. 2013), dioxane is a hard base implying that it is weakly polarizable, has high charge states, is less volatile, is nonpolar, and has small molecule (Davarani et al. 2012). Dioxane can react as a monodentate with uranium to produce useful materials (Fig. 4.4) for nuclear chemistry (Monreal et al. 2011).

Fig. 4.4
A structure of uranium-containing dioxane has a central U atom that is single bonded to two 1,4 dioxane compounds and dashed and wedge bonded to 2 I atoms.

Stable compounds of uranium-containing dioxane. [Image adopted with permission from Monreal et al. (2011)]

However, the bidentate ligand is complex due to bond constraints. The possible reaction reported in Fig. 4.5 indicates the reaction between dioxane and Grignard reagent to form a polymer (Che-revision Administrator 2014).

Fig. 4.5
A chemical reaction of 1,4 dioxane with Grignard reagent yields a polymer. A structure has a 1,4 dioxane in which C1 and C4 are replaced by O. C1 is bonded to a methyl group. C4 is bonded to M g, which is further wedged and dashed bonded to X. Also, M g is bonded to repeating units of dioxane.

Polymeric reaction of 1,4-dioxane

4.1.2 Sources and Occurrence of 1,4-Dioxane

Reports on the direct production of 1,4-dioxane are rare; however, it occurs in most consumer products such as in cosmetics as a trace contaminant (Hossein et al. 2022) that can penetrate human and animal skin (U.S. FDA 2016). Industrial manufacturing of chlorinated solvents, papers and pulp, agricultural pesticides, organic chemical, dyes, and rubber and textile processing are potential sources of dioxane (Sun et al. 2016). It is a by-product during the manufacturing of aircraft deicing fluids, dyes, antifreeze, greases, paint strippers, and other consumer products (Hossein et al. 2022). Its high water miscibility and use in domestic activities including cleansing and bathing, recreational swimming, industrial production, and automobile activities release substantial amount of dioxane in terrestrial and marine ecosystem (Karges et al. 2018). Through its environmental partitioning, dioxane is rarely available in the atmosphere due to low volatility (Adamson et al. 2021), surface water (Karges et al. 2020), soil (Hinchee et al. 2018), and groundwater (Yamamoto et al. 2018a). The oceanic occurrence of dioxane is also reported by Scaratti et al. (2020). The occurrence of dioxane in the body is rare since it is transformed to β-hydroxyethoxyacetic acid and eliminated through urination (ITRC 2020a). Its occurrence in the drinking water (Adamson et al. 2021), food (Broughton et al. 2019), and particulate matters (Lee and Choo 2013) has proved the occurrence of dioxane in all ecosystem compartments. Thus, mostly use and disposal of chlorinated compounds and consumer products, water and wastewater treatment plants, accidental releases, landfills, and pipe leakages are the major sources of dioxane to the environment.

The LD50 of dioxane in rats is 5170 mg/kg, while the no-observed adverse effect to human being is 400 mg/m3 (Supprenant 2012). Dioxane falls among EPA’s unregulated contaminant monitoring rule (UCMR) that qualifies it as an emerging contaminant (Suthersan et al. 2016; Sarma 2022), and in addition dioxins are classified as persistent organic pollutants (Miraji et al. 2021).

4.1.3 Effects and Fate of 1,4-Dioxane

Compared to other solvents such as ethanol whereby the USA alone produced 84% (6.3 million m3) of the world in the year 2020 (AFDC 2022), recent data of 1,4-dioxane production are limited. For example, in the year 1985, global production of dioxane was 14,000 tons (13.6 million liters) (Wiki 2022); in the year 1990, about 18 million pounds was produced (7.9 million liters); and in the year 2002, the USA alone produced ten million pounds (4.4 million liters) (Wilbur and Jones 2012) indicating a gradual decrease. The absence of recent data on the direct global production of dioxane is not clear, yet most consumer products such as automotive coolants, chemical manufacturing, and textile contain dioxane. Dioxane is volatile (EPA 2010) with atmospheric half-life of 1–3 days (US EPA 2014b). There are several federal standard guidelines for dioxane exposure. Some of them are shown in Table 4.1 (US EPA 2014a).

Table 4.1 Federal standard guidelines for 1,4-dioxane

Breathing of contaminated air (ATSDR 2006); skin adsorption via use of shampoo, toothpaste, and other cosmetics (Alsohaimi et al. 2020); ingestion via packaging material (Gi et al. 2018); food and food supplements/additives (Mo et al. 2022); pesticide remnants in agricultural products (Begum et al. 2016); groundwater (Chu et al. 2018); occupational exposure (US EPA 2014a); industrial effluents (Stepien et al. 2014); and use of spermicidal sponge (ATSDR 2012) are the potential roots of exposure to dioxane.

Experimental research on dioxane has established LD50 of about 2 g/kg for cats and rabbits, 5.7–5.9 g/kg in mice, and 3.15–4 g/kg in guinea pigs. In this study, kidney and liver were the chronic affected organs (OEHHA 1998). The United States Environmental Protection Agency (US EPA) has reported 35 μg/L of dioxane as a cancer risk level (US EPA 2012) and classified it as group B2 chemical that probably causes human carcinogenic effect (McElroy et al. 2019). The Office of Environmental Health Hazard Assessment (OEHHA) has recommended 3 ppb of dioxane as a maximum level in the drinking water (OEHHA 1998). Evidence of dioxane carcinogenicity is reported and summarized by the Scientific Committee on Consumer Safety in its report number SCCS/1570/15 indicating that while the EU classifies it as a carcinogen category 2, the IARC classifies it as a group 2B carcinogen chemical. The SCCS recommends 55 μg/day; Australia has the highest recommendation of 420 μg/day, while Japan has the lowest recommendation of 4.3 μg/day (US EPA 2012).

Short-term or acute exposure to 1,4-dioxane may result in various effects, but the common ones are drowsiness; nausea; irritation of the eyes, throat, and nose; and headache (ATSDR 2012; US EPA 2014a). Long-term chronic exposure is associated with drying and cracking of skin and liver, dermatitis, eczema, and kidney damage (ATSDR 2012; US EPA 2014a). Recently, there has been an exponential growth in the global reports regarding 1,4-dioxane as indicated in Fig. 4.6.

Fig. 4.6
A line graph plots documents versus year. The line begins at (1931, 0), remains constant until (1963, 0) with some fluctuations, follows an increasing trend, and ends at (2019, 335). All values are estimated.

Scopus analytic for the published papers mentioned 1,4-dioxane from 1931 to 2021

Exponential growth is an indication of global awareness on the environmental impact of 1,4-dioxane, with the leading research area being chemistry as indicated in Fig. 4.7.

Fig. 4.7
A pie chart provides the proportion of the subject areas for 1,4-dioxane. Chemistry, 35.9%. Material Science, 13.8%. Chemical Engineering, 12.1%. Biochemistry, 9.7%. Physics, 8.8%. Environmental science, 4.9%. Pharmacology, 4.3%. Engineering, 4.1%. Medical, 1.4%. Energy, 1.1%. Other, 3.9%.

Scopus analytic for the subject areas mentioned 1,4-dioxane from 1931 to 2021

Its occurrence as an impurity in most consumer products (Fuh et al. 2005) presents a challenge in its roots and mitigation approaches. Thus, apart from dioxane that will enter living organisms and causes adverse effects, the remaining will be contaminant in the environmental ecosystem that requires natural remediation for the ecosystem safety.

4.2 Remediation of 1,4-Dioxane

Chemical and physical properties of 1,4-dioxane lead to ineffective traditional remediation especially through chemical treatment methods (ITRC 2020b). Furthermore, these properties render conventional unit processes involved in drinking and wastewater treatment ineffective too (ITRC 2020b). Although under aerobic conditions 1,4-dioxane is biodegradable via direct metabolism or co-metabolism, the relative roles of these processes are dependent on both the concentration of 1,4-dioxane and the presence of co-contaminants (Zhang et al. 2017; He et al. 2018; Broughton et al. 2019; Polasko et al. 2019; ITRC 2020b; Johnson et al. 2020; Metabolism A, Cometabolism A, Biodegradation A, Tools MB 2021). Inhibitory co-contaminants frequently complicate the biological treatment of 1,4-dioxane (Johnson et al. 2020). Because of its uses as a solvent, wetting agent, and stabilizer for chlorinated solvents, used in metal vapor degreasing, 1,4-dioxane is frequently found with a variety of co-contaminants, including heavy metals like hexavalent chromium [Cr(VI)] (Zhang et al. 2017; He et al. 2018; Broughton et al. 2019; Polasko et al. 2019; ITRC 2020b; Johnson et al. 2020). Cr(VI) occurs naturally in groundwater due to geological formations, but it also has anthropogenic sources that can overlap with 1,4-dioxane.

Microbes that use 1,4-dioxane as a carbon or energy source, as well as those that co-metabolize it after growth on other organic substrates, can biodegrade it (Zhang et al. 2017; He et al. 2018; Broughton et al. 2019; Polasko et al. 2019; ITRC 2020b; Johnson et al. 2020). In this case, using microbes such as bacterial strains for biodegradation of 1,4-dioxane is of interest for environmental and public health safety. These techniques may be used to remediate 1,4-dioxane in contaminated water, wastewater, and soil. The composition of soil microbial community in contaminated soils is potential for the effectiveness of the remediation processes.

4.3 Soil Microbial Constituents

Soil comprises organic matters, minerals, gases, liquids, and living things, which contribute to supporting life (Haumaier et al. 2001; Pettit 2004; Amundson et al. 1998). In addition to a porous phase containing water and gases, soil consists of a solid phase comprised of minerals and organic matter (Haumaier et al. 2001; Amundson et al. 1998) as presented in Fig. 4.8. The solid, liquid, and gaseous states may coexist in the soil depending on the climatic conditions (Haumaier et al. 2001; Amundson et al. 1998). Climate; relief such as elevation, orientation, and slope of the terrain; organisms; and soil’s parent materials interact over time to produce soil. The produced soil continuously changes due to various physical, chemical, and biological processes, including weathering and related erosion (Velde 2013; Okon and Antia 2022). Soil ecologists view soil as an ecosystem having deep internal connections and complexity (Hopp and McDonnell 2009; Häring et al. 2012; Velde 2013; Reinhold-Hurek et al. 2015; Okon and Antia 2022).

Fig. 4.8
An illustration of the soil structure of the microbial constituents. The soil has a porous phase, and the presented constituents are water vapor, solid water, liquid water, soil, organic matters, and microbial communities.

Soil structure and some of its constituents

Soil microbial constituents are the main factor that determines the functions of soil microorganisms and how they impact soil characteristics and nutrient constituents (Haumaier et al. 2001; Amundson et al. 1998; Reinhold-Hurek et al. 2015). The first known bacteria and microbes on earth are thought to have originated in the waters between two and four billion years ago (Margulis and Sagan 1997; Nisbet and Sleep 2001; Altermann et al. 2006).

These bacteria could fix nitrogen, and as they grew over time, they released oxygen into the environment (Margulis and Sagan 1997; Nisbet and Sleep 2001; Altermann et al. 2006). This process is crucial because it impacts soil fertility and structure. There are several soil microorganisms including bacteria, actinomycetes, fungus, algae, and protozoa (Edwards and Fletcher 1988; Belnap 2001; Bhattarai et al. 2015; Balasubramanian 2017). Each group has traits that characterize and define their roles in the soil (Lopez de Ceballos et al. 1983; Shormanov et al. 2012; Sei et al. 2013; Li et al. 2018; Yamamoto et al. 2018b; Tusher et al. 2021; Wang et al. 2021b). Each gram of soil around and surrounding plant roots, or the rhizosphere, contains up to ten billion bacteria (Liu et al. 2007; Sei et al. 2013; Miao et al. 2019; Polasko et al. 2019), which can be utilized for remediation purposes and restoration of natural environments.

4.4 Bacteriological Transformation of 1,4-Dioxane in the Soil

In most environments, 1,4-dioxane does not biodegrade easily possibly due to their chemical stability and high miscibility to water (Klečka and Gonsior 1986; Suh and Mohseni 2004; Ghosh et al. 2010; Stepien et al. 2014). However, several microorganisms have been identified that can biodegrade 1,4-dioxane, either directly or through co-metabolism (Altermann et al. 2006; Inoue et al. 2016, 2018; Zhang et al. 2017; Li et al. 2020; Johnson et al. 2020; Ramalingam and Cupples 2020a; Zippilli et al. 2021; Murnane et al. 2021b; Kikani et al. 2021; Wang et al. 2021a). Microorganisms use 1,4-dioxane as a growth substrate during metabolic biodegradation, but growth is slow unless 1,4-dioxane concentrations are high to levels greater than 100 mg/L (Ramalingam and Cupples 2020a; Metabolism A, Cometabolism A, Biodegradation A, Tools MB 2021). An additional growth substrate such as methane must be supplied during co-metabolic biodegradation to support biomass growth and induce the appropriate 1,4-dioxane-degrading enzymes (Gedalanga et al. 2014; US EPA 2017; Zhang et al. 2017; Aoyagi et al. 2018; He et al. 2018; Hamid et al. 2020; Ramalingam and Cupples 2020a, b; Murnane et al. 2021b). Unlike metabolic degradation, co-metabolic processes can reduce 1,4-dioxane to very low concentrations, which is important for greener ecosystems. These processes may be used separately or in combination based on needs. Details of 1,4-dioxane degradation conditions are presented in Table 4.2.

Table 4.2 1,4-Dioxane degradation conditions

Some bacterial strains such as Amycolata sp. CB1190 and P. carboxydivorans RM-31 are capable of degradation of 1,4-dioxane in various reaction conditions. The utility of these microbes for environmental remediation of contaminants is undeniable. For example, Mahendra and colleagues investigated 1,4-dioxane, a probable human carcinogen and an important emerging contaminant (Mahendra and Alvarez-Cohen 2006). Their results revealed that among 20 bacteria strains, 13 were capable of biodegrading dioxane (Mahendra and Alvarez-Cohen 2006), indicating potential utility in remediation purposes.

The application of bacteria strains in the biodegradation of contaminants, including 1,4-dioxane, was also reported by several researchers (Metsä-Ketelä et al. 2013; Hamid et al. 2020; Murnane et al. 2021b).

4.4.1 Aerobic Biotransformation

Although they are uncommon, several microorganisms that can metabolize 1,4-dioxane in aerobic conditions have been isolated. Among the microbes that can biotransform 1,4-dioxane in aerobic conditions, Pseudonocardia dioxanivorans CB1190 is the best studied strain (Arve 2015; Isaka et al. 2016; Zhang et al. 2017; Aoyagi et al. 2018; Guan et al. 2018; He et al. 2018; Miao et al. 2018; Barajas-Rodriguez et al. 2019; Ramalingam and Cupples 2020b; Johnson et al. 2020; Dang and Cupples 2021; Metabolism A, Cometabolism A, Biodegradation A, Tools MB 2021). Industrial activated sludge was used to enrich the bacterium and then fed with tetrahydrofuran and finally 1,4-dioxane. Studies indicate that high levels of chlorinated solvents, their by-products, and some metals may inhibit 1,4-dioxane CB1190 strain degradation (Arve 2015; Isaka et al. 2016; Zhang et al. 2017; Aoyagi et al. 2018; Guan et al. 2018; He et al. 2018; Miao et al. 2018; Barajas-Rodriguez et al. 2019; Ramalingam and Cupples 2020b; Johnson et al. 2020; Dang and Cupples 2021; Metabolism A, Cometabolism A, Biodegradation A, Tools MB 2021). The most effective inhibitors of 1,4-dioxane degradation are presented in Fig. 4.9.

Fig. 4.9
4 chemical structures of 1,4 dioxane-based degradation inhibitors. The inhibitors are 1,1 dichloroethane, cis 1,2 dichloroethane, trichloroethane, and 1,1,1 trichloroethane.

Chemical structures of 1,4-dioxane degradation inhibitors (Source: Tomo Aoyagi et al.)

It is reported that about 5 mg/L of 1,1-dichloroethene is an inhibitor of 1,4-dioxane biodegradation (Arve 2015; Isaka et al. 2016; Zhang et al. 2017; Aoyagi et al. 2018; Guan et al. 2018; He et al. 2018; Miao et al. 2018; Barajas-Rodriguez et al. 2019; Ramalingam and Cupples 2020b; Johnson et al. 2020; Dang and Cupples 2021; Metabolism A, Cometabolism A, Biodegradation A, Tools MB 2021). The most effective metal inhibitor of 1,4-dioxane degradation by CB1190 was Cu(II), which lengthened the lag time at 1 mg/L and significantly decreased 1,4-dioxane degradation rates at 10 and 20 mg/L Cu(II). While Zn(II) did not impact biodegradation of 1,4-dioxane at the highest test concentration (20 mg/L Zn), Cd(II), Ni(II), and Ni(II) were less sensitive to 1,4-dioxane degradation (Arve 2015; Isaka et al. 2016; Zhang et al. 2017; Aoyagi et al. 2018; Guan et al. 2018; He et al. 2018; Miao et al. 2018; Barajas-Rodriguez et al. 2019; Ramalingam and Cupples 2020b; Johnson et al. 2020; Dang and Cupples 2021; Metabolism A, Cometabolism A, Biodegradation A, Tools MB 2021). This result indicates that 1,4-dioxane occurs as a mixture of several contaminants; therefore, the more effective approach will be to propose and improve techniques for co-biotransformation of 1,4-dioxane and its co-contaminants in the environments. For the treatment of industrial wastewater contaminated with 1,4-dioxane, biodegradation is a viable, economical, and environmentally benign solution. A schematic and proposed chemical equation for aerobic biotransformation is presented in Fig. 4.10.

Fig. 4.10
A pathway diagram for the production of diglycolic acid. The reactant dioxane reacts with P 450 to form dioxane 2 ol, which reacts reversibly to yield aldehyde. The aldehyde undergoes dehydrogenation to form beta hydroxyethoxyacetic acid, followed by diglycolic acid.

Series of chemical processes occurring during aerobic biotransformation of 1,4-dioxane (released under CC by 4)

In a study, Tusher and colleagues explored metabolic bacteria from a stable microbial community that degrades 1,4-dioxane (Tusher et al. 2021). Pseudonocardia sp. (TS28), Dokdonella sp. (TS32), and Afipia sp. (TS43) were three bacterial strains that were discovered to be capable to use 1,4-dioxane as their only source of carbon and energy to break down organic compounds (Tusher et al. 2021). This was a pioneer study that detailed the participation of the genus Dokdonella in the biodegradation of 1,4-dioxane as presented in Fig. 4.11. Genus Dokdonella possesses inducible 1,4-dioxane-degrading enzymes’ potential for remediation purposes (Tusher et al. 2021).

Fig. 4.11
A schematic representation of the biodegradation process of 1,4 dioxane in the presence of Dokdonella bacteria. The bacteria possess a 1,4 dioxane compound, resulting in fast intermediate, co-metabolism, and slow intermediate metabolisms.

Biodegradation of 1,4-dioxane by Dokdonella bacterium (released under CC by 4)

The findings of this study also contributed to our understanding of how various 1,4-dioxane degraders interact and cohabit in a consortium while utilizing a single carbon supply to create an effective biological 1,4-dioxane treatment system.

4.4.2 Aerobic Co-metabolic Biotransformation of 1,4-Dioxane

A simultaneous degradation method of two compounds is known as co-metabolism, whereby the degradation of a second compound always depends on the presence of the first compound. Organisms involved in co-metabolizing 1,4-dioxane, in contrast to 1,4-dioxane-metabolizing microorganisms, can break down 1,4-dioxane after growing on a main growth-supporting substrate (Kashimoto et al. 1989; Mahendra et al. 2013; US EPA 2014a, 2017; Arve 2015; Zhang et al. 2017, b; Johnson et al. 2020; Miao et al. 2020; Metabolism A, Cometabolism A, Biodegradation A, Tools MB 2021; Tusher et al. 2021). Figure 4.12 shows some main processes involved in co-metabolism.

Fig. 4.12
A schematic representation of the aerobic co-metabolic biotransformation process of 1,4 dioxane. The active bacterium breaks down the co-substrate, and it is soluble by monooxygenase, resulting in fast intermediate, co-metabolism, and slow intermediate metabolisms. M M O oxidizes the co-substrate.

Some main processes involved in co-metabolism (Mahendra et al. 2013)

Co-metabolic active bacteria such as R. rhodochrous ATCC 21198 can often break down co-substrates at lower concentrations below those attained by organisms that metabolize and thrive on them since co-substrate is not utilized as either primary carbon or energy sources (Kashimoto et al. 1989; Mahendra et al. 2013; US EPA 2014a, 2017; Arve 2015; Zhang et al. 2017, b; Johnson et al. 2020; Miao et al. 2020; Metabolism A, Cometabolism A, Biodegradation A, Tools MB 2021; Tusher et al. 2021). Numerous bacteria such as Mycobacterium austroafricanum JOB5 (Arve 2015; Zhang et al. 2017; Johnson et al. 2020; Tusher et al. 2021) can co-metabolize 1,4-dioxane in various conditions. These include model organisms that utilize well-characterized enzymes like soluble methane monooxygenase (sMMO) and grow on primary substrates like toluene or methane, and other 1,4-dioxane co-metabolizing strains reported to thrive on THF, ethane, and isobutane. Numerous bacterial monooxygenases have the capacity to oxidize numerous co-substrates at once; if fully studied and utilized, this may have a promising future in the remediation of environmental contaminants.

Hatzinger and colleagues assessed methane and ethane capacity to promote the aerobic co-metabolism that breaks down 1,4-dioxane in groundwater aquifers (Hatzinger et al. 2017). Ethane encouraged M. sphagni ENV482, which was isolated from a separate aquifer, to aerobically break down 1,4-dioxane (Hatzinger et al. 2017). According to this study, ethane, a common by-product of the biotic or abiotic reductive dechlorination of chlorinated ethanes and ethenes, may act as a substrate to speed up the breakdown of 1,4-dioxane in aquifers, especially in areas where these by-products combine with aerobic groundwater (Hatzinger et al. 2017). A similar study was conducted by Murnane and colleagues on co-metabolic transformation of 1,1,1-trichloroethane and 1,4-dioxane by pure cultures of R. rhodochrous (ATCC strain 21198) (Murnane et al. 2021a). Results indicated that the transformation of dioxane occurred without a lag phase for cells grown on 2-butanol, while an induction period of several hours was required for 1-butanol-grown cells (Murnane et al. 2021a). A similar observation was archived with activity-based labelling patterns for monooxygenase hydroxylase components and specific rates of tetrahydrofuran degradation (Murnane et al. 2021a). The study further reported lower rates of oxygen gas consumption in the reactors containing tetra-s-butylorthosilicate, which has benefits for in situ bioremediation. Indicating the structure of SRC is important when developing passive aerobic co-metabolic treatment systems.

4.4.3 Anaerobic Biotransformation of 1,4-Dioxane

Little metabolic or co-metabolic anaerobic 1,4-dioxane biodegradation has been observed thus far (Skinner et al. 2009; Göen et al. 2016; Guan et al. 2018). There was no evidence of 1,4-dioxane biodegradation in microcosm research employing samples of aquifer material from several 1,4-dioxane-impacted sites. On 1,4-dioxane, an iron-reducing bacterium did grow anaerobically with chemical processes (Arve 2015; Zhang et al. 2017; Hamid et al. 2020; Ramalingam and Cupples 2020a; Dang and Cupples 2021; Sengupta and Dhal 2021). Fe(III)-reducing facultative anaerobe S. oneidensis can produce hydroxyl radicals, which degrade 1,4-dioxane as indicated in Fig. 4.13 (Sekar and DiChristina 2014; Sekar et al. 2016).

Fig. 4.13
A schematic representation of the anaerobic biotransformation process of 1,4 dioxane. F e 3 + yields F e 2 + reversibly, which is further possessed by the anaerobic microbes to turn to the O H group, resulting in 1,4 dioxane. Finally, the chemical fenton reactions led to 1,4 dioxane degradation.

Facultative anaerobe transformation of 1,4-dioxane by S. oneidensis

S. oneidensis produces Fe(II) in anaerobic circumstances, which interacts chemically with H2O2 to yield hydroxyl radicals that can result in oxidative 1,4-dioxane breakdown (Sekar and DiChristina 2014; Sekar et al. 2016).

4.5 Roles of Soil Bacteria in the Biotransformation of 1,4-Dioxane

A cyclic molecule such as 1,4-dioxane with two ether moieties was thought to be relatively unmanageable to biodegradation, until a recent study that demonstrated that it may be efficiently biotransformed (Zhang et al. 2016, 2017; He et al. 2018). Dioxane-degrading bacteria are promising in situ bioremediation agents for cleaning up 1,4-dioxane-contaminated soils because it is less expensive than using advanced technologies. Several strains that may co-metabolically degrade dioxane have been isolated from wastewater treatment facilities or dioxane-affected areas (He et al. 2018). Numerous research emphasized the significance of monooxygenase enzymes in dioxane breakdown, and a gene cluster and sequence identified in P. dioxanivorans CB1190 that can aid in this process (i.e., thmADBC) (Mahendra and Alvarez-Cohen 2006; Kim et al. 2009; Sales et al. 2011). The gene cluster encodes dioxane monooxygenases, the dissolving di-iron (SDIMO) that initiates the breakdown of dioxane. The di-iron monooxygenase contains multiple bacterial enzyme components that can catalyze the oxidation of a variety of priority pollutants such as chlorinated solvents and aromatic hydrocarbons, among other contaminants, implying that they have inherent bioremediation potential (Notomista et al. 2003; Grostern et al. 2012).

A garden soil with unknown history of exposure to dioxane has shown a capability to degrade 500 mg/L dioxane as the sole source of carbon and energy to non-detectable levels within a week (He et al. 2018). Filtering through the aquifer material hinders the subsurface dispersion of inoculated bacteria, which is particularly problematic when the cultures aggregate. The archetypal dioxane degraders, which aggregate in suspension and include P. dioxanivorans (CB1190) and M. dioxanotrophicus (PH-06), are severely constrained by this (Grostern et al. 2012). Remarkably, the SDIMO consortia (A and B) enriched by He et al. (2018) from garden soil did not clump as much as the archetype, making them better candidates for in situ bioaugmentation at areas with inadequate indigenous dioxane biodegradation ability. Figure 4.14 shows findings indicating that there is a substantial growth of bacteria when exposed in the dioxane, while simultaneous decrease in dioxane concurrently occurs.

Fig. 4.14
A multi-line graph plots dioxane concentration and bacterial growth versus time in days. Control dioxane line begins at (0, 500) and ends at (7, 505). 4 lines for different dioxanes follow a decreasing trend and end at (7, 0). 3 lines for control and consortiums A and B follow an increasing trend.

Degradation of dioxane-enriched consortia and two archetypes (left) and concurrent growth of consortia (right). [Source: He et al. (2018)]

Tenfold serial dilution was adopted during conduction of experiments in Fig. 4.14 where ammonium mineral salts were used as growth medium with an initial dioxane concentration of 500 mg/L, while autoclaved bacteria were used as control.

4.6 Fate of 1,4-Dioxane Biotransformation Products

Numerous intermediates produced by the biotransformation of 1,4-dioxane in environmental matrices have the potential to interact with the environment and result in secondary pollution issues. Therefore, it is crucial to comprehend their fate and the effects of these items. Spectrometry, chromatography, nuclear magnetic resonance spectroscopy (NMR), gas chromatography-mass spectroscopy (GC-MS), high-pressure liquid chromatography (HPLC), and gas chromatography-flame ionization detection (GC-FID) are some techniques that can be used to quantify and qualitatively estimate the extent of changes in 1,4-dioxane (Kikani et al. 2022). The production of metabolites with are structurally distinct via redox mechanisms is the consequence of 1,4-dioxane transformation by biomaterials. There have been papers on microbial systems’ qualitative evaluation of 1,4-dioxane intermediates and metabolites. The intermediates produced by Cordyceps sinensis during the biodegradation of 1,4-dioxane are shown in Fig. 4.15 (Nakamiya et al. 2005; Mahendra et al. 2007; Kang and Doty 2014).

Fig. 4.15
An oxalic acid has a 2-carbon chain. Both are single bonded to O and double bonded to O H. Ethylene glycol has a 4-C chain. C1 and C4 are replaced by O H. C2 and C3 are replaced by C H 2. Glycolic acid has a 4-C chain. C1 and C4 are replaced by O H. C2 is double bonded to O. C3 is replaced by C H 2.

Some intermediate compounds produced by Cordyceps sinensis during biodegradation of 1,4-dioxane

Shaily and colleagues (Mahendra et al. 2007) described a process for dioxane oxidation by monooxygenase-expressing cells that does not result in the buildup of hazardous intermediate compounds in the environment. In this mechanism, dioxane is first transformed to 2-hydroxy-1,4-dioxane, which is then instantaneously oxidized to 2-hydroxyethoxyacetic acid. During a second monooxygenation step, 2-hydroxy-1,4-dioxane is further hydroxylated, yielding a mixture of dihydroxy-ethoxyacetic acids with a hydroxyl group in the ortho or para position. After the second ether bond is broken, small organic molecules such as ethylene glycol, glycolate, glyoxylate, and oxalate are progressively formed and mineralized to CO2 via common cellular metabolic mechanism.

4.7 Conclusion

The occurrence of 1,4-dioxane in the water, wastewater, soil, drinking water, and groundwater is globally reported. This chemical is classified as a carcinogenic compound, posing a threat to ecosystems upon its chronic exposure. Its occurrence in many consumer products as either an impurity or a solvent and its high miscibility to water lead to its widespread contamination, complicating its remediation methods. Bacteria such as Pseudonocardia dioxanivorans CB1190 through aerobic biotransformation, Rhodococcus rhodochrous ATCC 21198 through co-metabolic transformation, and Shewanella oneidensis through anaerobic transformation have shown a significant bioreduction of 1,4-dioxane concentration in the contaminated environments. The role of these bacteria is through either consumption of 1,4-dioxane as a source of energy, thus reducing its concentration, or transformation to different products that are harmless compared to parent molecule. In the future, relying of microbial remediation of 1,4-dioxane is expected to take a pace since nature does not leave behind unattended waste resulting in an ecological safety.