Fungal Biorem of PFAS

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

DOI: 10.1002/rem.

21550

RESEARCH ARTICLE

Fungal biotransformation of 6:2 fluorotelomer alcohol

Nancy Merino Meng Wang Rocio Ambrocio Kimberly Mak Ellen O'Connor
An Gao Elisabeth L. Hawley Rula A. Deeb Linda Y. Tseng Shaily Mahendra

Correspondence
Shaily Mahendra, Department of Civil and Abstract
Environmental Engineering, 5732J Boelter Fungal degradation of 6:2 fluorotelomer alcohol (6:2 FTOH, C6 F13 CH2 CH2 OH) by two wood-
Hall, University of California, Los Angeles, CA
decaying fungal strains and six fungal isolates from a site contaminated with per- and polyfluo-
90095.
Email: mahendra@seas.ucla.edu roalkyl substances (PFASs) was investigated. 6:2 FTOH is increasingly being used in FTOH-based
products, and previous reports on the microbial fate of 6:2 FTOH have focused on bacteria and
environmental microbial consortia. Prior to this study, one report demonstrated that the 6:2
FTOH biotransformation by the wood-decaying fungus, Phanerochaete chrysosporium, generated
more polyfluoroalkyl substances, such as 5:3 acid (F(CF2 )5 CH2 CH2 COOH), and diverted away
from producing the highly stable perfluorocarboxylic acids (PFCAs). Most of the fungi (Gloeophyl-
lum trabeum and isolates TW4-2, TW4-1, B79, and B76) examined in this study showed similar
degradation patterns, further demonstrating that fungi yield more 5:3 acid (up to 51 mol% of ini-
tial 6:2 FTOH dosed) relative to other metabolites (up to 12 mol% total PFCAs). However, medium
amendments can potentially improve 6:2 FTOH biotransformation rates and product profiles. The
six fungal isolates tolerated up to 100 or 1,000 milligrams per liter of perfluorooctanoic acid and
perfluorooctane sulfonic acid, and some isolates experienced increased growth with increasing
concentrations. This study proposes that fungal pathways must be considered for the biotransfor-
mation of potential PFAS precursors, such as 6:2 FTOH, and suggests the basis for selecting proper
microorganisms for remediation of fluoroalkyl-contaminated sites.

1 INTRODUCTION the ecosystem and account for up to 75 percent of the soil microbial
biomass (Ritz & Young, 2004), it is important to understand the various
6:2 Fluorotelomer alcohol (6:2 FTOH, C6 F13 CH2 CH2 OH) is part 6:2 FTOH fungal biotransformation pathways for predicting PFAS fate
of a larger group of compounds known as per- and polyfluoroalkyl in the environment and for devising remediation strategies.
substances (PFASs), which are widely used in consumer, industrial, In addition, 6:2 FTOH fungal biotransformation may be more favor-
and military applications due to their unique properties, including able for the remediation of 6:2 FTOH-based products compared to
resistance to water, oil, and stains (Baker & Rao, 1994). However, bacterial biotransformation pathways. The major metabolites of 6:2
long-chain perfluorocarboxylic acids (PFCAs), such as perfluorooc- FTOH biotransformation in soils, sediments, activated sludge, and
tanoic acid (PFOA), and their precursors, such as 8:2 FTOH, are being microbial consortia were PFCAs (e.g., perfluorobutanoic acid (PFBA)
phased out because of their toxicological and ecological impacts [F(CF2 )3 COOH], perflouropentanoic acid (PFPeA) [F(CF2 )4 COOH],
(Ding & Peijnenburg, 2013; Paul & Jones, 2009; U.S. Enviromental and perfluorohexanoic acid (PFHxA) [F(CF2 )5 COOH]) and x:3 acids,
Protection Agency, 2006). 6:2 FTOH is one of the shorter-chain such as 5:3 acid and 4:3 acid [F(CF2 )4 CH2 CH2 COOH] (Liu et al., 2010b;
substitutes (Ritter, 2010), but once released into the environment, it Zhao et al., 2013). For single bacterial strains, the major products were
can be transformed to PFCAs and other polyfluoroalkyl substances two transient, initial intermediates: 6:2 fluorotelomer unsaturated car-
by biological and physicochemical processes (Merino et al., 2016). boxylic acid (FTUCA) (F(CF2 )5 CF = CHCOOH), and 6:2 fluorotelomer
Biotransformation of 6:2 FTOH by bacteria and microbial consortia carboxylic acid (FTCA) (F(CF2 )6 CH2 COOH; Kim, Wang, & Chu, 2013;
has been previously reported (Butt, Muir, & Mabury, 2014). However, Kim, Wang, McDonald, & Chu, 2012). In contrast, fungal transforma-
one study examined the fungal biotransformation of 6:2 FTOH by the tion of 6:2 FTOH by P. chrysosporium led to higher production of 5:3 acid
wood-decaying fungus, Phanerochaete chrysosporium (Tseng, Wang, and 5:3 acid conjugates with minimal PFCA yields (Tseng et al., 2014).
Szostek, & Mahendra, 2014). Since fungi play an important role in Since polyfluoroalkyl substances do not contain a fully fluorinated

Remediation. 2018;28:59–70. wileyonlinelibrary.com/journal/rem 2018


c Wiley Periodicals, Inc. 59
60 MERINO ET AL .

carbon chain, these products can then be more easily transformed in zone (5–6 ft interval) with acetate liners. All equipment was rinsed
the environment by other processes. thoroughly first with tap water and then with deionized water before
While P. chrysosporium is the model white-rot basidiomycete for use. Samples were shipped and stored at 4 ◦ C and were used within
the degradation of a wide range of contaminants, the phylum Basid- one month of receiving the samples to isolate fungi capable of growing
iomycota accounts for about 34 percent of all described fungal species on high concentrations of PFOA and PFOS.
(Harms, Schlosser, & Wick, 2011). On the other hand, the largest fungal
group, Ascomycota, also includes species that can degrade pollutants, 2.3 Isolation of fungi
including toluene, aliphatic hydrocarbons, polycyclic aromatic hydro-
Culturable fungi were first enriched on nutrient-rich agar, followed
carbons (PAHs), pesticides, and chlorophenols (Harms et al., 2011).
by several transfers of morphologically different colonies in Kirk agar
Several studies have isolated soil fungi from contaminated sites to
(Tien & Kirk, 1988; Tseng et al., 2014) containing 10 milligrams (mg)
degrade organic chemicals, such as 𝛽-hexachlorocyclohexane (Ceci
PFOA or PFOS/L. Nutrient-rich agar consisted of either Yeast Pep-
et al., 2015), polyvinyl chloride (Ali et al., 2014), and dieldrin (Kataoka,
tone Dextrose (Becton Dickinson, Franklin Lakes, NJ), Malt Extract
Takagi, Kamei, Kiyota, & Sato, 2010). In order to better understand
(ME), or Potato Dextrose (Becton Dickinson, Franklin Lakes, NJ) agar.
the 6:2 FTOH fungal transformation pathways in the environment, it
ME agar contained (per L): 16 grams (g) glucose, 16 g ME, 1.8 g mal-
is important to examine various fungal taxonomic groups.
tose, 2 g peptone, 3.2 g yeast extract, 1 g asparagine, 2 g KH2 PO4 , 1 g
The objectives of this study were to determine the 6:2 FTOH bio-
MgSO4 ⋅7H2 O, 1 mg thiamine–HCl, and 15 g agar. Kirk agar was cho-
transformation potential of two fungal strains, Gloeophyllum trabeum
sen as the nutrient-defined medium to select for fungal isolates related
and Trametes versicolor, and six fungal isolates obtained from sites
to wood-decaying fungi, which are known to degrade contaminants,
historically contaminated with PFASs from the use of aqueous film-
such as pharmaceuticals, azo dyes, and explosives (Harms et al., 2011).
forming foam (AFFF). Additionally, this work aimed to identify whether
For soil samples, 10 g soil was immersed in 50 milliliters (mL) of filter
the six fungal isolates could tolerate exposure to PFOA and perfluo-
sterilized Phosphate Buffer Saline and wrist shaken for ∼30 minutes
rooctane sulfonic acid (PFOS).
(min) to resuspend soil-bound microorganisms. Afterwards, each sam-
ple was aseptically streaked onto nutrient-rich agar with a 10 micro-
liter (𝜇L) sterile inoculating loop. Groundwater samples were shaken
2 MATERIALS AND METHODS
and directly streaked onto nutrient-rich agar.

2.1 Chemicals and fungal pure cultures


The per- and polyfluoroalkyl chemicals described in this article are The objectives of this study
listed in Exhibit S1 with the chemical names, acronyms, and formulae.
The parent compound, 6:2 FTOH (99 percent purity) and standards
were to determine the 6:2
for the biotransformation products were provided by DuPont (Liu
et al., 2010b; Zhang et al., 2013). Internal standards 2H-perfluoro-[1,2-
FTOH biotransformation
13 C
2 ]-2-octenoic acid (M6:2 FTUCA,+98 percent; Wellington Labo- potential of two fungal
ratories, Ontario, Canada) and perfluoro-n-[1,2-13 C2 ] hexanoic acid
(MPFHxA, +98 percent; Wellington Laboratories, Ontario, Canada) strains, Gloeophyllum
were used for liquid chromatography-tandem mass spectrometry (LC-
MS/MS) quantitative analysis. Solvents for LC-MS/MS analysis were
trabeum and Trametes
HPLC grade or higher. All medium components had purities ≥98.0 per-
cent and were purchased from either Sigma-Aldrich (St. Louis, MO) or
versicolor, and six fungal
Thermo Fisher Scientific (Waltham, MA). isolates obtained from sites
G. trabeum was purchased from the American Type Culture Collec-
tion (ATCC 11539; Manassas, VA). Trametes versicolor (MAD697-R) was historically contaminated
generously provided by Professor Chad Jafvert of Purdue University.
with PFASs from the use of
2.2 Site description aqueous film-forming foam
Groundwater (4 liters [L] total) and soil samples (2-foot [ft] linear
core soil sample) were collected from two U.S. military bases by
(AFFF).
ARCADIS/Malcolm Pirnie Inc. Both sites utilized AFFF extensively
for fire training and other fire-related incidents. Detailed information, Triplicate nutrient-rich agar streak plates were prepared for each
including dissolved PFASs concentrations and other environmental groundwater and soil sample and incubated at 30◦ C for up to
parameters, are included in Exhibit S2. Groundwater was collected 4 weeks. Individual colonies were then selected and transferred to
from temporary wells and stored in polypropylene bottles with min- fresh nutrient-rich agar to isolate each morphologically different fun-
imal headspace. Soil samples were collected in the shallow vadose gal colony. All fungal isolates were then transferred to Kirk agar with
MERINO ET AL . 61

10 g glucose/L, followed by several transfers on Kirk agar without glu- tration was ≤18 percent, as measured by headspace oxygen analyzer
cose and with 10 to 100 mg PFOA or PFOS/L. All fungi were incu- (Model 902D, Quantek Instruments, Grafton, MA), all bottles were aer-
bated at 30◦ C to select for those that could tolerate these highly sta- ated for 3 min with filter sterilized air at ∼150 cubic centimeters per
ble PFASs. Afterwards, each fungal isolate was incubated in liquid Kirk minute flow rate.
medium with 10 g glucose/L and 10 to 100 mg PFOA or PFOS/L at A sterile group and a PFAS-free (matrix) control group were
30 ◦ C with 150 revolutions per minute (rpm) orbital shaking. Glu- prepared to account for abiotic effects and for background PFAS
cose was removed in future transfers to further isolate those fungi concentrations, respectively. The sterile control group consisted of
that could grow in liquid medium on high PFAS concentrations. Among resuspended fungi autoclaved at 121 ◦ C for 30 min instead of live
several hundred initial fungi isolated on nutrient-rich agar, six were fungi. The matrix control group was the same as the live group except
chosen (three from each site) for biotransformation studies and DNA that an equivalent volume (30 𝜇L) of 50 percent ethanol solution was
sequencing. added instead of the 6:2 FTOH stock solution. Each experimental
group consisted of triplicate samples sacrificed at each time point
(days 0, 7, 14, and 28).
2.4 6:2 FTOH biotransformation studies
All six fungal isolates were tested to biotransform 6:2 FTOH. Each iso-
2.5 Tolerance toward PFOA and PFOS
late was grown on Potato Dextrose or Yeast Peptone Dextrose agar for
1 week at 30 ◦ C before collecting the spores, as described previously All fungal isolate starter cultures were grown and collected in the same
(Tseng et al., 2014). Briefly, the spores on the agar plate were resus- manner as in Section 2.4. PFOA and PFOS tolerance studies were con-
pended in 10 mL sterile deionized water and filtered through sterile ducted in serum bottles (160 mL), crimp-sealed with butyl rubber stop-
glass wool. The spores were then transferred to 1 L liquid Kirk medium pers containing 10 mL total volume of resuspended fungi in fresh Kirk
(Tien & Kirk, 1988) containing 10 g glucose/L and grown at 30 ◦ C with medium without glucose. The experiments were initiated on day 0 by
150 rpm orbital shaking for 7 to 10 days and blended (ethanol and UV- adding 0.1 mL of PFOA and PFOS stock solutions (1 to 100 g/L made
sterilized Hamilton Beach Power Elite Multi-Function Blender) for a in 62 percent ethanol) to have a final concentration of 0 to 1,000 mg
total of 2 min and 20 seconds (sec; 30 sec blend with 5 sec break inter- PFOA and PFOS/L. Immediately after the addition of PFOA and PFOS,
vals) to obtain a starter culture. The fungal cultures, G. trabeum and each bottle was crimp-sealed with butyl rubber stoppers and two 0.22
T. versicolor, were grown in a similar manner except that Newcombe micron (𝜇m)-pore sterile nylon filters were inserted into the headspace
medium (Newcombe et al., 2002) and Tišma medium (Tišma, Znidarsic- to allow for aeration (measured headspace oxygen concentrations
Plazl, Vasic-Racki, & Zelic, 2012) were used. Before starting each bio- remained ≥20 percent). A sterile control group and a PFAS-free con-
transformation study, the fungal starter cultures were blended a final trol group were prepared as described in Section 2.4 to account for
time and washed thrice by centrifuging at 7,000 rpm for 5 to 20 min abiotic effects and for background PFAS concentrations, respectively.
before decanting the supernatant and resuspending the pellet in sterile The matrix control group was the same as the live group except that
deionized water. After the final washing step, the cell pellet was resus- an equivalent volume of 62 percent ethanol stock solution was added
pended in 1 L volume of fresh Kirk, Newcombe, or Tišma medium with instead of the PFOA and PFOS stock solution. Each experimental group
0 g glucose/L. consisted of duplicate samples sacrificed at each time point (days 0
Biotransformation studies were prepared as previously described and 14).
with a few modifications (Tseng et al., 2014). All experiments were con- At each time point, each sample was filtered through a pre-weighed
ducted in serum bottles (Wheaton, 160-mL), crimp-sealed with grey glass fiber GF/C Whatman filter and placed in an oven at 50 ◦ C for
butyl rubber stoppers, that contained 10 mL total volume of resus- 20 hours. The dried filter and fungal biomass was then weighed and the
pended fungi, 40 mg activated C18 powder from Maxi-Clean C18 car- pre-weighed mass of the filter was subtracted to obtain the dry weight
tridges (Grace Davison Discovery Sciences; Deerfield, IL), and addi- (mg) of the fungal biomass.
tional organic nutrient components. The C18 powder was added to
maintain 6:2 FTOH in the liquid phase since 6:2 FTOH is very volatile
2.6 Collection and processing of samples
(vapor pressure = 108 Pa at 35◦ C; Krusic et al., 2005) and was acti-
vated by eluting 10 mL acetonitrile through the cartridge, followed by Triplicate bottles were analyzed at each time point from the bio-
10 mL air, and let dry overnight. Two nutrient conditions were tested: transformation studies using a modified collection method described
(1) addition of 10 mg glucose (G), 0.25 mg yeast extract (Y), and 2 mg in Tseng et al. (2014). Briefly, prior to sample collection, all bot-
cellulose (C; +GYC) and (2) no added organic nutrients (–GYC). The tles were aerated with ∼1.9 L air (125 cubic centimeters per minute
experiments were initiated on day 0 by adding 30 microliters of a 1 g for 15 min) through the C18 cartridges inserted into the headspace
6:2 FTOH/L stock solution (final concentration 3 milligrams per liter to ensure that headspace 6:2 FTOH and volatile metabolites were
[mg/L]) made in 50 percent ethanol. Immediately after the addition of captured by the cartridges. Afterwards, the captured compounds
6:2 FTOH, each bottle was crimp-sealed with butyl rubber stoppers were eluted with 5 mL acetonitrile, followed by 5 mL air through
and two C18 cartridges were inserted into the headspace to capture each cartridge and collected inside the bottle (“first and cartridge
6:2 FTOH and any volatile intermediates (Tseng et al., 2014). The two extract”). The cartridges and the sample stopper were then removed,
C18 cartridges also allowed for aeration, and when oxygen concen- and the bottle was re-crimped with a fresh butyl rubber stopper
62 MERINO ET AL .

and incubated horizontally overnight at 50◦ C with 150 rpm shak- of (1) 64 ◦ C to 55 ◦ C (−1 ◦ C per cycle) for 55 sec (10 cycles) and
ing. After incubation, the bottles were stored at 4 ◦ C overnight to (2) 54 ◦ C for 55 sec (30 cycles), extending temperature of 72 ◦ C for
allow for settling of the fungal mycelium and any other large par- 1 min, final extension temperature 72 ◦ C for 10 min, hold tempera-
ticles. The supernatant was then filtered (0.22 𝜇m nylon) for LC- ture 4 ◦ C. The PCR product was verified using gel electrophoresis and
MS/MS quantitative analysis, and a “second extract” on the bottle cleaned using the UltraClean PCR Clean-Up Kit (Mo Bio Laboratories,
was conducted by adding 200 𝜇L 1 molar (M) NaOH and 5 mL ace- Carlsbad, CA). The cleaned PCR product was sequenced with ITS1-F
tonitrile. The second extract was then incubated and collected in the and ITS4-R primers (Schoch et al., 2012). Sequence data were subse-
same manner as the first extract. The extract from the sample stop- quently aligned against the NCBI database using standard nucleotide
per was obtained by incubating upright in a 20 mL scintillation vial BLAST and were submitted to NCBI under the accession numbers
overnight at 50 ◦ C with 150 rpm shaking in 5 mL acetonitrile (“stop- KU533844–KU533849.
per extract”), and was then combined with the second extract. All sam-
ples were stored at –20 ◦ C before analysis and at –40 ◦ C for long-
term storage. Additional extractions were conducted, as described in
The six fungal isolates
Discussion S1.
(TW1-3, TW4-2, TW4-1,
2.7 LC-MS/MS quantitative analysis B76, B78, and B79) were
LC-MS/MS was performed to determine the concentrations of 6:2 obtained from the culturable
FTOH and transformation products (Ruan et al., 2013; Zhang et al.,
2013). Briefly, 200 𝜇L filtered sample was spiked with 2 𝜇L inter- microbial communities found
nal standard containing (per mL) 1 nanogram (1,1-D; 1,2-13 C) 6:2
FTUCA and 1 nanogram (1,2-13 C) PFHxA. Quantitative analysis of 6:2
at two U.S. military bases
FTOH and transformation products were carried out with an Applied historically contaminated
Biosystems-MDS Sciex 4000 Qtrap equipped with a C8 column (ZOR-
BAX RX-C8, 5-𝜇m particle size, 2.1 × 150 millimeter) by Agilent (Santa with PFASs due to extensive
Clara, CA). The mass spectrometer was operated in negative electro-
spray ionization mode using scheduled multiple reaction monitoring. use of AFFF.
Samples (11 𝜇L) were injected via an autosampler and eluted with
0.15 percent acetic acid in HPLC grade water (solvent A) and 0.15 per-
3 RESULTS AND DISCUSSION
cent acetic acid in HPLC grade acetonitrile (solvent B) at a flow rate
of 400 microliters per minute. The gradient started with 90 percent
3.1 Identification of fungal isolates
A and 10 percent B (1.50 min), followed by 45 percent A and 55 per-
cent B (1.60 to 2.50 min), 20 percent A and 80 percent B (2.50 to The six fungal isolates (TW1-3, TW4-2, TW4-1, B76, B78, and B79)
8.00 min), 0 percent A and 100 percent B (8.10 to 10.10 min), and were obtained from the culturable microbial communities found at
ending with 90 percent A and 10 percent B (10.20 to 10.50 min). two U.S. military bases historically contaminated with PFASs due
Polytetrafluoroethylene (TeflonTM ) and other fluoropolymer materi- to extensive use of AFFF. Since these bases had been exposed to
als were avoided as much as possible to minimize background con- several decades of PFASs and other contaminants, such as chlorinated
tamination. More detailed information and ion transitions are listed in solvents and petroleum hydrocarbons, these fungal isolates were
Exhibit S3. part of a selected group of microorganisms that must have evolved
to tolerate or detoxify PFASs. In addition, these fungi were isolated
on nutrient-limited agar (Tien & Kirk, 1988) without glucose and high
2.8 Identification of fungal isolates
concentrations (mg/L) of PFOA or PFOS. The ITS region of each fungus
All fungal isolates were identified based on the DNA sequence of the was sequenced and compared against the NCBI database (Exhibit 1).
Internal Transcribed Spacer (ITS) region encompassing the 5.8S rRNA BLAST results indicated that four isolates, TW1-3, TW4-2, TW4-1, and
and 28S rRNA (Schoch et al., 2012). Total DNA (ZR Fungal/Bacterial B79, could be identified as Fusarium sp., Penicillium sp., or Aspergillus sp.
DNA MiniPrep kit, Zymo Research, Irvine, CA) was first extracted (Exhibit 1). Species in these genera have been shown to degrade a wide
from all fungal isolates grown in ME liquid medium at 30◦ C with range of contaminants (Exhibit 1). The majority of the results for B76
150 rpm orbital shaking for 4 days. Afterwards, polymerase chain and B78 were related to Fusarium sp. and Penicillium sp., respectively.
reaction (PCR) using the Promega GoTaq Kit (Promega, Madison, WI) However, B76 may also be related to Lachnum sp. (BLAST Ident 99
with ITS1-F and ITS4-R primers (Embong et al., 2008; Prewitt, Diehl, percent), and B78 may be related to Aspergillus sp. (BLAST Ident 100
McElroy, & Diehl, 2008) was conducted to amplify the ITS region. percent), Paecilomyces sp. (BLAST Ident 100 percent), or Hypocrea
The PCR conditions followed touchdown thermal cycling parameters sp. (BLAST Ident 100 percent). Species in Paecilomyces sp. have been
and contained 40 cycles: initial denaturation temperature 95 ◦C for shown to degrade benzene, toluene, ethylbenzene, xylene, and phenol
35 sec, melt temperature 95 ◦ C for 35 sec, annealing temperature (Estévez, Veiga, & Kennes, 2005; García-Peña, Ortiz, Hernández, &
MERINO ET AL . 63

EXHIBIT 1 BLAST identities of six fungal isolates

BLAST Closely related Biotransformation of other


Isolate Accession # ident genus contaminants Ref.
TW1-3 KU533849 100% Fusarium sp. Phenol, PAHs, lindane, plasticizers, Hidayat, Tachibana, and Itoh (2012);
B76a KU533845 99% dyes, polyamide-6 Luo et al. (2015); Sorkhoh,
Ghannoum, Ibrahim, Stretton, and
Radwan (1990)
B78b KU533847 100% Penicillium sp. PAHs, anthracene, endosulfan, Ceci et al. (2015); Hofrichter, Bublitz,
B79 KU533846 100% 𝛽-hexachlorocyclohexane, and Fritsche (1994); Jove et al.
halogenated phenols, imidazolium (2016); Mohsenzadeh, Rad, and
compounds, quaternary Akbari (2012); Song, Tian, Fan, and
ammonium compounds, dyes, He (2010)
polyethylene, polyamide-6,
aliphatic polyester resin Bionolle
TW4-2 KU533844 100% Aspergillus sp. PAHs, endosulfan, n-alkanes, Mohsenzadeh et al. (2012); Song
TW4-1 KU533848 99% dimethoate, dyes, et al. (2010)
carboxymethylchitosan-g-medium
chain length
polyhydroxyalkanoates
(mcl-PHA), polyamide-6, aliphatic
polyester resin Bionolle
a
B76 ITS rRNA region BLAST results contained one hit highly similar (Ident 99%) to Lachnum sp.
b
B78 ITS rRNA region BLAST results contained one hit each: Aspergillus sp (Ident 100%), Paecilomyces sp. (Ident 100%), and Hypocrea sp. (Ident 100%).

Revah, 2008; Wang et al., 2010), while species in Hypocrea sp. have Total recovery of 6:2 FTOH and metabolites are shown in Exhibits
been shown to degrade PAHs, pyrene, biopolymers, and azo dyes S5 to S13 and was around 38 to 133 mol% for all live fungal cultures.
(Baldrian et al., 2010; Gajera, Bambharolia, Hirpara, Patel, & Golakiya, The loss in recovery for certain fungi may indicate unknown metabo-
2015; Hong, Park, & Gadd, 2010). Although few studies have reported lites due to lack of authentic analytical standards or the formation
the biotransformation of organic pollutants by species in Lachnum of bound-residues to fungal, cellular organic components (Liu et al.,
sp., one report demonstrated that Lachnum spartinae could produce 2010b; Tseng et al., 2014; Wang et al., 2009; Zhang et al., 2013). The
laccase, an extracellular fungal enzyme known to biotransform many recovery of 6:2 FTOH in the sterile controls remained around 85 to
environmental pollutants (Desai & Nityanand, 2011) and may be able 116 mol% of the initial 6:2 FTOH dosed for all fungi at most time
to transform PFOA (Luo et al., 2015; Luo et al., 2017). These previous points. However, due to sacrificial sampling, instrumental variation, or
reports and the ability of the six fungal isolates to tolerate high PFOA lower efficiency in some C18 solid phase extraction (SPE) cartridges,
and PFOS concentrations suggest that these isolates may be able to the recovery for 4 out of 32 time points for sterile control bottles
transform fluorinated substances. Furthermore, previous research was below 70 percent or above 120 percent. Similar ranges in recov-
utilizing forest soil suggested that fungi played an essential role in ery for both the live and sterile control has been observed in previous
the initial rapid breakdown of 8:2 fluorotelomer stearate monoester studies (Kim et al., 2013; Kim et al., 2012; Liu et al., 2010a; Liu et al.,
(Dasu, Lee, Turco, & Nies, 2012). 2010b; Zhang et al., 2013; Zhao et al., 2013). Furthermore, the lack of
metabolite formation in the sterile control and the satisfactory recov-
ery for most time points indicate the integrity of the fungal system and
3.2 Biotransformation of 6:2 FTOH by fungi
extraction method.
Previously, it was shown that P. chrysosporium, a white-rot fungus, could From the biotransformation studies, the eight fungi could be
transform 6:2 FTOH to 11 metabolites with 5:3 acid (43 mol%) as ranked from least to highest residual 6:2 FTOH: G. trabeum (16 ±
the highest transformation product by day 14 (Tseng et al., 2014). 5 mol%) < TW4-2 (33 ± 4 mol%) < B76 (46 ± 16 mol%) < B78 (62 ±
Exhibit S4 depicts the fungal 6:2 FTOH transformation pathway. Since 6 mol%) < T. versicolor (63 ± 8 mol%) < B79 (65 ± 16 mol%) < TW4-
fungi play a large role in various ecosystems (Harms et al., 2011), this 1 (82 ± 5 mol%) < TW1-3 (111 ± 7 mol%). While the amount of 6:2
study further evaluated 6:2 FTOH biotransformation by other fungi, FTOH in TW1-3 cultures was higher than the initial 6:2 FTOH dose,
including a brown-rot fungus, G. trabeum; a soft-rot fungus, T. versicolor; increasing concentrations of 6:2 FTUCA (up to 8 mol% on day 28) were
and six fungal isolates. G. trabeum and T. versicolor were observed to observed over time. The lack of 6:2 FTOH transformation for some
transform 6:2 FTOH to 9 and 6 quantifiable transformation products, fungi could have also resulted from the partitioning of 6:2 FTOH into
respectively (Exhibits 2a, S4, S6, S7). All six fungal isolates were capable the headspace, where C18 SPE cartridges captured 6:2 FTOH, caus-
of transforming 6:2 FTOH to 5 to 9 quantifiable transformation prod- ing lower bioavailability. Since the extractions for the C18 SPE car-
ucts within 28 days at various removal molar yields (mol%; Exhibits 2a, tridge and the first liquid extraction were mixed, the recovery of 6:2
S4, S8–S13) under the conditions tested. The biotransformation path- FTOH into the headspace versus the liquid medium could not be dis-
ways and metabolite production profiles could change, for example, tinguished in this study. However, a previous report on 6:2 FTOH fun-
with different media formulations (Merino, Wang, Wang, & Mahendra, gal biotransformation demonstrated that after 28 days of incubation,
2018), initial 6:2 FTOH concentrations (Zhang, Merino, Wang, Ruan, & up to 46 mol% was observed in the headspace with only 2 to 16 mol%
Lu, 2017), or the concentrations of fungal inocula. remaining in the liquid culture (Tseng et al., 2014). In environmental
64 MERINO ET AL .

EXHIBIT 2 Transformation of 6:2 FTOH and major metabolites in nine fungal cultures

processes, 6:2 FTOH could still be bioavailable for transformation via a fungal culture, P. chrysosporium (Merino et al., 2018) and four bac-
microbial processes due to sorption to soil and sediment, slowing the terial cultures, Mycobacterium vaccae JOB5, Pseudomonas oleovorans,
rate of volatilization over time (Liu et al., 2010a; Zhao et al., 2013). Pseudomonas butanovora, and Pseudomonas fluorescens DSM 8341 (Kim
Previous studies have also shown that microbial processes in soil can et al., 2013).
transform volatile, fluorinated parent compounds, such as 6:2 FTOH, Among the eight fungi tested, five fungi, TW4-2, G. trabeum, TW4-
toward non-volatile intermediates and soil-bound residues (Liu, Lee, 1, B79, and B76, transformed 6:2 FTOH toward more abundant
Nies, Nakatsu, & Turco, 2007; Liu et al., 2010a; Liu et al., 2010b; Zhao 5:3 acid (9 to 51 mol%) relative to the other metabolites detected
et al., 2013). (Exhibits 2b, 3, S4–S14). 5:3 Acid is one of the major downstream
intermediates that has only been observed during microbial 6:2 FTOH
transformation (Liu et al., 2010a; Liu et al., 2010b; Tseng et al., 2014;
3.3 Polyfluoroalkyl acids were the major
Zhang et al., 2013; Zhao et al., 2013). The highest molar yields were
transformation products
observed in TW4-2 cultures (51 ± 16 mol%) after 16 days of incu-
Each fungi could also be ranked by the sum total of detected metabo- bation, which slightly increased to 62 ± 13 mol% after 63 days. The
lites at the end of the experiment from highest to lowest: TW4-2 (86 ± reduced rate of 5:3 acid formation after 16 days suggests that neces-
18 mol% total metabolites formed) > B79 (46 ± 5 mol%) > G. trabeum sary nutrients were depleted, and TW4-2 could not produce enough
(23 ± 4 mol%) > B76 (13 ± 4 mol%) > TW1-3 (9 ± 3 mol%) > B78 (9 ± reducing power to further transform the transient intermediate, 6:2
2 mol%) > TW4-1 (8 ± 4 mol%) > T. versicolor (6 ± 1 mol%; Exhibits 3, FTUCA (Exhibits S4 and S8). More research is needed to develop a
S6–S14). For fungal isolates B76 and B78, slightly more transformation substrate or nutrient supplement for long-term (>16 days) 6:2 FTOH
products (23 ± 10 mol% and 12 ± 3 mol%, respectively) were observed transformation.
when the medium was supplemented with GYC (Exhibits S10B–S11B). Other fungi, including G. trabeum and isolates B79 and B76, may
This suggests that B76 and B78 may need additional nutrients and also benefit from medium optimization to further transform 6:2 FTOH.
medium amendments to efficiently transform 6:2 FTOH. Indeed, the G. trabeum and isolates B79 and B76 had higher molar yields of 5:3
aerobic transformation of 6:2 FTOH requires a reducing power, such acid after 28 days of incubation (∼9 to 12 mol%) compared to total
as NAD(P)H (Wang, Buck, Szostek, Sulecki, & Wolstenholme, 2012), PFCAs production (<1 to 6.7 mol%; Exhibit 3). Preliminary experi-
suggesting the medium for isolates B76 and B78 may be lacking nec- ments using G. trabeum also observed similar 5:3 acid (20 ± 4 mol%)
essary nutrient components. Furthermore, variation in nutrients have and PFCAs (<2 mol%) yields (Exhibit S14). This suggests that these
previously affected the biotransformation pathway of 6:2 FTOH by fungi are potential candidates to produce enough reducing power to
MERINO ET AL . 65

EXHIBIT 3 Yields of 6:2 FTOH transformation products in pure cultures and environmental matrices

Initial Transient
intermediates products toward Stable products and intermediates
(mol%) PFCAs (mol%) (mol%)
𝚺 6:2 FTCA, 6:2 𝚺 5:2 ketone, 5:2 𝚺 PFPeA, PFHxA, 𝚺 5:3 acid, 4:3
FTUCA sFTOH PFBA acida Ref.
Gloeophyllum trabeum 3.4 <1 6.7 12.0 This study
Trametes versicolor <1 2.6 <1 2.0 This study
TW1-3 7.9 <1 <1 <1 This study
TW4-2 22.5 4.5 4.6 51.4 This study
TW4-1b 1.1 nd 1.4 4.5 This study
B79b 30.1 nd 4.6 9.4 This study
B79 4.7 nd 2.1 12.6 This study
B76 4.4 nd <1 9.2 This study
b
B78 4.4 nd 1.9 4.8 This study
Phanerochaete chrysosporiumc <1 10.7 5.9 33.5 Tseng et al. (2014)
Aerobic sedimentd nd 21.7 20.3 25.1 Zhao et al. (2013)
Aerobic soil 30.7 17.0 6.1 5.5 Liu et al. (2010b)
Activated sludgee nd 49.4 15.4 15.4 Zhao et al. (2013a)
Anaerobic digester sludgef 42.9 2.5 <1 21.0 Zhang et al. (2013)
Mycobacterium vaccae JOB5g 49.0 21.1 1.4 3.8 Kim et al. (2013)
h
Pseudomonas oleovorans 26.9 42.0 2.7 2.1 Kim et al. (2013)
Pseudomonas oleovoransi 26.9 35.0 1.7 1.7 Kim et al. (2013)
Pseudomonas butanovoraj 18.6 22.0 1.5 7.3 Kim et al. (2013)
Pseudomonas fluorescens 50.0 52.0 1.9 5.7 Kim et al. (2013)
DSM 8341k
nd = not detected.
a 5:3 acid and 4:3 acid can be further transformed via the one-carbon removal pathway37 or 5:3 acid conjugation7 .
b Supplemented with (per 10 mL) 0.25 mg yeast extract, 2 mg cellulose, and 10 mg glucose.
c Modified Kirk medium without glucose and supplemented (per 10 mL) with 100 mg lignocellulosic powder, 0.25 mg yeast extract,

2 mg cellulose, or 10 mg glucose.
d Day 100 data in aerobic river sediment.
e Day 60 data in activated sludge.
f Day 90 data in anaerobic digester sludge.
g
Supplemented with 1-butanol+formate.
h
Supplemented with n-octane+formate.
i
Supplemented with n-octane+dicyclopropylketone.
j
Supplemented with 1-butanol+lactate.
k
Supplemented with sodium chloroacetate+yeast extract+format.

transform 6:2 FTOH toward 5:3 acid, but supplemental substrates may gation with cellular organic compounds (Tseng et al., 2014). The one-
be needed to increase 6:2 FTOH biotransformation. For example, the carbon removal pathway is dominant in bacterial transformation of 6:2
amount of the initial intermediate 6:2 FTUCA observed on day 28 FTOH (Kim et al., 2013) and has also been observed for environmen-
for B79 cultures increased from 4.7 ± 1.2 to 30.1 ± 1.3 mol% after tal microbial consortia (Liu et al., 2010b; Zhao et al., 2013). However,
the removal of GYC, suggesting that GYC has some negative impact since two of the major products of the one-carbon removal pathway,
on 6:2 FTOH biotransformation by B79 cultures. However, the same 4:3 acid and 3:3 acid, were not observed in any of the fungal systems
nutrients were previously shown to improve the biotransformation in this study, it is unlikely that this pathway played a major role in the
of 6:2 FTOH with P. chrysosporium (Merino et al., 2018; Tseng et al., fungal transformation of 6:2 FTOH. Instead, 5:3 acid conjugation may
2014). have occurred, especially for those fungi, such as G. trabeum and B76,
Unknown transformation metabolites were likely produced by with more than 60 mol% loss in total recovery of the initial 6:2 FTOH
some fungi tested since the total recovery of 6:2 FTOH and metabo- dosed. Similar losses were observed in preliminary experiments using
lites could not be achieved. Additionally, when 3 mg 5:3 acid/L was G. trabeum (Exhibit S14). The 5:3 acid conjugation pathway resulted in
dosed as the parent compound, three fungal isolates, TW4-2, TW4- the production of three new metabolites when 6:2 FTOH was biotrans-
1, and B79, may have transformed 5:3 acid to unquantifiable metabo- formed by P. chrysosporium: 5:3 THPA, 5:3 TKE, and 5:3 TKHPA (Tseng
lites (Exhibit S4). Further transformation of 5:3 acid could occur via et al., 2014). Due to a lack of authentic analytical standards, these con-
one-carbon removal pathways (Wang et al., 2012) or 5:3 acid conju- jugates could not be quantified.
66 MERINO ET AL .

Other unknown fungal metabolites may have also occurred as some Kim et al., 2012), which have the potential to be further transformed
of the fungi investigated in this study are likely to utilize different to PFCAs by other microbial processes (Liu et al., 2010b). The effec-
biochemical pathways compared to P. chrysosporium. While G. trabeum tive transformation of 6:2 FTOH by the fungal isolate TW4-2 and G.
and T. versicolor belong to the same phylum (Basidiomycota) as P. trabeum (Exhibit 2a) indicates that other fungi besides P. chrysosporium
chrysosporium, the six fungal isolates belong to the phylum Ascomy- are capable of oxidizing and reducing 6:2 FTOH. Furthermore, the 5:3
cota. Furthermore, G. trabeum, T. versicolor, P. chrysosporium, Fusarium acid yield for both isolate TW4-2 and P. chrysosporium was higher than
sp., Penicillium sp, and Aspergillus sp. (Exhibit 1) employ different that observed for environmental matrices (Liu et al., 2010b; Zhang
strategies for the biotransformation of organic pollutants and produce et al., 2013; Zhao et al., 2013), indicating these two fungi could employ
a different suite of intracellular and extracellular enzymes, including similar biotransformative mechanisms as entire microbial consortia.
cytochrome P450 monooxygenases and peroxidases (Harms et al., However, for the other fungal cultures in this study, medium amend-
2011). The diversity of P450s in fungi are vast, with at least 150 ments may be needed for enhanced 6:2 FTOH degradation toward
P450 genes identified in P. chrysosporium (Yadav, Doddapaneni, & more degradable polyfluoroalkyl acids, preventing the accumulation
Subramanian, 2006) and 250 P450 gene candidates in a brown-rot of PFCAs. The medium used in this study may have been less effec-
fungi, Postia placenta (Ide, Ichinose, & Wariishi, 2012). Furthermore, tive for the production of the necessary enzymes and reducing power
G. trabeum is capable of producing a hydroquinone-driven Fenton to degrade 6:2 FTOH. For example, replacing nitrilotriacetic acid with
system, generating hydroxyl radicals and recycling Fe2+ /Fe3+ at ethylenediaminetetraacetic acid as the trace elements chelator might
low pH conditions (Hyde & Wood, 1997; Jensen, Houtman, Ryan, & enhance 6:2 FTOH biotransformation and increase 5:3 acid yields
Hammel, 2001). While the Fenton reaction produced by G. trabeum (Merino et al., 2018).
is not completely understood, these reactions may have influenced Compared to bacterial cultures (Kim et al., 2013), the fungal cul-
the transformation of 6:2 FTOH in this study as FeSO4 ∙7H2 O (25 𝜇M tures were much slower to oxidize 6:2 FTOH to the initial intermedi-
final concentration) was added to the medium (pH 4.5), and Fenton ate, 6:2 FTCA. Within three days, P. fluorescens DSM 8431 transformed
reaction systems have already been extensively utilized to degrade 64 mol% 6:2 FTOH, with the major products as 6:2 FTCA (55 mol%),
perfluoroalkyl acids (Merino et al., 2016). 6:2 FTUCA (8.9 mol%), when sodium chloroacetate, yeast extract, and
The remaining three fungal cultures (fungal isolates TW1-3 and B78 formate were added (Kim et al., 2013). P. oleovorans also transformed
and T. versicolor) did not transform 6:2 FTOH toward 5:3 acid under the at least 95 mol% 6:2 FTOH within 0.5 day to 6:2 FTCA (6.5 mol%) and
conditions tested. Fungal isolate TW1-3 could not transform 6:2 FTOH 6:2 FTUCA (25 mol%) when n-octane and dicyclopropylketone were
beyond one of the initial metabolites, 6:2 FTUCA (7.9 ± 1.4 mol%), over added (Kim et al., 2013). Similar results were observed for P. butanovora
28 days of incubation (Exhibit 3). This is likely due to a lack of reduc- and M. vaccae JOB5 (Kim et al., 2013). In contrast, fungi likely need at
ing power (Liu et al., 2010b; Wang et al., 2012) and may be rectified least 7 to 14 days to achieve similar results when accounting for only
by amending the medium (Kim et al., 2013; Merino et al., 2018). T. ver- the bioavailable portion of 6:2 FTOH (Merino et al., 2018; Tseng et al.,
sicolor and fungal isolate B78 also minimally transformed 6:2 FTOH, 2014). Since the microbial consortia from soil, sediment, and sludge
resulting in similar molar yields of 5:3 acid (∼2 and ∼5 mol%, respec- achieved high molar yields of 5:3 acid (Liu et al., 2010b; Zhang et al.,
tively) to the sum of 5:2 sFTOH and 5:2 ketone (∼3 and ∼4 mol%, 2013; Zhao et al., 2013), similar consortia made up of selected fun-
respectively), which are the direct precursors of the PFCAs, PFPeA, gal and bacterial cultures could be employed to efficiently degrade
and PFHxA (Exhibit S4; Liu et al., 2010b). However, production of 6:2 FTOH and similar PFASs in remediation systems. These systems
lower amounts of PFCAs and more transformation of 6:2 FTOH toward may utilize bacteria, such as P. fluorescens DSM 8431, P. oleovorans,
5:3 acid and other polyfluoroalkyl substances may be achieved by and P. butanovora, to quickly transform 6:2 FTOH to 6:2 FTUCA. 6:2
altering the medium. For example, the addition of 2,6-dimethoxy-1,4- FTUCA could then be further degraded by fungal degradation path-
benzoquinone and chelated ferric ion (Fe3+ -oxalic acid) may improve ways toward 5:3 acid and 5:3 acid conjugates (Tseng et al., 2014) or
6:2 FTOH transformation by T. versicolor due to the production of by one-carbon removal pathways to form x:3 acids (Wang et al., 2012).
hydroxyl radicals (Harms et al., 2011; Marco-Urrea, Aranda, Caminal, Although the toxicity of polyfluoroalkyl acids, such as x:3 acids and
& Guillen, 2009; Vilaplana, García, Caminal, Guillén, & Sarrà, 2012). 5:3 acid conjugates, is still unknown, these substances have the poten-
tial to be further biotransformed in the environment toward less toxic
compounds, unlike PFCAs that are environmentally persistent. Indeed,
3.4 Comparison of fungal cultures with other
more studies are needed to identify and describe the toxicological
microorganisms
properties of known and unknown polyfluoroalkyl acids before engi-
Although 6:2 FTOH biotransformation by most fungi tested in this neering biological remediation systems.
study was low compared to previous reports using pure cultures and
environmental matrices (Exhibit 3; Butt et al., 2014; Tseng et al., 2014),
3.5 Tolerance toward PFOA and PFOS
the amount of 5:3 acid observed was relatively higher than the produc-
tion of PFCAs and transient intermediates after 28 days of incubation. Since PFCAs are likely to be found with polyfluoroalkyl substances in
In comparison, three Pseudomonas strains and M. vaccae JOB5 accu- consumer products and industrial applications (Moody & Field, 1999)
mulated either the initial intermediates, 6:2 FTCA and 6:2 FTUCA, or and are also transformation end products in the aerobic degrada-
the transient products, 5:2 ketone and 5:2 sFTOH (Kim et al., 2013; tion of polyfluoroalkyl substances (Butt et al., 2014), tolerance toward
MERINO ET AL . 67

were negatively impacted by perfluoroalkyl acids (Weathers, Harding-


Marjanovic, Higgins, Alvarez-Cohen, & Sharp, 2016; Weathers, Hig-
gins, & Sharp, 2015), and the bacterial community in aerobic river
sediment was observed to decrease in diversity with PFOA expo-
sure (>100 nanograms per gram dry weight in sediment; Sun, Wang,
Peng, Wang, & Lu, 2016). Hydrogenotrophic methanogens from anaer-
obic sludge were also inhibited by 500 mg PFOS/L, but other PFASs
tested, such as PFBS, PFPrA, and PFPeA, did not inhibit methanogene-
sis (Ochoa-Herrera, Field, Luna-Velasco, & Sierra-Alvarez, 2016).
The increased growth experienced by most fungal isolates up to
100 or 1,000 mg PFOA and PFOS/L may be attributed to increased
oxygen transfer. Other perfluorinated compounds have been shown
to improve the solubility of respiratory gases, including perfluo-
rotributylamine, perfluorodecalin, perfluorotripropylamine, and
perfluoro-octyl bromide (Lowe, Davey, & Power, 1998; Mitsuno,
Ohyanagi, & Naito, 1982). These compounds have also previously been
used to enhance the growth of prokaryotic and eukaryotic cell culture
systems (Lowe et al., 1998), including P. chrysosporium batch cultures
(Ntwampe, Williams, & Sheldon, 2010). Due to enhanced oxygen trans-
fer, improved growth and bioproductivity was sustained over longer
periods compared to cultures without perfluorinated compounds
(Lowe et al., 1998; Ntwampe et al., 2010). Perfluorinated compounds
have also been shown to impart other positive side effects, such as
decreased mechanical damage to biomass, increased nutrient uptake,
and increased enzyme productivity (Lowe et al., 1998; Ntwampe et al.,
2010). Furthermore, polytetrafluoroethylene (TeflonTM ) membranes
have also been used to oxygenate animal-hybridoma cells (Schneider,
Reymond, Marison, & Vonstockar, 1995) and plant cells (Lowe et al.,
EXHIBIT 4 Effect of PFOA and PFOS on growth of fungal isolates 1998).

PFOA and PFOS by the six fungal isolates was examined. Fungi were
exposed to increasing concentrations of PFOA and PFOS (0, 10, 100, 4 CONCLUSIONS
and 1,000 mg/L) to determine the approximate concentration that
would affect fungal growth over 14 days of incubation (Exhibit 4). Since 6:2 FTOH is increasingly being used to manufacture FTOH-based
The isolates exhibited different growth behaviors based on their origi- products, it is important to understand the variety of microbial pro-
nal location. Those from site TW experienced increasing growth with cesses that can transform 6:2 FTOH. This study examined the fun-
increasing concentrations of PFOA and PFOS, with maximum dry gal biotransformation of 6:2 FTOH by three well-characterized fungal
weight observed at 1,000 mg PFOA and PFOS/L. In contrast, those cultures, P. chrysosporium, G. trabeum, and T. versicolor, and six fungal
from site B were negatively impacted by 1,000 mg PFOA and PFOS/L, isolates. Among the nine fungal cultures, TW4-2 isolate readily bio-
and only B78 and B79 were positively affected by 100 mg PFOA transformed 6:2 FTOH and produced the highest yields of 5:3 acid.
and PFOS/L. Generally, growth of all fungal isolates were not neg- Other fungal cultures transformed 6:2 FTOH at different removal
atively impacted by the lowest concentration dosed (10 mg/L), indi- yields, and most cultures formed higher amounts of polyfluoroalkyl
cating lower concentrations of PFOA, PFOS, and similar PFASs found substances, such as 5:3 acid, rather than terminal PFCAs. While more
in the environment (Moody & Field, 1999) would not impact growth studies are needed to characterize the environmental conditions to
of indigenous or bioaugmented fungi. Although all six fungal isolates improve 6:2 FTOH transformation rates and product profiles, the
were unable to biotransform PFOA (Exhibits S15 and S16), tolerance diversion of the 6:2 FTOH biotransformation pathway away from ter-
toward high concentrations of PFOA, PFOS, and similar PFASs sug- minal PFCAs and toward PFAAs by fungal cultures suggests that fungi
gests that these fungi are potential candidates for PFASs remediation. possess favorable mechanisms for the biotransformation of PFCA pre-
Furthermore, for sites contaminated with PFASs and metals, such as cursors. Since PFCAs are still likely to be produced during aerobic 6:2
chromium, most fungal isolates tested in this study were capable of FTOH transformation processes and may be present at PFAS contam-
sorbing metals (Exhibit S17). However, the effects of perfluoroalkyl inated sites, the six fungal isolates were able to tolerate high con-
substances on fungal stress responses and the fungal transformation centrations of PFOA and PFOS, further demonstrating that fungi are
of 6:2 FTOH and related compounds has not yet been examined. Com- potential candidates for PFAS remediation strategies. These results
paratively, bacteria (e.g., Dehalococcoides and Rhodococcus jostii RHA1) provide additional information on the role of bacteria and fungi on the
68 MERINO ET AL .

fate and transport of PFASs and could also be employed for developing com/track/pdf/10.1186/1471-2415-8-7?site=bmcophthalmol.
treatment trains involving biostimulation approaches or bioaugmenta- biomedcentral.com

tion cocultures containing fungi and bacteria to efficiently degrade 6:2 Estévez, E., Veiga, M. C., & Kennes, C. (2005). Biodegradation of toluene by
the new fungal isolates Paecilomyces variotii and Exophiala oligosperma.
FTOH and other PFASs in waste streams and at contaminated sites.
Journal of Industrial Microbiology and Biotechnology, 32(1), 33–37.
Gajera, H. P., Bambharolia, R. P., Hirpara, D. G., Patel, S. V., & Golakiya, B. A.
ACKNOWLEDGMENTS
(2015). Molecular identification and characterization of novel Hypocrea
This work was supported by the Air Force Civil Engineer Center con- koningii associated with azo dyes decolorization and biodegradation of
tract #FA8903-11-C-8009 and performed in a renovated collabora- textile dye effluents. Process Safety and Environmental Protection, 98,
406–416.
tory funded by National Science Foundation Grant #0963183, which
was awarded under the American Recovery and Reinvestment Act García-Peña, I., Ortiz, I., Hernández, S., & Revah, S. (2008). Biofiltration of
BTEX by the fungus Paecilomyces variotii. International Biodeterioration &
of 2009 (ARRA). The LC-MS/MS (UCLA Molecular Instrumentation
Biodegradation, 62(4), 442–447.
Center) was funded by the NSF Grant #S10RR024605. N. Merino
Harms, H., Schlosser, D., & Wick, L. Y. (2011). Untapped potential: Exploiting
received U.S. Environmental Protection Agency Science to Achieve fungi in bioremediation of hazardous chemicals. Nature Reviews Microbi-
Results (EPA-STAR) Fellowship, the UCLA Dissertation Year Fellow- ology, 9(3), 177–192.
ship, SWE Scholarships, and the Earth-Life Science Institute Origin of Hidayat, A., Tachibana, S., & Itoh, K. (2012). Determination of chrysene
Life (EON) Postdoctoral Fellowship, which is supported by a grant from degradation under saline conditions by Fusarium sp F092, a fungus
the John Templeton Foundation. The opinions expressed in this pub- screened from nature. Fungal Biology, 116(6), 706–714.

lication are those of the author(s) and do not necessarily reflect the Hofrichter, M., Bublitz, F., & Fritsche, W. (1994). Unspecific degradation of
halogenated phenols by the soil fungus Penicillium frequentans Bi 7/2.
views of the John Templeton Foundation. E. O'Connor received UCLA
Journal of Basic Microbiology, 34(3), 163–172.
Eugene V. Cota-Robles Fellowship and UCLA Competitive Edge Grant
Hong, J. W., Park, J. Y., & Gadd, G. M. (2010). Pyrene degradation and
(funded by a National Science Foundation Alliance for Graduate Educa-
copper and zinc uptake by Fusarium solani and Hypocrea lixii isolated
tion and the Professoriate grant). S. Mahendra received UCLA Hellman from petrol station soil. Journal of Applied Microbiology, 108(6), 2030–
Fellowship, the DuPont Young Professor Award, and the NSF CAREER 2040.
Award. We also thank Dr. Ning Wang for providing advice on experi- Hyde, S. M., & Wood, P. M. (1997). A mechanism for production of hydroxyl
mental design, analytical measurements, and manuscript preparation. radicals by the brown-rot fungus Coniophora puteana: Fe(III) reduction
by cellobiose dehydrogenase and Fe(II) oxidation at a distance from the
hyphae. Microbiology-UK, 143, 259–266.
Ide, M., Ichinose, H., & Wariishi, H. (2012). Molecular identification and
REFERENCES functional characterization of cytochrome P450 monooxygenases from
Ali, M. I., Ahmed, S., Robson, G., Javed, I., Ali, N., Atiq, N., & Hameed, A. the brown-rot basidiomycete Postia placenta. Archives of Microbiology,
(2014). Isolation and molecular characterization of polyvinyl chloride 194(4), 243–253.
(PVC) plastic degrading fungal isolates. Journal of Basic Microbiology,
Jensen, K. A., Houtman, C. J., Ryan, Z. C., & Hammel, K. E. (2001). Path-
54(1), 18–27.
ways for extracellular fenton chemistry in the brown rot basidiomycete
Baker, B. E., & Rao, N. S. (1994). Textile finishes and fluorosurfactants. In B. E. Gloeophyllum trabeum. Applied and Environmental Microbiology, 67(6),
Banks, B. E. Smart, & J. C. Tatlow (Eds.), Organofluorine chemistry: Princi- 2705–2711.
ples and commercial applications (pp. 321–338). New York, NY: Springer.
Jove, P., Olivella, M. A., Camarero, S., Caixach, J., Planas, C., Cano, L., & De Las
Baldrian, P., Voříšková, J., Dobiášová, P., Merhautová, V., Lisá, L., & Valášková, Heras, F. X. (2016). Fungal biodegradation of anthracene-polluted cork:
V. (2010). Production of extracellular enzymes and degradation of A comparative study. Journal of Environmental Science and Health, Part A,
biopolymers by saprotrophic microfungi from the upper layers of forest 51(1), 70–77.
soil. Plant and Soil, 338(1-2), 111–125.
Kataoka, R., Takagi, K., Kamei, I., Kiyota, H., & Sato, Y. (2010). Biodegradation
Butt, C. M., Muir, D. C. G., & Mabury, S. A. (2014). Biotransformation of dieldrin by a soil fungus isolated from a soil with annual endosulfan
pathways of fluorotelomer-based polyfluoroalkyl substances: A review. applications. Environmental Science & Technology, 44(16), 6343–6349.
Environmental Toxicology and Chemistry, 33(2), 243–267.
Kim, M. H., Wang, N., & Chu, K. H. (2013). 6:2 Fluorotelomer alcohol (6:2
Ceci, A., Pierro, L., Riccardi, C., Pinzari, F., Maggi, O., Persiani, A. M., … Papini, FTOH) biodegradation by multiple microbial species under different
M. (2015). Biotransformation of 𝛽-hexachlorocyclohexane by the sapro- physiological conditions. Applied Microbiology and Biotechnology, 98(4),
trophic soil fungus Penicillium griseofulvum. Chemosphere, 137, 101–107. 1831–1840.
Dasu, K., Lee, L. S., Turco, R. F., & Nies, L. F. (2012). Aerobic biodegradation Kim, M. H., Wang, N., McDonald, T., & Chu, K.-H. (2012). Biodeflu-
of 8:2 fluorotelomer stearate monoester and 8:2 fluorotelomer citrate orination and biotransformation of fluorotelomer alcohols by two
triester in forest soil. Chemosphere, 91(3), 399–405. alkane-degrading Pseudomonas strains. Biotechnology and Bioengineering,
Desai, S. S., & Nityanand, C. (2011). Microbial laccases and their applica- 109(12), 3041–3048.
tions: A review. Asian Journal of Biotechnology, 3(2), 98–124. Krusic, P. J., Marchione, A. A., Davidson, F., Kaiser, M. A., Kao, C.-P. C.,
Ding, G., & Peijnenburg, W. J. G. M. (2013). Physicochemical properties and Richardson, R. E., … Buck, R. C. (2005). Vapor pressure and intramolec-
aquatic toxicity of poly- and perfluorinated compounds. Critical Reviews ular hydrogen bonding in fluorotelomer alcohols. Journal of Physical
in Environmental Science and Technology, 43(6), 598–678. Chemistry A, 109(28), 6232–6241.

Embong, Z., Hitam, W. H. W., Yean, C. Y., Rashid, N. H. A., Kamarudin, B., Liu, J., Lee, L. S., Nies, L. F., Nakatsu, C. H., & Turco, R. F. (2007). Biotransfor-
Abidin, S. K. Z., … Ravichandran, M. (2008). Specific detection of fungal mation of 8:2 fluorotelomer alcohol in soil and by soil bacteria isolates.
pathogens by 18S rRNA gene PCR in microbial keratitis. BMC Ophthal- Environmental Science & Technology, 41(23), 8024–8030.
mology, 8. Retrieved from https://bmcophthalmol.biomedcentral.
MERINO ET AL . 69

Liu, J., Wang, N., Buck, R. C., Wolstenholme, B. W., Folsom, P. W., Sulecki, Ruan, T., Szostek, B., Folsom, P., Wolstenholme, B. W., Liu, R., Liu, J.,
L. M., & Bellin, C. A. (2010a). Aerobic biodegradation of [14 C] 6:2 flu- … Buck, R. C. (2013). Aerobic soil biotransformation of 6:2 fluo-
orotelomer alcohol in a flow-through soil incubation system. Chemo- rotelomer iodide. Environmental Science & Technology, 47(20), 11504–
sphere, 80(7), 716–723. 11511.
Liu, J., Wang, N., Szostek, B., Buck, R. C., Panciroli, P. K., Folsom, P. W., … Schneider, M., Reymond, F., Marison, I. W., & Vonstockar, U. (1995). Bubble-
Bellin, C. A. (2010b). 6-2 Fluorotelomer alcohol aerobic biodegradation free oxygenation by means of hydrophobic porous membranes. Enzyme
in soil and mixed bacterial culture. Chemosphere, 78(4), 437–444. and Microbial Technology, 17(9), 839–847.
Lowe, K. C., Davey, M. R., & Power, J. B. (1998). Perfluorochemicals: Their Schoch, C. L., Seifert, K. A., Huhndorf, S., Robert, V., Spouge, J. L., Levesque, C.
applications and benefits to cell culture. Trends in Biotechnology, 16(6), A., … Schindel, D. (2012). Nuclear ribosomal internal transcribed spacer
272–277. (ITS) region as a universal DNA barcode marker for Fungi. Proceedings of
Luo, Q., Lu, J., Zhang, H., Wang, Z., Feng, M., Chiang, S.-Y. D., … Huang, Q. the National Academy of Sciences, 109(16), 6241–6246.
(2015). Laccase-catalyzed degradation of perfluorooctanoic acid. Envi- Song, F. Q., Tian, X. J., Fan, X. X., & He, X. B. (2010). Decomposing ability
ronmental Science & Technology Letters, 2(7), 198–203. of filamentous fungi on litter is involved in a subtropical mixed forest.
Luo, Q., Wang, Z., Feng, M., Chiang, D., Woodward, D., Liang, S., … Huang, Q. Mycologia, 102(1), 20–26.
(2017). Factors controlling the rate of perfluorooctanoic acid degrada- Sorkhoh, N. A., Ghannoum, M. A., Ibrahim, A. S., Stretton, R. J., & Radwan,
tion in laccase-mediator systems: The impact of metal ions. Environmen- S. S. (1990). Crude-oil and hydrocarbon-degrading strains of Rhodococ-
tal Pollution, 224, 649–657. cus rhodochrous isolated from soil and marine environments in Kuwait.
Marco-Urrea, E., Aranda, E., Caminal, G., & Guillen, F. (2009). Induction of Environmental Pollution, 65(1), 1–17.
hydroxyl radical production in Trametes versicolor to degrade recalci- Sun, Y., Wang, T., Peng, X., Wang, P., & Lu, Y. (2016). Bacterial community
trant chlorinated hydrocarbons. Bioresource Technology, 100(23), 5757– compositions in sediment polluted by perfluoroalkyl acids (PFAAs) using
5762. Illumina high-throughput sequencing. Environmental Science and Pollu-
Merino, N., Qu, Y., Deeb, R. A., Hawley, E. L., Hoffmann, M. R., & Mahen- tion Research, 23(11), 10556–10565.
dra, S. (2016). Degradation and removal methods for perfluoroalkyl and Tien, M., & Kirk, T. K. (1988). Lignin peroxidase of Phanerochaete chrysospo-
polyfluoroalkyl substances in water. Environmental Engineering Science, rium. Methods in Enzymology, 161, 238–249.
33(9), 615–649. Tišma, M., Znidarsic-Plazl, P., Vasic-Racki, D., & Zelic, B. (2012). Optimiza-
Merino, N., Wang, N., Wang, M., & Mahendra, S. (2018, In Review). Roles tion of laccase production by Trametes versicolor cultivated on industrial
of various nutrient components on the biotransformation of 6:2 fluo- waste. Applied Biochemistry and Biotechnology, 166(1), 36–46.
rotelomer alcohol (6:2 FTOH) on a wood-decaying fungi. Tseng, N., Wang, N., Szostek, B., & Mahendra, S. (2014). Biotransformation
Mitsuno, T., Ohyanagi, H., & Naito, R. (1982). Clinical-studies of a perfluoro- of 6:2 fluorotelomer alcohol (6:2 FTOH) by a wood-rotting fungus. Envi-
chemical whole-blood substitute (Fluosol-Da): Summary of 186 cases. ronmental Science & Technology, 48(7), 4012–4020.
Annals of Surgery, 195(1), 60–69. U.S. Enviromental Protection Agency. (2006). 2010/2015 PFOA stewardship
Mohsenzadeh, F., Rad, A. C., & Akbari, M. (2012). Evaluation of oil removal program. Washington, DC: Author.
efficiency and enzymatic activity in some fungal strains for bioremedia- Vilaplana, M., García, A. B., Caminal, G., Guillén, F., & Sarrà, M. (2012). Opti-
tion of petroleum-polluted soils. Iranian Journal of Environmental Health misation of the operational conditions of trichloroethylene degradation
Science & Engineering, 9. doi: 10.1186/1735-2746-9-26 using Trametes versicolor under quinone redox cycling conditions using
Moody, C. A., & Field, J. A. (1999). Determination of perfluorocarboxylates central composite design methodology. Biodegradation, 23(2), 333–
in groundwater impacted by fire-fighting activity. Environmental Science 341.
& Technology, 33(16), 2800–2806. Wang, L., Li, Y., Yu, P., Xie, Z., Luo, Y., & Lin, Y. (2010). Biodegradation of phe-
Newcombe, D., Paszczynski, A., Gajewska, W., Kroger, M., Feis, G., & Craw- nol at high concentration by a novel fungal strain Paecilomyces variotii
ford, R. (2002). Production of small molecular weight catalysts and JH6. Journal of Hazardous Materials, 183(1–3), 366–371.
the mechanism of trinitrotoluene degradation by several Gloeophyllum Wang, N., Buck, R. C., Szostek, B., Sulecki, L. M., & Wolstenholme, B. W.
species. Enzyme and Microbial Technology, 30(4), 506–517. (2012). 5:3 Polyfluorinated acid aerobic biotransformation in activated
Ntwampe, S. K. O., Williams, C. C., & Sheldon, M. S. (2010). Influence of per- sludge via novel “one-carbon removal pathways”. Chemosphere, 87(5),
fluorocarbons on Phanerochaete chrysosporium biomass development, 527–534.
substrate consumption and enzyme production. Chemical and Biochemi- Wang, N., Szostek, B., Buck, R. C., Folsom, P. W., Sulecki, L. M., & Gannon, J.
cal Engineering Quarterly, 24(2), 187–194. T. (2009). 8-2 Fluorotelomer alcohol aerobic soil biodegradation: Path-
Ochoa-Herrera, V., Field, J. A., Luna-Velasco, A., & Sierra-Alvarez, R. (2016). ways, metabolites, and metabolite yields. Chemosphere, 75(8), 1089–
Microbial toxicity and biodegradability of perfluorooctane sulfonate 1096.
(PFOS) and shorter chain perfluoroalkyl and polyfluoroalkyl substances Weathers, T. S., Harding-Marjanovic, K., Higgins, C. P., Alvarez-Cohen, L., &
(PFASs). Environmental Science Process Impacts, 18(9), 1236–1246. Sharp, J. O. (2016). Perfluoroalkyl acids inhibit reductive dechlorination
Paul, A. G., & Jones, K. C. (2009). A first global production, emission, and of trichloroethene by repressing Dehalococcoides. Environmental Science
environmental inventory for perfluorooctane sulfonate. Environmental & Technology, 50(1), 240–248.
Science & Technology, 43(2), 386–392. Weathers, T. S., Higgins, C. P., & Sharp, J. O. (2015). Enhanced biofilm pro-
Prewitt, M. L., Diehl, S. V., McElroy, T. C., & Diehl, W. J. (2008). Comparison of duction by a toluene-degrading Rhodococcus observed after exposure
general fungal and basidiomycete-specific ITS primers for identification to perfluoroalkyl acids. Environmental Science & Technology, 49(9), 5458–
of wood decay fungi. Forest Products Journal, 58(4), 66–71. 5466.
Ritter, S. K. (2010). Fluorochemicals go short. Chemical Engineering News, 88, Yadav, J. S., Doddapaneni, H., & Subramanian, V. (2006). P450ome of the
12–17. white rot fungus Phanerochaete chrysosporium: Structure, evolution and
Ritz, K., & Young, I. M. (2004). Interactions between soil structure and fungi. regulation of expression of genomic P450 clusters. Biochemical Society
Mycologist, 18, 52–59. Transactions, 34(6), 1165–1169.
70 MERINO ET AL .

Zhang, S., Merino, N., Wang, N., Ruan, T., & Lu, X. (2017). Impact of 6:2 flu- Elisabeth L. Hawley, PE, is a senior consultant at Geosyntec Consul-
orotelomer alcohol aerobic biotransformation on a sediment microbial tants in Oakland, California. She is a professional engineer with more
community. Science of The Total Environment, 575, 1361–1368.
than 15 years of experience with environmental remediation, focusing
Zhang, S., Szostek, B., McCausland, P. K., Wolstenholme, B. W., Lu, X., Wang,
on PFASs and other emerging contaminants including applied research,
N., & Buck, R. C. (2013). 6:2 and 8:2 fluorotelomer alcohol anaerobic bio-
site characterization, fate and transport, and innovative treatment. She
transformation in digester sludge from a WWTP under methanogenic
conditions. Environmental Science & Technology, 47, 4227–4235. earned an MS degree in civil and environmental engineering and a BS
Zhao, L., Folsom, P. W., Wolstenholme, B. W., Sun, H., Wang, N., & Buck, R. C. in environmental engineering science from the University of California,
(2013). 6:2 Fluorotelomer alcohol biotransformation in an aerobic river Berkeley.
sediment system. Chemosphere, 90(2), 203–209.

Rula A. Deeb, PhD, is a senior principal at Geosyntec Consultants in


AUTHOR'S BIOGRAPHIES Oakland, California. She is a professional civil and environmental engi-
neer with more than 25 years of experience in private practice and
Nancy Merino, PhD, is currently a research fellow at the Earth Life academia addressing the cross-media fate and transport of contami-
Science Institute and the University of Southern California analyz- nants and remediation of complex soil and groundwater sites impacted
ing the microbial community of two terrestrial serpentinizing systems by non-aqueous phase liquids. She earned her PhD and postdoctoral
(pH∼10.5 to 12). Nancy obtained her PhD from University of Califor- fellowship from the University of California, Berkeley.
nia, Los Angeles (UCLA) in 2016, and was working on the fungal bio-
transformation of per- and polyfluoroalkyl compounds in Dr. Shaily Linda Y. Tseng, PhD, is an assistant professor at Colgate University in
Mahendra's laboratory. the Environmental Studies Program and the Department of Physics
and Astronomy. She received her PhD in environmental engineering
Meng Wang is a PhD student in Dr. Shaily Mahendra's laboratory at from the University of California, Irvine, in 2012, and was a postdoc-
UCLA. His research interests include application of bionanomaterials toral research associate in Dr. Shaily Mahendra's laboratory at UCLA.
and enzymes in environmental remediation and water treatment. Her research interests include extracellular polymeric substances, and
the fate and transport of emerging contaminants.

Rocio Ambrocio was an undergraduate research assistant in Dr. Shaily


Shaily Mahendra, PhD, is an associate professor in the UCLA Depart-
Mahendra's laboratory. She received her bachelor degree in civil and
ment of Civil and Environmental Engineering. She received PhD from
environmental engineering from UCLA in 2016, and is currently a sci-
University of California, Berkeley, and postdoctoral fellowship from
ence instructor for the Science Discovery Center in Orange County,
Rice University. She recently won the Paul L. Busch Award, National
California.
Science Foundation CAREER Award, DuPont Young Professor Award,
Northrop Grumman Excellence in Teaching Award, Samueli Fellowship,
Kimberly Mak was an undergraduate research assistant in Dr. Shaily
Hellman Fellowship, Poptech Science and Public Leadership Fellow-
Mahendra's laboratory. She received her bachelor degree in environ-
ship, and Environmental Science & Technology Excellence in Review
mental science from UCLA in 2016, and is currently an independent
Award. Her research areas are microbial processes in natural and engi-
compliance monitor with the environmental consulting firm Environ-
neered systems, applications of molecular and isotopic tools in envi-
mental Intelligence.
ronmental microbiology, environmental applications of nanomaterials,
and biotransformation of water contaminants.
Ellen O'Connor is a graduate student obtaining her PhD in molecular
toxicology from UCLA where she did work on per- and polyfluoroalkyl
SUPPORTING INFORMATION
compounds during a rotation in Dr. Shaily Mahendra's laboratory. Ms.
O'Connor is now studying the role of oxidized lipids in the development Additional Supporting Information may be found online in the support-
of pulmonary hypertension. ing information tab for this article.

An Gao was an undergraduate research assistant under project “Grand How to cite this article: Merino N, Wang M, Ambrocio R, et al.
Challenges” in Dr. Shaily Mahendra's laboratory. She received her Fungal biotransformation of 6:2 fluorotelomer alcohol. Remedi-
bachelor degree in civil and environmental engineering from UCLA in ation. 2018;28:59–70. https://doi.org/10.1002/rem.21550
2016, and is currently a junior estimator at Bernards Bros Construc-
tion.

You might also like