Search
2025 Volume 10
Article Contents
ARTICLE   Open Access    

Taxonomy, phylogeny, and preliminary screening of fungal isolates for cadmium tolerance from Iloilo Ferry Terminal, Iloilo, Philippines

More Information
  • Received: 21 October 2024
    Revised: 24 December 2024
    Accepted: 25 December 2024
    Published online: 23 January 2025
    Studies in Fungi  10 Article number: e001 (2025)  |  Cite this article
  • Water pollution by heavy metals poses a growing threat to ecosystems and human health. Bioremediation using fungi offers a promising solution due to their tolerance and absorption capabilities. This study aimed to isolate and identify cadmium-tolerant fungi from the Iloilo River, a potentially polluted river in Iloilo City, Philippines using morpho-phylogenetic analysis. Water and sediment samples were collected, and fungi were isolated by serial dilution on PDA plates supplemented with cadmium ions [Cd(II)] (25 ppm). Isolates were further tested for tolerance to varying cadmium concentrations (50 and 100 ppm). Nine fungal species were identified, including three ascomycetous taxa (Pseudopestalotiopsis elaeidis, Peroneutypa scoparia, Diaporthe sp.) and six basidiomycetous taxa (Cerrena sp., Irpex laceratus, Phlebia sp., Pyrrhoderma noxium, Schizophyllum commune, Trichosporon asahii). The majority of the isolates originated from sediment samples. Morphological illustrations, descriptions, and taxonomic notes are provided for each taxon. These findings provide insight into the fungal diversity of the Iloilo River and their tolerance to cadmium.
  • 加载中
  • [1]

    Vardhan KH, Kumar PS, Panda RC. 2019. A review on heavy metal pollution, toxicity and remedial measures: Current trends and future perspectives. Journal of Molecular Liquids 290:111197

    doi: 10.1016/J.MOLLIQ.2019.111197

    CrossRef   Google Scholar

    [2]

    Zamora-Ledezma C, Negrete-Bolagay D, Figueroa F, Zamora-Ledezma E, Ni M, et al. 2021. Heavy metal water pollution: A fresh look about hazards, novel and conventional remediation methods. Environmental Technology and Innovation 22:101504

    doi: 10.1016/J.ETI.2021.101504

    CrossRef   Google Scholar

    [3]

    Kapahi M, Sachdeva S. 2019. Bioremediation options for heavy metal pollution. Journal of Health and Pollution 9(24):191203

    doi: 10.5696/2156-9614-9.24.191203

    CrossRef   Google Scholar

    [4]

    Briffa J, Sinagra E, Blundell R. 2020. Heavy metal pollution in the environment and their toxicological effects on humans. Heliyon 6(9):e04691

    doi: 10.1016/j.heliyon.2020.e04691

    CrossRef   Google Scholar

    [5]

    Afzaal M, Hameed S, Liaqat I, Ali Khan AA, Abdul-Manan H, et al. 2022. Heavy metals contamination in water, sediments and fish of freshwater ecosystems in Pakistan. Water Practice and Technology 17(5):1253−72

    doi: 10.2166/WPT.2022.039

    CrossRef   Google Scholar

    [6]

    Muhammad M, Khan S, Shehzadi SA, Gul Z, Al-Saidi HM et al. 2022. Recent advances in colorimetric and fluorescent chemosensors based on thiourea derivatives for metallic cations: A review. Dyes and Pigments 205:110477

    doi: 10.1016/j.dyepig.2022.110477

    CrossRef   Google Scholar

    [7]

    El-Gendi H, Saleh AK, Badierah R, Redwan EM, El-Maradny YA, et al. 2021. A comprehensive insight into fungal enzymes: structure, classification, and their role in mankind's challenges. Journal of Fungi 8(1):23

    doi: 10.3390/jof8010023

    CrossRef   Google Scholar

    [8]

    Paul D. 2017. Research on heavy metal pollution of river Ganga: A review. Annals of Agrarian Science 15(2):278−86

    doi: 10.1016/J.AASCI.2017.04.001

    CrossRef   Google Scholar

    [9]

    He Z, Shentu J, Yang X, Baligar V, Zhang T, et al. 2015. Heavy metal contamination of soils: sources, indicators, and assessment. Journal of Environmental Indicators 9:17−18

    Google Scholar

    [10]

    Montaño-López F, Biswas A. 2021. Are heavy metals in urban garden soils linked to vulnerable populations? A case study from Guelph, Canada. Scientific Reports 11(1):11286

    doi: 10.1038/s41598-021-90368-3

    CrossRef   Google Scholar

    [11]

    Baysal A, Ozbek N, Akm S. 2013. Determination of trace metals in waste water and their removal processes. Waste Water - Treatment Technologies and Recent Analytical Developments, eds. Einschlag FSG, Carlos L. Rijeka: IntechOpen. doi:10.5772/52025

    [12]

    Hama Aziz KH, Mustafa FS, Omer KM, Hama S, Hamarawf RF, et al. 2023. Heavy metal pollution in the aquatic environment: efficient and low-cost removal approaches to eliminate their toxicity: a review. RSC Advances 13(26):17595−610

    doi: 10.1039/D3RA00723E

    CrossRef   Google Scholar

    [13]

    Razzak SA, Faruque MO, Alsheikh Z, Alsheikhmohamad L, Alkuroud D, et al. 2022. A comprehensive review on conventional and biological-driven heavy metals removal from industrial wastewater. Environmental Advances 7:100168

    doi: 10.1016/J.ENVADV.2022.100168

    CrossRef   Google Scholar

    [14]

    Dell'anno F, Rastelli E, Buschi E, Barone G, Beolchini F, et al. 2022. Fungi can be more effective than bacteria for the bioremediation of marine sediments highly contaminated with heavy metals. Microorganisms 10(5):993

    doi: 10.3390/microorganisms10050993/S1

    CrossRef   Google Scholar

    [15]

    Talukdar D, Jasrotia T, Sharma R, Jaglan S, Kumar R, et al. 2020. Evaluation of novel indigenous fungal consortium for enhanced bioremediation of heavy metals from contaminated sites. Environmental Technology and Innovation 20:101050

    doi: 10.1016/J.ETI.2020.101050

    CrossRef   Google Scholar

    [16]

    Talukdar D, Sharma R, Jaglan S, Vats R, Kumar R, et al. 2020. Identification and characterization of cadmium resistant fungus isolated from contaminated site and its potential for bioremediation. Environmental Technology and Innovation 17:100604

    doi: 10.1016/J.ETI.2020.100604

    CrossRef   Google Scholar

    [17]

    Jayaraman M, Arumugam R. 2014. Metal tolerance analysis of microfungi isolated from metal contaminated soil and waste water. Journal of Microbiology, Biotechnology and Food Sciences 4(1):63−66

    doi: 10.15414/JMBFS.2014.4.1.63-66

    CrossRef   Google Scholar

    [18]

    Joshi PK, Swarup A, Maheshwari S, Kumar R, Singh N. 2011. Bioremediation of heavy metals in liquid media through fungi isolated from contaminated sources. Indian Journal of Microbiology 51(4):482−87

    doi: 10.1007/s12088-011-0110-9

    CrossRef   Google Scholar

    [19]

    Dusengemungu L, Kasali G, Gwanama C, Ouma KO. 2020. Recent advances in biosorption of copper and cobalt by filamentous fungi. Frontiers in Microbiology 11:582016

    doi: 10.3389/fmicb.2020.582016

    CrossRef   Google Scholar

    [20]

    Gajewska J, Floryszak-Wieczorek J, Sobieszczuk-Nowicka E, Mattoo A, Arasimowicz-Jelonek M. 2022. Fungal and oomycete pathogens and heavy metals: an inglorious couple in the environment. IMA Fungus 13(1):6

    doi: 10.1186/s43008-022-00092-4

    CrossRef   Google Scholar

    [21]

    Rajapaksha RMCP, Tobor-Kapłon MA, Bååth E. 2004. Metal toxicity affects fungal and bacterial activities in soil differently. Applied Environmental Microbiology 70(5):2966−73

    doi: 10.1128/AEM.70.5.2966-2973.2004

    CrossRef   Google Scholar

    [22]

    Taha A, Hussien W, Gouda S. 2023. Bioremediation of heavy metals in wastewaters: a concise review. Egyptian Journal of Aquatic Biology and Fisheries 27:143−66

    doi: 10.21608/EJABF.2023.284415

    CrossRef   Google Scholar

    [23]

    Pantidos N, Horsfall LE. 2014. Biological synthesis of metallic nanoparticles by bacteria, fungi and plants. Journal of Nanomedicine & Nanotechnology 5:5

    doi: 10.4172/2157-7439.1000233

    CrossRef   Google Scholar

    [24]

    Valdez JBT. 2021. Iloilo River: Annual Assessment Report, CY 2021. Report, CY 2021. Department of Environment and Natural Resources. Environmental Management Bureau – R6

    [25]

    Senanayake IC, Rathnayaka AR, Marasinghe DS, Calabon MS, Gentekaki E et al. 2020. Morphological approaches in studying fungi: collection, examination, isolation, sporulation and preservation. Mycosphere 11:2678−54

    doi: 10.5943/mycosphere/11/1/20

    CrossRef   Google Scholar

    [26]

    White TJ, Bruns T, Lee S, Taylor J. 1990. Amplification and direct sequencing of fungal ribosomal RNA genes for phylogenetics. In PCR-protocols: a guide to methods and applications, eds. MA Innis, DH Gelfand, JJ Shinsky, JJ White. San Diego: Academic Press Inc. pp. 315–22. doi: 10.1016/B978-0-12-372180-8.50042-1

    [27]

    Trifinopoulos J, Nguyen LT, von Haeseler A, Minh BQ. 2016. W-IQ-TREE: a fast online phylogenetic tool for maximum likelihood analysis. Nucleic Acids Research 44(W1):W232−W235

    doi: 10.1093/nar/gkw256

    CrossRef   Google Scholar

    [28]

    Miller MA, Pfeiffer W, Schwartz T. 2010. Creating the CIPRES Science Gateway for inference of large phylogenetic trees. 2010 Gateway Computing Environments Workshop (GCE), 14 November 2010, New Orleans, LA, USA. USA: IEEE. pp. 1–8. doi: 10.1109/GCE.2010.5676129

    [29]

    Dissanayake AJ, Bhunjun CS, Maharachchikumbura SSN, Liu JK. 2020. Applied aspects of methods to infer phylogenetic relationships amongst fungi. Mycosphere 11:2652−76

    doi: 10.5943/MYCOSPHERE/11/1/18

    CrossRef   Google Scholar

    [30]

    Liu F, Bonthond G, Groenewald JZ, Cai L, Crous PW. 2019. Sporocadaceae, a family of coelomycetous fungi with appendage-bearing conidia. Studies in Mycology 92:287−415

    doi: 10.1016/J.SIMYCO.2018.11.001

    CrossRef   Google Scholar

    [31]

    Liu F, Hou L, Raza M, Cai L. 2017. Pestalotiopsis and allied genera from Camellia, with description of 11 new species from China. Scientific Reports 7:866

    doi: 10.1038/s41598-017-00972-5

    CrossRef   Google Scholar

    [32]

    Nozawa S, Yamaguchi K, Hoang Yen LT, Van Hop D, Nyunt P, et al. 2017. Identification of two new species and a sexual morph from the genus Pseudopestalotiopsis. Mycoscience 58(5):328−37

    doi: 10.1016/J.MYC.2017.02.008

    CrossRef   Google Scholar

    [33]

    Carmarán CC, Romero AI, Giussani LM. 2006. An approach towards a new phylogenetic classification in Diatrypaceae. Fungal Diversity 23:67−87

    Google Scholar

    [34]

    Tiffany L, Gilman J. 1965. Iowa Ascomycetes IV. Diatrypaceae. Iowa State College Journal of Science 40:121−61

    Google Scholar

    [35]

    Rappaz F. 1987. Taxonomy and nomenclature of the octosporous Diatrypaceae. Mycologia Helvetica 2:285−648

    Google Scholar

    [36]

    Yuan ZQ. 1996. Fungi and associated tree diseases in Melville Island, Northern Territory, Australia. Australian Systematic Botany 9(3):337−60

    doi: 10.1071/SB9960337

    CrossRef   Google Scholar

    [37]

    Urbez-Torres JR, Peduto F, Striegler RK, Urrea-Romero KE, Rupe JC, et al. 2012. Characterization of fungal pathogens associated with grapevine trunk diseases in Arkansas and Missouri. Fungal Diversity 52(1):169−89

    doi: 10.1007/S13225-011-0110-4

    CrossRef   Google Scholar

    [38]

    Mayorquin JS, Wang DH, Twizeyimana M, Eskalen A. 2016. Identification, distribution, and pathogenicity of Diatrypaceae and Botryosphaeriaceae associated with citrus branch canker in the Southern California desert. Plant Disease 100(12):2402−13

    doi: 10.1094/PDIS-03-16-0362-RE

    CrossRef   Google Scholar

    [39]

    Castilla-Cayuman A, Lolas M, Díaz G. 2019. First report of Peroneutypa scoparia causing cane dieback in kiwifruit in Chile. Plant Disease 103(2):373

    doi: 10.1094/PDIS-06-18-1092-PDN

    CrossRef   Google Scholar

    [40]

    Eichmeier A, Pecenka J, Spetik M, Necas T, Ondrasek I, et al. 2020. Fungal trunk pathogens associated with Juglans regia in the Czech Republic. Plant Disease 104(3):761−71

    doi: 10.1094/PDIS-06-19-1308-RE

    CrossRef   Google Scholar

    [41]

    Shang QJ, Hyde KD, Jeewon R, Khan S, Promputtha I, et al. 2018. Morpho-molecular characterization of Peroneutypa (Diatrypaceae, Xylariales) with two novel species from Thailand. Phytotaxa 356(1):1−18

    doi: 10.11646/phytotaxa.356.1.1

    CrossRef   Google Scholar

    [42]

    Ghimire SR, Charlton ND, Bell JD, Krishnamurthy YL, Craven KD. 2011. Biodiversity of fungal endophyte communities inhabiting switchgrass (Panicum virgatum L. ) growing in the native tallgrass prairie of northern Oklahoma. Fungal Diversity 47(1):19−27

    doi: 10.1007/S13225-010-0085-6

    CrossRef   Google Scholar

    [43]

    Pandi A, Kuppuswami G, Ramudu K, Palanivel, S. 2019. A sustainable approach for degradation of leather dyes by a new fungal laccase. Journal of Cleaner Production 211:590−97

    doi: 10.1016/j.jclepro.2018.11.048

    CrossRef   Google Scholar

    [44]

    Bilal M, Ashraf SS, Iqbal HMN. 2020. Laccase-Mediated bioremediation of dye-based hazardous pollutants. In Methods for Bioremediation of Water and Wastewater Pollution. Environmental Chemistry for a Sustainable World, eds. Inamuddin, Ahamed MI, Lichtfouse E, Asiri AM. vol 51. Cham: Springer. pp. 137−60. doi: 10.1007/978-3-030-48985-4_6

    [45]

    Gao Y, Liu F, Duan W, Crous PW, Cai L. 2017. Diaporthe is paraphyletic. IMA Fungus 8(1):153−87

    doi: 10.5598/imafungus.2017.08.01.11

    CrossRef   Google Scholar

    [46]

    Hyde KD, Norphanphoun C, Maharachchikumbura SSN, Bhat DJ, Jones EBG, et al. 2020. Refined families of Sordariomycetes. Mycosphere 11(1):305−1059

    doi: 10.5943/mycosphere/11/1/7

    CrossRef   Google Scholar

    [47]

    Norphanphoun C, Gentekaki E, Hongsanan S, Jayawardena R, Senanayake IC, et al. 2022. Diaporthe: formalizing the species-group concept. Mycosphere 13(1):752−819

    doi: 10.5943/mycosphere/13/1/9

    CrossRef   Google Scholar

    [48]

    Dissanayake AJ, Phillips AJL, Hyde KD, Yan JY, Li XH. 2017. The current status of species in Diaporthe. Mycosphere 8(5):1106−56

    doi: 10.5943/mycosphere/8/5/5

    CrossRef   Google Scholar

    [49]

    Phukhamsakda C, McKenzie EHC, Phillips AJL, Gareth Jones EB, Jayarama Bhat D. 2020. Microfungi associated with Clematis (Ranunculaceae) with an integrated approach to delimiting species boundaries. Fungal diversity 102(1):1−203

    doi: 10.1007/s13225-020-00448-4

    CrossRef   Google Scholar

    [50]

    Zervakis G, Dimou D, Balis C. 1998. A check-list of the Greek macrofungi, including hosts and biogeographic distribution: I. Basidiomycotina. Mycotaxon 66:273−336

    Google Scholar

    [51]

    Gafforov Y, Ordynets A, Langer E, Yarasheva M, de Mello Gugliotta A, et al. 2020. Species diversity with comprehensive annotations of wood-inhabiting poroid and corticioid fungi in Uzbekistan. Frontiers in Microbiology 11:598321

    doi: 10.3389/fmicb.2020.598321

    CrossRef   Google Scholar

    [52]

    Miettinen O, Vlasák J, Larsson E, Vlasák J Jr, Seelan JSS, et al. 2023. A revised genus-level classification for Cerrenaceae (Polyporales, Agaricomycetes). Fungal Systematics and Evolution 12:271−322

    doi: 10.3114/FUSE.2023.12.14

    CrossRef   Google Scholar

    [53]

    Pawlik A, Ruminowicz-Stefaniuk M, Frąc M, Mazur A, Wielbo J, et al. 2019. The wood decay fungus Cerrena unicolor adjusts its metabolism to grow on various types of wood and light conditions. PLoS One 14(2):e0211744

    doi: 10.1371/journal.pone.0211744

    CrossRef   Google Scholar

    [54]

    Cha HJ, Chiang MWL, Guo SY, Lin SM, Pang KL. 2021. Culturable fungal community of Pterocladiella capillacea in Keelung, Taiwan: Effects of surface sterilization method and isolation medium. Journal of Fungi 7:651

    doi: 10.3390/jof7080651

    CrossRef   Google Scholar

    [55]

    Glen M, Yuskianti V, Puspitasari D, Francis A, Agustini L, et al. 2014. Identification of basidiomycete fungi in Indonesian hardwood plantations by DNA barcoding. Forest Pathology 44:496−508

    doi: 10.1111/efp.12146

    CrossRef   Google Scholar

    [56]

    Pang KL, Guo SY, Chen IA, Burgaud G, Luo ZH, et al. 2019. Insights into fungal diversity of a shallow-water hydrothermal vent field at Kueishan Island, Taiwan by culture-based and metabarcoding analyses. PLOS ONE 14:e0226616

    doi: 10.1371/journal.pone.0226616

    CrossRef   Google Scholar

    [57]

    D'Souza-Ticlo D, Sharma D, Raghukumar C. 2009. A thermostable metal-tolerant laccase with bioremediation potential from a marine-derived fungus. Marine Biotechnology 11(6):725−37

    doi: 10.1007/s10126-009-9187-0

    CrossRef   Google Scholar

    [58]

    Xu X, Huang X, Liu D, Lin J, Ye X, et al. 2018. Inhibition of metal ions on Cerrena sp. laccase: Kinetic, decolorization and fluorescence studies. Journal of the Taiwan Institute of Chemical Engineers 84:1−10

    doi: 10.1016/j.jtice.2017.12.028

    CrossRef   Google Scholar

    [59]

    Hassan A, Periathamby A, Ahmed A, Innocent O, Hamid FS. 2020. Effective bioremediation of heavy metal–contaminated landfill soil through bioaugmentation using consortia of fungi. Journal of Soils and Sediments 20:66−80

    doi: 10.1007/s11368-019-02394-4

    CrossRef   Google Scholar

    [60]

    Viswanath B, Rajesh B, Janardhan A, Kumar AP, Narasimha G. 2014. Fungal laccases and their applications in bioremediation. Enzyme Research 2014:163242

    doi: 10.1155/2014/163242

    CrossRef   Google Scholar

    [61]

    Chandra R, Chowdhary P. 2015. Properties of bacterial laccases and their application in bioremediation of industrial wastes. Environmental Science. Processes & Impacts 17(2):326−42

    doi: 10.1039/c4em00627e

    CrossRef   Google Scholar

    [62]

    Robene-Soustrade I, Lung-Escarmant B. 1997. Laccase isoenzyme patterns of European Armillaria species from culture filtrates and infected woody plant tissues. European Journal of Forest Pathology 27(2):105−14

    doi: 10.1111/j.1439-0329.1997.tb01361.x

    CrossRef   Google Scholar

    [63]

    Yang J, Wang G, Ng TB, Lin J, Ye X. 2016. Laccase production and differential transcription of laccase genes in Cerrena sp. in response to metal ions, aromatic compounds, and nutrients. Frontiers in Microbiology 6:1558

    doi: 10.3389/fmicb.2015.01558

    CrossRef   Google Scholar

    [64]

    Baldrian P, Gabriel J. 2002. Copper and cadmium increase laccase activity in Pleurotus ostreatus. FEMS Microbiology Letters 206(1):69−74

    doi: 10.1111/j.1574-6968.2002.tb10988.x

    CrossRef   Google Scholar

    [65]

    Goligar N, Saadatmand S, Khavarinejad RA. 2023. Mycoremediation of lead and cadmium by lignocellulosic enzymes of Pleurotus eryngii. AMB Express 13:127

    doi: 10.1186/s13568-023-01626-8

    CrossRef   Google Scholar

    [66]

    Rodríguez Couto S, Sanromán M, Gübitz GM. 2005. Influence of redox mediators and metal ions on synthetic acid dye decolourization by crude laccase from Trametes hirsuta. Chemosphere 58(4):417−22

    doi: 10.1016/j.chemosphere.2004.09.033

    CrossRef   Google Scholar

    [67]

    Chen CC, Chen CY, Wu SH. 2021. Species diversity, taxonomy and multi-gene phylogeny of phlebioid clade (Phanerochaetaceae, Irpicaceae, Meruliaceae) of Polyporales. Fungal Diversity 111(1):337−442

    doi: 10.1007/s13225-021-00490-w

    CrossRef   Google Scholar

    [68]

    Suhara H, Maekawa N, Kaneko S, Hattori T, Sakai K, et al. 2003. A new species, Ceriporia lacerata, isolated from white-rotted wood. Mycotaxon 86:335−47

    Google Scholar

    [69]

    Wu B, Hussain M, Zhang W, Stadler M, Liu X, et al. 2019. Current insights into fungal species diversity and perspective on naming the environmental DNA sequences of fungi. Mycology 10(3):127−40

    doi: 10.1128/microbiolspec.FUNK-0052-2016

    CrossRef   Google Scholar

    [70]

    Buzina W, Lass-Flörl C, Kropshofer G, Freund MC, Marth E. 2005. The polypore mushroom Irpex lacteus, a new causative agent of fungal infections. Journal of Clinical Microbiology 43(4):2009−11

    doi: 10.1128/JCM.43.4.2009-2011.2005

    CrossRef   Google Scholar

    [71]

    Zhao Y, Li SQ, Li HJ, Lan WJ. 2013. Lanostane triterpenoids from the fungus Ceriporia lacerate associated with Acanthaster planci. Chemistry of Natural Compounds 49(4):653−56

    doi: 10.1007/s10600-013-0701-2

    CrossRef   Google Scholar

    [72]

    Fang ST, Miao FP, Yin XL, Ji NY. 2024. A new lanostane-type triterpenoid from the marine shellfish symbiotic fungus Ceriporia lacerata CD7-5. Natural Product Research 38:1510−16

    doi: 10.1080/14786419.2022.2154345

    CrossRef   Google Scholar

    [73]

    Gareth Jones EB, Pang KL, Abdel-Wahab MA, Scholz B, Hyde KD, et al. 2019. An online resource for marine fungi. Fungal Diversity 96(1):347−433

    doi: 10.1007/S13225-019-00426-5

    CrossRef   Google Scholar

    [74]

    Calabon MS, Gareth Jones EB, Pang KL, Abdel-Wahab MA, Jin J, et al. 2023. Updates on the classification and numbers of marine fungi. Botanica Marina 66(4):213−38

    doi: 10.1515/bot-2023-0032

    CrossRef   Google Scholar

    [75]

    Trianto A, Radjasa OK, Subagiyo S, Purnaweni H, Bahry MS, et al. 2021. Potential of fungi isolated from a mangrove ecosystem in Northern Sulawesi, Indonesia: Protease, cellulase and anti-microbial capabilities. Biodiversitas 22(4):1717−24

    doi: 10.13057/biodiv/d220415

    CrossRef   Google Scholar

    [76]

    Kuuskeri J, Mäkelä MR, Isotalo J, Oksanen I, Lundell T. 2015. Lignocellulose-converting enzyme activity profiles correlate with molecular systematics and phylogeny grouping in the incoherent genus Phlebia (Polyporales, Basidiomycota). BMC microbiology 15:217

    doi: 10.1186/s12866-015-0538-x

    CrossRef   Google Scholar

    [77]

    Li X, Kondo R, Sakai K. 2002. Biodegradation of sugarcane bagasse with marine fungus Phlebia sp. MG-60. Journal of Wood Science 48(2):159−62

    doi: 10.1007/BF00767294

    CrossRef   Google Scholar

    [78]

    Akar T, Aydın P, Celik S, Akar ST. 2020. Phlebia gigantea cells immobilized on renewable biomass matrix as potential ecofriendly scavenger for lead contamination. Environmental Science and Pollution Research International 27(14):16177−88

    doi: 10.1007/s11356-020-07889-z

    CrossRef   Google Scholar

    [79]

    Sharma KR, Giri R, Sharma RK. 2020. Lead, cadmium and nickel removal efficiency of white-rot fungus Phlebia brevispora. Letters in Applied Microbiology 71(6):637−44

    doi: 10.1111/lam.13372

    CrossRef   Google Scholar

    [80]

    Abdel-Hamid AM, Solbiati JO, Cann IKO. 2013. Insights into lignin degradation and its potential industrial applications. Advances in Applied Microbiology 82:1−28

    doi: 10.1016/b978-0-12-407679-2.00001-6

    CrossRef   Google Scholar

    [81]

    Chen L, Zhang X, Zhang M, Zhu Y, Zhuo R. 2022. Removal of heavy-metal pollutants by white rot fungi: Mechanisms, achievements, and perspectives. Journal of Cleaner Production 354:131681

    doi: 10.1016/j.jclepro.2022.131681

    CrossRef   Google Scholar

    [82]

    Munk L, Sitarz AK, Kalyani DC, Mikkelsen JD, Meyer AS. 2015. Can laccases catalyze bond cleavage in lignin? Biotechnology Advances 33(1):13−24

    doi: 10.1016/j.biotechadv.2014.12.008

    CrossRef   Google Scholar

    [83]

    Zhou LW, Ji XH, Vlasák J, Dai YC. 2018. Taxonomy and phylogeny of Pyrrhoderma: a redefinition, the segregation of Fulvoderma, gen. nov. , and identifying four new species. Mycologia 110(5):872−89

    doi: 10.1080/00275514.2018.1474326

    CrossRef   Google Scholar

    [84]

    Chang TT. 1995. Decline of nine tree species associated with brown root rot caused by Phellinus noxius. Plant Disease 79:962−65

    doi: 10.1094/PD-79-0962

    CrossRef   Google Scholar

    [85]

    Ann PJ, Chang TT, Ko WH. 2002. Phellinus noxius brown root rot of fruit and ornamental trees in Taiwan. Plant Disease 86(8):820−826

    doi: 10.1094/PDIS.2002.86.8.820

    CrossRef   Google Scholar

    [86]

    Schwarze FWMR, Jauss F, Spencer C, Hallam C, Schubert M. 2012. Evaluation of an antagonistic Trichoderma strain for reducing the rate of wood decomposition by the white rot fungus Phellinus noxius. Biological Control 61(2):160−68

    doi: 10.1016/J.BIOCONTROL.2012.01.016

    CrossRef   Google Scholar

    [87]

    Adikaram NKB, Yakandawala DMD. 2020. A checklist of plant pathogenic fungi and Oomycota in Sri Lanka. Ceylon Journal of Science 49(1):93−123

    doi: 10.4038/CJS.V49I1.7709

    CrossRef   Google Scholar

    [88]

    Wang JW, Liu DY, Zhang HZ, Tan Z, Zheng CJ, et al. 2024. Drimane-type sesquiterpenoids and their anti-inflammatory evaluation from Pyrrhoderma noxium HNNU0524. Natural Product Research 38(10):1711−18

    doi: 10.1080/14786419.2023.2218008

    CrossRef   Google Scholar

    [89]

    Raper JR, Miles PG. 1958. The genetics of Schizophyllum commune. Genetics 43(3):530

    doi: 10.1093/genetics/43.3.530

    CrossRef   Google Scholar

    [90]

    Arifeen MZU, Yang X, Li F, Xue Y, Gong P, et al. 2020. Growth behaviors of deep subseafloor Schizophyllum commune in response to various environmental conditions. Acta Microbiologica Sinica 60(9):1882−92

    doi: 10.13343/J.CNKI.WSXB.20200157

    CrossRef   Google Scholar

    [91]

    Arifeen MZU, Chu C, Yang X, Liu J, Huang X, et al. 2021. The anaerobic survival mechanism of Schizophyllum commune 20R-7-F01, isolated from deep sediment 2 km below the seafloor. Environmental Microbiology 23(2):1174−85

    doi: 10.1111/1462-2920.15332

    CrossRef   Google Scholar

    [92]

    Zhang X, Li Y, Yu Z, Liang X, Qi S. 2021. Phylogenetic diversity and bioactivity of culturable deep-sea-derived fungi from Okinawa Trough. Journal of Oceanology and Limnology 39(3):892−902

    doi: 10.1007/s00343-020-0003-z

    CrossRef   Google Scholar

    [93]

    Liu X, Huang X, Chu C, Xu H, Wang L, et al. 2022. Genome, genetic evolution, and environmental adaptation mechanisms of Schizophyllum commune in deep subseafloor coal-bearing sediments. iScience 25(6):104417

    doi: 10.1016/j.isci.2022.104417

    CrossRef   Google Scholar

    [94]

    Gnavi G, Garzoli L, Poli A, Prigione V, Burgaud G, et al. 2017. The culturable mycobiota of Flabellia petiolata: First survey of marine fungi associated to a Mediterranean green alga. PLoS One 12(4):e0175941

    doi: 10.1371/journal.pone.0175941

    CrossRef   Google Scholar

    [95]

    Garzoli L, Poli A, Prigione V, Gnavi G, Varese GC. 2018. Peacock's tail with a fungal cocktail: first assessment of the mycobiota associated with the brown alga Padina pavonica. Fungal Ecology 35:87−97

    doi: 10.1016/j.funeco.2018.05.005

    CrossRef   Google Scholar

    [96]

    Panno L, Bruno M, Voyron S, Anastasi A, Gnavi G, et al. 2013. Diversity, ecological role and potential biotechnological applications of marine fungi associated to the seagrass Posidonia oceanica. New Biotechnology 30(6):685−694

    doi: 10.1016/J.NBT.2013.01.010

    CrossRef   Google Scholar

    [97]

    Vu D, Groenewald M, de Vries M, Gehrmann T, Stielow B, et al. 2019. Large-scale generation and analysis of filamentous fungal DNA barcodes boosts coverage for kingdom fungi and reveals thresholds for fungal species and higher taxon delimitation. Studies in Mycology 92:135−54

    doi: 10.1016/j.simyco.2018.05.001

    CrossRef   Google Scholar

    [98]

    Wollenberg A, Kretzschmar J, Drobot B, Hübner R, Freitag L, et al. 2021. Uranium (VI) bioassociation by different fungi – a comparative study into molecular processes. Journal of Hazardous Materials 411:125068

    doi: 10.1016/j.jhazmat.2021.125068

    CrossRef   Google Scholar

    [99]

    Yan ZY, Zhao MR, Huang CY, Zhang LJ, Zhang JX. 2021. Trehalose alleviates high-temperature stress in Pleurotus ostreatus by affecting central carbon metabolism. Microbial Cell Factories 20(1):82

    doi: 10.1186/s12934-021-01572-9

    CrossRef   Google Scholar

    [100]

    Xu F, Chen P, Li H, Qiao S, Wang J, et al. 2021. Comparative transcriptome analysis reveals the differential response to cadmium stress of two Pleurotus fungi: Pleurotus cornucopiae and Pleurotus ostreatus. Journal of Hazardous Materials 416:125814

    doi: 10.1016/j.jhazmat.2021.125814

    CrossRef   Google Scholar

    [101]

    Günther A, Raff J, Merroun ML, Roßberg A, Kothe E, et al. 2014. Interaction of U(VI) with Schizophyllum commune studied by microscopic and spectroscopic methods. BioMetals 27(4):775−85

    doi: 10.1007/s10534-014-9772-1

    CrossRef   Google Scholar

    [102]

    Kleijburg FEL, Safeer AA, Baldus M, Wösten HAB. 2023. Binding of micro-nutrients to the cell wall of the fungus Schizophyllum commune. The Cell Surface 10:100108

    doi: 10.1016/j.tcsw.2023.100108

    CrossRef   Google Scholar

    [103]

    Kirtzel J, Scherwietes EL, Merten D, Krause K, Kothe E. 2019. Metal release and sequestration from black slate mediated by a laccase of Schizophyllum commune. Environmental Science and Pollution Research International 26(1):5−13

    doi: 10.1007/s11356-018-2568-z

    CrossRef   Google Scholar

    [104]

    Javaid A, Bajwa R, Javaid A. 2010. Biosorption of heavy metals using a dead macro fungus Schizophyllum commune Fries: evaluation of equilibrium and kinetic models. Pakistan Journal of Botany 42(3):2105−18

    Google Scholar

    [105]

    Traxler L, Shrestha J, Richter M, Krause K, Schäfer T, et al. 2022. Metal adaptation and transport in hyphae of the wood-rot fungus Schizophyllum commune. Journal of Hazardous Materials 425:127978

    doi: 10.1016/j.jhazmat.2021.127978

    CrossRef   Google Scholar

    [106]

    Guarro J, Gené J, Stchigel AM. 1999. Developments in fungal taxonomy. Clinical Microbiology Review 12(3):454−500

    doi: 10.1128/CMR.12.3.454

    CrossRef   Google Scholar

    [107]

    Li H, Guo M, Wang C, Li Y, Fernandez AM, et al. 2020. Epidemiological study of Trichosporon asahii infections over the past 23 years. Epidemiology and Infection 148:e169

    doi: 10.1017/S0950268820001624

    CrossRef   Google Scholar

    [108]

    Vogel C, Rogerson A, Schatz S, Laubach H, Tallman A, et al. 2007. Prevalence of yeasts in beach sand at three bathing beaches in South Florida. Water Research 41(9):1915−20

    doi: 10.1016/J.WATRES.2007.02.010

    CrossRef   Google Scholar

    [109]

    Singh P, Raghukumar C, Verma P, Shouche Y. 2012. Assessment of fungal diversity in deep-sea sediments by multiple primer approach. World Journal of Microbiology and Biotechnology 28(2):659−67

    doi: 10.1007/s11274-011-0859-3

    CrossRef   Google Scholar

    [110]

    Singh P, Raghukumar C, Meena RM, Verma P, Shouche Y. 2012. Fungal diversity in deep-sea sediments revealed by culture-dependent and culture-independent approaches. Fungal Ecology 5(5):543−53

    doi: 10.1016/J.FUNECO.2012.01.001

    CrossRef   Google Scholar

    [111]

    Chi ZM, Liu TT, Chi Z, Liu GL, Wang ZP. 2012. Occurrence and diversity of yeasts in the mangrove ecosystems in Fujian, Guangdong and Hainan provinces of China. Indian Journal of Microbiology 52(3):346−53

    doi: 10.1007/s12088-012-0251-5

    CrossRef   Google Scholar

    [112]

    Nhi-Cong LT, Mai CTN, Minh NN, Ha HP, Lien DT, et al. 2016. Degradation of sec-hexylbenzene and its metabolites by a biofilm-forming yeast Trichosporon asahii B1 isolated from oil-contaminated sediments in Quangninh coastal zone, Vietnam. Journal of Environmental Science and Health, Part A 51(3):267−75

    doi: 10.1080/10934529.2015.1094351

    CrossRef   Google Scholar

    [113]

    Vidya P, Sebastian CD. 2022. Yeast diversity in the mangrove sediments of North Kerala, India. European Journal of Biology 81(1):50−57

    doi: 10.26650/EURJBIOL.2022.1027475

    CrossRef   Google Scholar

    [114]

    Ilyas S, Rehman A. 2018. Metal resistance and uptake by Trichosporon asahii and Pichia kudriavzevii isolated from industrial effluents. Archives of Environmental Protection 44(3):77−84

    doi: 10.24425/aep.2018.122291

    CrossRef   Google Scholar

    [115]

    Elahi A, Rehman A. 2017. Oxidative stress, chromium-resistance and uptake by fungi: isolated from industrial wastewater. Brazilian Archives of Biology and Technology 60:e17160394

    doi: 10.1590/1678-4324-2017160394

    CrossRef   Google Scholar

    [116]

    Ilyas S, Rehman A, Varela AC, Sheehan D. 2014. Redox proteomics changes in the fungal pathogen Trichosporon asahii on arsenic exposure: identification of protein responses to metal-induced oxidative stress in an environmentally-sampled isolate. PloS One 9(7):e102340

    doi: 10.1371/journal.pone.0102340

    CrossRef   Google Scholar

    [117]

    Ezzouhri L, Castro E, Moya M, Espínola F, Lairini K. 2009. Heavy metal tolerance of filamentous fungi isolated from polluted sites in Tangier Morocco. African Journal of Microbiology Research 3(2):35−48

    Google Scholar

    [118]

    Iram S, Zaman A, Iqbal Z, Shabbir R. 2013. Heavy metal tolerance of fungus isolated from soil contaminated with sewage and industrial wastewater. Polish Journal of Environmental Studies 22(3):691−97

    Google Scholar

    [119]

    Mohammadian E, Babai Ahari A, Arzanlou M, Oustan S, Khazaei SH. 2017. Tolerance to heavy metals in filamentous fungi isolated from contaminated mining soils in the Zanjan Province, Iran. Chemosphere 185:290−96

    doi: 10.1016/j.chemosphere.2017.07.022

    CrossRef   Google Scholar

    [120]

    Iram S, Parveen K, Usman J, Nasir K, Akhtar N, et al. 2012. Heavy metal tolerance of filamentous fungal strains isolated from soil irrigated with industrial wastewater. Biologija 58(3):107−16

    doi: 10.6001/biologija.v58i3.2527

    CrossRef   Google Scholar

    [121]

    Qayyum S, Khan I, Maqbool F, Zhao Y, Gu Q, et al. 2016. Isolation and characterization of heavy metal resistant fungal isolates from Industrial soil, China. Pakistan Journal of Zoology 48(5):1241−47

    Google Scholar

    [122]

    Zafar S, Aqil F, Ahmad I. 2007. Metal tolerance and biosorption potential of filamentous fungi isolated from metal contaminated agricultural soil. Bioresource Technology 98(13):2557−61

    doi: 10.1016/j.biortech.2006.09.051

    CrossRef   Google Scholar

    [123]

    Muñoz AJ, Ruiz E, Abriouel H, Gálvez A, Ezzouhri L, et al. 2012. Heavy metal tolerance of microorganisms isolated from wastewaters: Identification and evaluation of its potential for biosorption. Chemical Engineering Journal, 210:325−32

    doi: 10.1016/j.cej.2012.09.007

    CrossRef   Google Scholar

    [124]

    Taberna HS, Nillos MG, Gamarcha L, Pahila IG, Gastar AN. 2014. Analysis of road-deposited sediments for heavy metal pollutants in bridge sidewalks of Iloilo City, Philippines. Advances in Environmental Sciences-International Journal of the Bioflux Society 6(1):69−75

    Google Scholar

    [125]

    Sarinas BGS, Gellada LD, Jamolangue EB, Teruñez MR, et al. 2014. Assessment of heavy metals in sediments of Iloilo Batiano River, Philippines. International Journal of Environmental Science and Development 5(6):543−46

    doi: 10.7763/ijesd.2014.v5.542

    CrossRef   Google Scholar

    [126]

    Borlongan I, Golez N, Lorque F. 2010. Physico-chemical assessment of the Jalaur River system, Iloilo, Philippines. Siliman Journal 51(1):224−46

    Google Scholar

    [127]

    Sarinas BGS, Gellada L, Alfonsa JK, Domiquel K, Gumawa LR, et al. 2014. Heavy metal concentration in seawater at Villa Beach, Iloilo City, Philippines. IAMURE International Journal of Ecology and Conservation 11(1):41−53

    doi: 10.7718/ijec.v11i1.806

    CrossRef   Google Scholar

    [128]

    Birch G, Taylor S. 1999. Source of heavy metals in sediments of the Port Jackson estuary, Australia. Science of the Total Environment 227(2−3):123−38

    doi: 10.1016/s0048-9697(99)00007-8

    CrossRef   Google Scholar

    [129]

    Abdollahi S, Raoufi Z, Faghiri I, Savari A, Nikpour Y, et al. 2013. Contamination levels and spatial distributions of heavy metals and PAHs in surface sediment of Imam Khomeini Port, Persian Gulf, Iran. Marine Pollution Bulletin 71(1-2):336−45

    doi: 10.1016/j.marpolbul.2013.01.025

    CrossRef   Google Scholar

    [130]

    Sany S, Salleh A, Sulaiman A, Sasekumar A, Rezayi M, et al. 2012. Heavy metal contamination in water and sediment of the Port Klang coastal area, Selangor, Malaysia. Environmental Earth Sciences 69(6):2013−25

    doi: 10.1007/s12665-012-2038-8

    CrossRef   Google Scholar

    [131]

    Pejman A, Nabi Bidhendi G, Ardestani M, Saeedi M, Baghvand A. 2015. A new index for assessing heavy metals contamination in sediments: a case study. Ecological Indicators 58:365−73

    doi: 10.1016/J.ECOLIND.2015.06.012

    CrossRef   Google Scholar

    [132]

    Huang Z, Zhao W, Xu T, Zheng B, Yin D. 2019. Occurrence and distribution of antibiotic resistance genes in the water and sediments of Qingcaosha Reservoir, Shanghai, China. Environmental Sciences Europe 31(1):81

    doi: 10.1186/s12302-019-0265-2

    CrossRef   Google Scholar

    [133]

    Hauer C, Leitner P, Unfer G, Pulg U, Habersack H, et al. 2018. The role of sediment and sediment dynamics in the aquatic environment. In Riverine Ecosystem Management. Aquatic Ecology Series, eds. Schmutz S, Sendzimir J. Vol 8. Cham: Springer . pp 151–69. doi: 10.1007/978-3-319-73250-3_8

    [134]

    Shyleshchandran MN, Mohan M, Ramasamy EV. 2018. Risk assessment of heavy metals in Vembanad Lake sediments (south-west coast of India), based on acid-volatile sulfide (AVS)-simultaneously extracted metal (SEM) approach. Environmental Science and Pollution Research 25(8):7333−45

    doi: 10.1007/s11356-017-0997-8

    CrossRef   Google Scholar

    [135]

    Liu M, Zhong J, Zheng X, Yu J, Liu D, et al. 2018. Fraction distribution and leaching behavior of heavy metals in dredged sediment disposal sites around Meiliang Bay, Lake Taihu (China). Environmental Science and Pollution Research 25(10):9737−9744

    doi: 10.1007/S11356-018-1249-2

    CrossRef   Google Scholar

    [136]

    Chen CY, Wu ZC, Liu TY, Yu SS, Tsai JN, et al. 2023. Investigation of asymptomatic infection of Phellinus noxius in herbaceous plants. Phytopathology 113(3):460−69

    doi: 10.1094/PHYTO-08-22-0281-R

    CrossRef   Google Scholar

    [137]

    Liu M, Chen J, Sun X, Hu Z, Fan D. 2019. Accumulation and transformation of heavy metals in surface sediments from the Yangtze River estuary to the East China Sea shelf. Environmental Pollution 245:111−21

    doi: 10.1016/J.ENVPOL.2018.10.128

    CrossRef   Google Scholar

    [138]

    Idris AM, Eltayeb MAH, Potgieter-Vermaak SS, Van Grieken R, Potgieter JH. 2007. Assessment of heavy metals pollution in Sudanese harbours along the Red Sea Coast. Microchemical Journal 87(2):104−12

    doi: 10.1016/J.MICROC.2007.06.004

    CrossRef   Google Scholar

    [139]

    Salati S, Moore F. 2010. Assessment of heavy metal concentration in the Khoshk River water and sediment, Shiraz, Southwest Iran. Environmental monitoring and assessment 164:677−89

    doi: 10.1007/s10661-009-0920-y

    CrossRef   Google Scholar

    [140]

    Jakimska A, Konieczka P, Skóra K, Namieśnik J. 2011. Bioaccumulation of metals in tissues of marine animals, Part II: metal concentrations in animal tissues. Polish Journal of Environmental Studies 20(5):1127−46

    Google Scholar

    [141]

    Takarina V, Adiwibowo A. 2011. Impact of heavy metals contamination on the biodiversity of marine benthic organisms in Jakarta Bay. Journal of Coastal Development 14(2):168−171

    Google Scholar

    [142]

    Liu J, Wang J, Gao G, Bartlam MG, Wang Y. 2015. Distribution and diversity of fungi in freshwater sediments on a river catchment scale. Frontiers in Microbiology 6:329

    doi: 10.3389/fmicb.2015.00329

    CrossRef   Google Scholar

    [143]

    Shang Y, Wu X, Wang X, Dou H, Wei Q, et al. 2022. Environmental factors and stochasticity affect the fungal community structures in the water and sediments of Hulun Lake, China. Ecology and Evolution 12(11):e9510

    doi: 10.1002/ece3.9510

    CrossRef   Google Scholar

    [144]

    Dockrey J, Lindsay M, Mayer K, Beckie R, Norlund K, et al. 2014. Acidic microenvironments in waste rock characterized by neutral drainage: Bacteria–mineral interactions at sulfide surfaces. Minerals 4(1):170−90

    doi: 10.3390/min4010170

    CrossRef   Google Scholar

    [145]

    Manguilimotan L, Bitacura J. 2018. Biosorption of cadmium by filamentous fungi isolated from coastal water and sediments. Journal of Toxicology 2018:170510

    doi: 10.1155/2018/7170510

    CrossRef   Google Scholar

    [146]

    Passarini MRZ, Ottoni JR, Costa PEDS, Hissa DC, Falcão RM, et al. 2022. Fungal community diversity of heavy metal contaminated soils revealed by metagenomics. Archives of Microbiology 204(5):255

    doi: 10.1007/s00203-022-02860-7

    CrossRef   Google Scholar

    [147]

    Bonugli-Santos RC, Durrant LR, Sette LD. 2020. Analysis of fungal composition in mine-contaminated soils in Hechi City. Current Microbiology 77(10):2685−93

    doi: 10.1007/s00284-020-02044-w

    CrossRef   Google Scholar

    [148]

    Bonugli-Santos R, Durrant L, Sette L. 2012. The production of ligninolytic enzymes by marine-derived Basidiomycetes and their biotechnological potential in the biodegradation of recalcitrant pollutants and the treatment of textile effluents. Water, Air, and Soil Pollution 223(5):2333−45

    doi: 10.1007/s11270-011-1027-y

    CrossRef   Google Scholar

    [149]

    Cui Z, Zhang X, Yang H, Sun L. 2017. Bioremediation of heavy metal pollution utilizing composite microbial agent of Mucor circinelloides, Actinomucor sp. , and Mortierella sp. Journal of Environmental Chemical Engineering 5(4):3616−21

    doi: 10.1016/j.jece.2017.07.021

    CrossRef   Google Scholar

    [150]

    Lin Y, Xiao W, Ye Y, Wu C, Hu Y, et al. 2020. Adaptation of soil fungi to heavy metal contamination in paddy fields—a case study in eastern China. Environmental Science and Pollution Research International 27(22):27819−30

    doi: 10.1007/s11356-020-09049-9

    CrossRef   Google Scholar

    [151]

    Li J, Zheng Q, Liu J, Pei S, Yang Z, et al. 2024. Bacterial–fungal interactions and response to heavy metal contamination of soil in agricultural areas. Frontiers in Microbiology 15:1395154

    doi: 10.3389/fmicb.2024.1395154

    CrossRef   Google Scholar

    [152]

    Lin Y, Ye Y, Hu Y, Shi H. 2019. The variation in microbial community structure under different heavy metal contamination levels in paddy soils. Ecotoxicology and Environmental Safety 180:557−64

    doi: 10.1016/j.ecoenv.2019.05.057

    CrossRef   Google Scholar

    [153]

    Chandra P, Enespa. 2019. Mycoremediation of environmental pollutants from contaminated soil. In Mycorrhizosphere and Pedogenesis, eds Varma A, Choudhary D. Singapore: Springer. pp 239–74. doi: 10.1007/978-981-13-6480-8_15

    [154]

    Singh A, Gauba P. 2014. Mycoremediation: A treatment for heavy metal pollution of soil. Journal of Civil Engineering and Environmental Technology 1(4):59−61

    Google Scholar

    [155]

    Dashtban M, Schraft H, Syed TA, Qin W. 2010. Fungal biodegradation and enzymatic modification of lignin. International Journal of Biochemistry and Molecular Biology 1(1):36

    Google Scholar

    [156]

    Thippeswamy B, Shivakumar CK, Krishnappa M. 2012. Bioaccumulation potential of Aspergillus niger and Aspergillus flavus for removal of heavy metals from paper mill effluent. Journal of Environmental Biology 33(6):1063−68

    Google Scholar

    [157]

    Tripathi P, Khare P, Barnawal D, Shanker K, Srivastava PK, et al. 2020. Bioremediation of arsenic by soil methylating fungi: Role of Humicola sp. strain 2WS1 in amelioration of arsenic phytotoxicity in Bacopa monnieri L. Science of the Total Environment 716:136758

    doi: 10.1016/j.scitotenv.2020.136758

    CrossRef   Google Scholar

  • Cite this article

    Dela Cruz FM, Bermeo-Capunong MRA, Bagacay JFE, Canto CM, Calabon MS. 2025. Taxonomy, phylogeny, and preliminary screening of fungal isolates for cadmium tolerance from Iloilo Ferry Terminal, Iloilo, Philippines. Studies in Fungi 10: e001 doi: 10.48130/sif-0025-0001
    Dela Cruz FM, Bermeo-Capunong MRA, Bagacay JFE, Canto CM, Calabon MS. 2025. Taxonomy, phylogeny, and preliminary screening of fungal isolates for cadmium tolerance from Iloilo Ferry Terminal, Iloilo, Philippines. Studies in Fungi 10: e001 doi: 10.48130/sif-0025-0001

Figures(7)  /  Tables(1)

Article Metrics

Article views(283) PDF downloads(119)

ARTICLE   Open Access    

Taxonomy, phylogeny, and preliminary screening of fungal isolates for cadmium tolerance from Iloilo Ferry Terminal, Iloilo, Philippines

Studies in Fungi  10 Article number: e001  (2025)  |  Cite this article

Abstract: Water pollution by heavy metals poses a growing threat to ecosystems and human health. Bioremediation using fungi offers a promising solution due to their tolerance and absorption capabilities. This study aimed to isolate and identify cadmium-tolerant fungi from the Iloilo River, a potentially polluted river in Iloilo City, Philippines using morpho-phylogenetic analysis. Water and sediment samples were collected, and fungi were isolated by serial dilution on PDA plates supplemented with cadmium ions [Cd(II)] (25 ppm). Isolates were further tested for tolerance to varying cadmium concentrations (50 and 100 ppm). Nine fungal species were identified, including three ascomycetous taxa (Pseudopestalotiopsis elaeidis, Peroneutypa scoparia, Diaporthe sp.) and six basidiomycetous taxa (Cerrena sp., Irpex laceratus, Phlebia sp., Pyrrhoderma noxium, Schizophyllum commune, Trichosporon asahii). The majority of the isolates originated from sediment samples. Morphological illustrations, descriptions, and taxonomic notes are provided for each taxon. These findings provide insight into the fungal diversity of the Iloilo River and their tolerance to cadmium.

    • The global population boom in the 21st century has led to a significant increase in water consumption. Unfortunately, this rise has been accompanied by escalating water pollution[1,2]. Heavy metals are among the most worrisome pollutants affecting both aquatic and terrestrial environments[3]. These pollutants persist in water and sediment, posing a continuous threat due to their toxicity, even in low concentrations[3]. The non-biodegradability and low solubility of heavy metals, including their ionic forms such as cadmium, make them difficult to eliminate, allowing them to accumulate in aquatic ecosystems. These ionic forms are readily absorbed by living organisms, where they can induce oxidative stress, disrupt steroidogenesis, impair spermatogenesis, and trigger apoptotic toxicity, ultimately impairing biological processes[46].

      As a result, heavy metals can concentrate within organisms, amplifying as they move up the food chain[7]. This contamination not only reduces biodiversity and ecosystem productivity but also undermines the self-repair abilities of affected water bodies[8]. In addition to the ecological damage, the global economic costs of heavy metal pollution in aquatic systems exceed USD $10 billion annually[9]. Ensuring cleaner water through the removal of these pollutants could preserve biodiversity and yield significant economic benefits[10].

      Traditional methods for removing heavy metals from water bodies, such as physicochemical or electrical treatments, are often costly and inefficient[11,12]. Bioremediation offers a promising alternative, utilizing biological agents to transform or degrade pollutants into less harmful substances[3,13]. Fungi, in particular, are effective bioremediation agents due to their high tolerance and absorption capacity for heavy metals, outperforming bacteria and chemical methods in bioleaching studies[3,14].

      Researchers have found that fungal isolates sourced from polluted environments, such as industrially impacted rivers, tend to exhibit superior bioremediation capabilities compared to lab-cultured fungi[15,16]. This suggests that environmental pressures may enhance the effectiveness of fungi in mitigating heavy metal contamination. Therefore, exploring diverse natural water sources for potential fungal isolates is crucial[17].

      Fungi have shown significant potential as bioremediation agents for environments contaminated with heavy metals due to their remarkable tolerance to toxic substances. They can thrive in extreme conditions, with some species tolerating heavy metal concentrations as high as 400 ppm[18]. Studies have demonstrated that fungi like Aspergillus terreus and Trichoderma viride are capable of effectively removing heavy metals from contaminated media, with A. terreus removing 59.67 mg/g of lead and T. viride removing 16.25 mg/g[18]. Their ability to tolerate and accumulate heavy metals is attributed to their unique structural features, including a rigid cell wall made of chitin, which enhances metal uptake[19]. Fungi also exhibit a high biomass yield and are relatively inexpensive to cultivate, making them an attractive alternative to bacteria for bioremediation. Unlike bacteria, which exhibit a significant decline in activity under high metal concentrations, fungi can maintain or even increase their activity, especially for metals like arsenic and cadmium[20,21]. This increased tolerance is linked to their metabolic differences and high osmotic pressure in their cell walls[19]. Fungi also provide the advantage of a modifiable genetic makeup, which can be optimized for large-scale production and enhanced metal biosorption[22]. Their ability to secrete various metabolites further enhances their potential for treating industrial effluents, making them a promising candidate for the bioremediation of heavy metals[23].

      Rivers, particularly those exposed to industrial and urban waste, are prime sampling sites for identifying fungi capable of bioremediation. The Iloilo River in Iloilo City, Philippines, is an example of such a site, as it is frequently subjected to heavy metal pollution due to its location in a densely populated and industrialized area[24]. Water and sediment samples from this river offer valuable opportunities to study the bioremediation potential of local fungal species in combating heavy metal pollution.

      This study aims to employ ITS rDNA barcoding for the accurate identification of fungal isolates collected from the Iloilo Ferry Terminal, Iloilo, Philippines, and to conduct a preliminary assessment of their tolerance to cadmium exposure. Through the analysis of these isolates, this research seeks to enhance the understanding of fungal roles in bioremediation within aquatic ecosystems impacted by heavy metal contamination.

    • Water and sediment samples were collected at the river mouth of the Iloilo Ferry Terminal Port, Iloilo City, Philippines (Fig. 1). A total of 500 mL of water samples and 500 g of sediment were collected for analysis. Sediment samples were collected using a sterile corer and stored in sterile zip-lock bags to prevent contamination. Water samples were collected in sterile amber bottles to minimize light exposure, thereby reducing the potential degradation of photosensitive compounds and preserving the integrity of any microorganisms present. Both sediment and water samples were transported to the University of the Philippines Visayas - Microbiology Laboratory, Miagao, Iloilo, Philippines. Water samples were maintained at 4 °C during transport using a cooler to minimize any physicochemical changes that could affect the analysis.

      Figure 1. 

      Map showing the sampling sites (yellow circle) in Iloilo-Guimaras Ferry Terminal Port, Iloilo City, Iloilo, Philippines. The map was created using Google Earth Pro.

    • Three different concentrations (25, 50, and 100 ppm) were prepared for stock solutions of cadmium [Cd(II)]. Cadmium nitrate tetrahydrate was used as the cadmium source. A 1,000 ppm stock solution of cadmium was prepared by dissolving 2.774 g of cadmium nitrate tetrahydrate in 1 L of distilled water. The prepared 1000 ppm stock solution was used to create the desired working concentrations (25, 50, and 100 ppm) using the dilution formula: C1V1 = C2V2 (where C1 and V1 are initial concentration and volume before dilution, respectively; C2 and V2 are final concentration and volume after dilution, respectively). The diluted solutions were sterilized using 0.22 μm filters with a vacuum filtration setup consisting of a vacuum filter flask and a vacuum pump.

      Potato dextrose agar (PDA), supplemented with Pen-Strep, was prepared following the manufacturer's instructions. To achieve a final concentration of 25 ppm cadmium in the medium, a 4 mL cadmium (25 ppm) solution was added per 1 L of cooled, molten PDA. This concentration was used for the initial isolation of fungi tolerant to cadmium.

    • Water and sediment samples collected from the port were serially diluted up to 10−6. Aliquots of 0.1 mL from each dilution were then spread onto potato dextrose agar (PDA) supplemented with Cd(II) at concentrations of 25 ppm using a spread plate technique. Each dilution was plated in triplicate to minimize random error. Petri plates were incubated for a week at 28 ± 2 °C and observed daily for fungal growth. Unique fungal colonies were purified and characterized using both colonial characteristics and micromorphological features. Colonial traits include color, texture, elevation, margin, and characteristics of aerial hyphae. For microscopic examination, the slide culture technique was employed to observe traits such as spore morphology, shape, and conidial formation, which were analyzed based on the cultural characteristics of media supplemented with 20 ppm Cd(II)[25].

    • All fungal isolates were assessed for their tolerance to cadmium [Cd(II)] by culturing them on PDA media supplemented with Cd(II) at higher concentrations (i.e., 50 and 100 ppm). For comparison, fungal isolates grown on PDA medium without Cd(II) served as controls. All cultures were incubated for 5 d at a temperature of 28 ± 2 °C. The growth of each fungal isolate was visually assessed and recorded based on relative growth compared to the control and categorized as follows: ++++ (normal growth), +++ (slightly less than normal), ++(less than normal), + (least growth) and – (absent growth)[15].

    • Fungal mycelia from pure cultures grown in PDA for 15 d were scraped using a sterilized scalpel and kept in a sterilized 2 mL microcentrifuge tube. Genomic DNA was extracted using the DNeasy Blood & Tissue Kit (Qiagen) following the manufacturer's protocol. Polymerase chain reaction (PCR) was used to amplify the internal transcribed spacers (ITS) of rDNA using the primers ITS4 and ITS5[26]. Polymerase chain reaction was performed in a volume of 50 μl, which contained 5 μL 10 XPCR buffer, 1 μL of forward primer 10 μM, 1μL reverse primer 10 μM, 1 μL of dNTP, 0.25 μL of Hotstart DNA polymerase, 36.75 μL nuclease-free water, and 5 μL of genomic DNA (5 ng/μL). The PCR thermal cycle program for ITS amplification was as follows: an initial denaturing step of 95 °C for 15 min, followed by 35 cycles of denaturation at 94 °C for 1 min, annealing at 60°C for 30 s, elongation at 72 °C for 1 min, and final extension at 72 °C for 10 min. Agarose gel electrophoresis was carried out to confirm the presence of amplicons at the expected molecular weight. PCR products were purified and sequenced with the primers mentioned above at the Philippine Genome Center – Visayas, University of the Philippines Visayas, Miagao, Iloilo, Philippines.

    • Taxonomic classifications were established based on the closest matches obtained from nucleotide BLASTn searches. The ITS sequence data was analyzed in conjunction with additional sequences retrieved from GenBank. Phylogenetic analysis was conducted using maximum likelihood (ML) and Bayesian inference (BI) methods. ML analyses were carried out with IQ-TREE webserver (www.hiv.lanl.gov/content/sequence/IQTREE/iqtree.html), while BI analyses were performed with MrBayes v.3.2.6 on XSEDE[27], accessed through the CIPRES web portal (www.phylo.org/portal2)[28], following the procedures of Dissanayake et al.[29]. The resulting phylogenetic trees were visualized with FigTree v. 1.4.4, and the final layout was created in Microsoft PowerPoint 365. Newly generated sequences have been deposited in GenBank.

    • Of the colonies exhibiting tolerance to 25 ppm cadmium supplementation, 15 displayed distinct morphological characteristics. Among these, nine isolates were successfully sequenced and became the primary focus of this study. Table 1 provides detailed information on their taxonomic identities, cadmium tolerance levels, and the original sample sources from which they were isolated. Six isolates (Pseudopestalotiopsis elaeidis, Peroneutypa scoparia, Diaporthe sp., Phlebia sp., Pyrrhoderma noxium, and Trichosporon asahii) from sediment exhibited tolerance to both 50 and 100 ppm cadmium concentrations, indicating their resilience to heavy metal stress. In contrast, water isolates displayed mixed results, with only Trichosporon asahii showing tolerance to both cadmium concentrations. The remaining water isolates (Cerrena sp., Irpex laceratus, and Schizophyllum commune) were not tolerant to 50 and 100 ppm cadmium. This suggests that sediment-derived fungi generally demonstrate a higher tolerance to cadmium contamination compared to those from water sources.

      Table 1.  Tolerance of isolates to 50 ppm and 100 ppm cadmium supplementation with sample sources.

      Taxon Source Growth*
      25 ppm 50 ppm 100 ppm
      Pseudopestalotiopsis elaeidis Sediment ++++ ++++ ++++
      Peroneutypa scoparia Sediment ++++ ++++ ++++
      Cerrena sp. Water +++
      Diaporthe sp. Sediment ++++ ++++ ++++
      Phlebia sp. Sediment ++++ ++++ ++++
      Irpex laceratus Water +++
      Schizophyllum commune Water +++
      Pyrrhoderma noxium Sediment ++++ ++++ ++++
      Trichosporon asahii Water ++++ ++++ ++++
      * ++++ (normal growth); +++ (slightly less than normal); ++ (less than normal); + (least growth); – (absent growth).
    • The ITS gene dataset comprised 75 taxa from Ascomycota, with Parasympodiella elongata (CBS 522.93) and P. eucalypti (CBS 124767) as the outgroup taxa (Fig. 2). The analyzed dataset, after trimming, comprised a total of 531 characters. Alignment has 75 sequences with 531 columns, 321 distinct patterns, 268 parsimony-informative, 31 singleton sites, and 232 constant sites. The ML analysis for the combined dataset provided the best scoring tree (Fig. 2) with a final ML optimization likelihood value of -5305.142 (ln). Parameters for the GTR model of the combined ITS dataset are as follows: estimated base frequencies; A = 0.250, C = 0.250, G = 0.250, T = 0.250; substitution rates AC = 1.00000, AG = 1.60667, AT = 1.00000, CG = 1.00000, CT = 2.99928, GT = 1.00000; gamma distribution shape parameter α = 0.346. Support values for maximum likelihood (ML) above 75% are given at the nodes.

      Figure 2. 

      Phylogram generated from maximum likelihood analysis based on ITS sequence data of Ascomycota taxa. Isolates from this study are highlighted in blue font. Ex-type strains are in bold.

      For Basidiomycota, the ITS gene dataset comprised of 242 taxa, with Ustilago abaconensis (CBS 8380) and U. maydis (CBS 380.32) as the outgroup taxa (Fig. 3). The analyzed dataset, after trimming, comprised a total of 1,725 characters. Alignment has 242 sequences with 1,725 columns, 928 distinct patterns, 693 parsimony-informative, 192 singleton sites, and 840 constant sites. The ML analysis for the combined dataset provided the best scoring tree (Fig. 3) with a final ML optimization likelihood value of –22,547.577 (ln). Parameters for the GTR model of the combined ITS dataset are as follows: estimated base frequencies; A = 0.263, C = 0.206, G = 0.227, T = 0.305; substitution rates AC = 1.58477, AG = 2.87131, AT = 1.58477, CG = 1.00000, CT = 4.26806, GT = 1.00000; gamma distribution shape parameter α = 0.611. Support values for maximum likelihood (ML) above 75% are given at the nodes.

      Figure 3. 

      Phylogram generated from maximum likelihood analysis based on ITS sequence data of Basidiomycota taxa. Isolates from this study are highlighted in blue font. Ex-type strains are in bold.

    • Ascomycota Caval.-Sm., Biol. Rev. 73: 247 (1998)

      Sordariomycetes O.E. Erikss. & Winka, Myconet 1(1): 10 (1997)

      Amphisphaeriales D. Hawksw. & O.E. Erikss., Syst. Ascom. 5(1): 177 (1986)A

      Pestalotiopsidaceae Maharachch. & K.D. Hyde, Fungal Diversity 73: 106 (2015)

      Pseudopestalotiopsis Maharachch., K.D. Hyde & Crous, Stud. Mycol. 79: 180 (2014)

      Pseudopestalotiopsis elaeidis (C. Booth & J.S. Robertson) F. Liu, L. Cai & Crous, Stud. Mycol. 92: 374 (2018) [2019], Fig. 4ac

      Figure 4. 

      Colonial morphology on 25 ppm Cd (II). Pseudopestalotiopsis elaeidis UPVMICC 23-010. Colony on PDA from (a) surface, and (b) reverse. (c) Septate hyphae. Peroneutypa scoparia UPVMICC 23-015. Colony on PDA from (d) surface, and (e) reverse. (f) Septate hyphae. Diaporthe sp. UPVMICC 23-017. Colony on PDA from (g) surface, and (h) reverse. (i) Septate hyphae. Scale bars: (c), (f), (i) = 25 μm.

      Description: Colonies on PDA reaching 20–25 mm diameter after 7 d at 25 °C, colonies circular, margin entire, flat, velvety to sparse cottony appearance, colony from surface: initially white, became greyish white; reverse: verrucose, yellow. No sporulation even on prolonged incubation (6 months). Hyphae 1.69–3.63 μm ($\bar {\rm x} $ = 2.67 μm), hyaline, septate.

      Material examined: PHILIPPINES, Iloilo, Iloilo City, Iloilo Ferry Terminal Port, on sediment samples, 31 October 2023, F.M. Dela Cruz, SHD-633, culture UPVMICC 23-010.

      GenBank number: PQ304372 (ITS)

      Notes: Pseudopestalotiopsis elaeidis was introduced by Liu et al.[30] based on their phylogenetic analysis and comparison of nucleotide differences of the ex-type strains of Leptosphaeria elaeidis (CBS 413.62) and P. myanmarina (NBRC 112264). Based on nomenclatural priority, P. myanmarina was synonymized as P. elaeidis. The taxon was reported in China, Indonesia, Myanmar, and Nigeria on various hosts[3032]. In the phylogenetic analysis, UPVMICC 23-010 clade with other strains of P. elaeidis (NBRC 112260; NBRC 112264; NBRC 112265; NBRC 112269) with high bootstrap support (100% ML, 1.00 BYPP) (Fig. 2). This is the first country report of P. elaeidis in the Philippine aquatic habitats. The isolate was able to grow in all 25, 50, and 100 ppm cadmium and thus can potentially be applied for cadmium bioremediation.

      No other studies were found that report heavy metal tolerance or bioremediation applications for this species, making this the first study to document its tolerance to heavy metals.

      Xylariales Nannf., Nova Acta R. Soc. Scient. upsal., Ser. 4 8(no. 2): 66 (1932)

      Diatrypaceae Nitschke [as 'Diatrypeae'], Verh. naturh. Ver. preuss. Rheinl. 26: 73 (1869)

      Peroneutypa Berl., Icon. fung. (Abellini) 3(3-4): 80 (1902)

      Peroneutypa scoparia (Schwein.) Carmarán & A.I. Romero, Fungal Diversity 23: 84 (2006), Fig. 4df

      Description: Colonies on PDA reaching 27–29 mm diameter after 7 d at 25 °C, colonies circular, margin entire, flat, cottony, effuse, fluffy but are flat and dense, colony from surface: denser mycelia at the center, initially white, with grey pigmentations appearing at the center when mature; reverse: white with grey at center. No sporulation even on prolonged incubation (6 months). Hyphae 1.34–4.27 μm ($\bar {\rm x} $ = 2.23 μm), hyaline.

      Material examined: PHILIPPINES, Iloilo, Iloilo City, Iloilo Ferry Terminal Port, on sediment samples, 31 October 2023, F.M. Dela Cruz, SHD-432, culture UPVMICC 23-015.

      Notes: Carmarán et al.[33] resurrected the genus Peroneutypa based on morphological and phylogenetic analyses of Diatrypaceae, transferring Sphaeria scoparia to Peroneutypa. Peroneutypa scoparia is a known pathogen causing diseases on various plant hosts[3440]. Additionally, the taxon has been reported as saprobic[33,41], and endophytic[42] in nature. The isolated P. scoparia UPVMICC 23-015 represents the first record of this species in the Philippine aquatic habitats. Phylogenetic analysis revealed that it clustered with other documented P. scoparia species with strong bootstrap support (99% ML, 1.00 BYPP) (Fig. 2). Growth in 25, 50, and 100 ppm cadmium was observed, which suggests potential application for cadmium bioremediation.

      There is limited documentation on the use of P. scoparia as a bioremediation agent, with most studies focusing on its enzyme-based remediation of dyes, particularly using laccase. While no applications for heavy metal remediation have been documented, there is considerable potential given that the enzymes employed for dye remediation have demonstrated effectiveness against other toxic substances, including heavy metals[43,44].

      Diaporthales Nannf., Nova Acta R. Soc. Scient. upsal., Ser. 4 8 (no. 2): 53 (1932)

      Diaporthaceae Höhn. ex Wehm., Am. J. Bot. 13: 638 (1926)

      Diaporthe Nitschke, Pyrenomyc. Germ. 2: 240 (1870)

      Diaporthe sp., Fig. 4gi

      Description: Colony on PDA has a diameter ranging from 25–27 mm after 7 d at 25 °C incubation, dense and velvety, white on the surface view, similar to its reverse side. Chlamydospores 4.97–6.58 μm ($\bar {\rm x} $ = 5.93 μm) present. Hyphae, 2.30–5.13 μm ($\bar {\rm x} $ = 3.60 μm), hyaline, septate.

      Material examined: PHILIPPINES, Iloilo, Iloilo City, Iloilo Ferry Terminal Port, on sediment samples, 31 October 2023, F.M. Dela Cruz, SHD-431, culture UPVMICC 23-017.

      Distribution: Diaporthe is a genus with a wide geographical distribution, encompassing pathogens, endophytes, and saprobes that infect diverse hosts[45,46]. It has been associated with economically important crops like citrus and coffee, as well as ornamental plants such as Clematis[4749]. Phylogenetic analyses placed this taxon within the Diaporthe sojae species complex[47] and clustered with other unidentified Diaporthe spp. (99% ML, 0.97 BYPP) (Fig. 2). Growth in 25, 50, and 100 ppm cadmium was also observed, which suggests potential application for cadmium bioremediation.

      Basidiomycota R.T. Moore, Bot. Mar. 23(6): 371 (1980)

      Agaricomycetes Doweld, Prosyllabus Tracheophytorum, Tentamen Systematis Plantarum Vascularium (Tracheophyta) (Moscow): LXXVIII (2001)

      Polyporales Gäum., Vergl. Morph. Pilze (Jena): 503 (1926)

      Cerrenaceae Miettinen, Justo & Hibbett, Fungal Biology 121(9): 817 (2017)

      Cerrena Gray, Nat. Arr. Brit. Pl. (London) 1: 649 (1821)

      Cerrena sp., Fig. 5ac

      Figure 5. 

      Colonial morphology on 25 ppm Cd (II). Cerrena sp. UPVMICC 23-008. Colony on PDA from (a) surface, and (b) reverse. (c) Hyphae. Irpex laceratus UPVMICC 23-007. Colony on PDA from (d) surface, and (e) reverse. (f) Hyphae. Phlebia sp. UPVMICC 23-009. Colony on PDA from (g) surface, and (h) reverse. (i) Hyphae. Schizophyllum commune UPVMICC 23-012. Colony on PDA from (g) surface, and (h) reverse. (i) Hyphae. Scale bars: (c), (f), (i), (l) = 25 μm.

      Description: Colony on PDA reaching a diameter of 35 to 36 mm after 7 d at 25 °C, white (surface and obverse), cottony, and round; Mycelia fluffy, short, dense; septate hyphae 3.15–4.76 μm ($\bar {\rm x} $ = 3.37 μm); binding hyphae intercrossed with each other, creating clumps, dimitic, or trimitic systems.

      Material examined: PHILIPPINES, Iloilo, Iloilo City, Iloilo Ferry Terminal Port, on water samples, 31 October 2023, F.M. Dela Cruz, WHD-431, culture UPVMICC 23-008.

      Notes: Cerrena species are widely distributed across Europe, Africa, and South America[5052]. Known for its potent wood-decaying ability, this fungus commonly parasitizes living trees before adopting a saprobic lifestyle on dead wood[53]. The isolate clustered with other unidentified Cerrena strains [e.g., NTOU5117; NTOU 4204; 7-SU-3-B-77(M)-B; 6-L-3-C-32(M)][5456] with strong bootstrap support (100% ML, 1.00 BYPP) (Fig. 3). However, the presence of Cordyceps militans (CH9) within this cluster raises the possibility of misidentification and must be correctly identified as Cerrena sp. In the phylogenetic analysis, given its distinct lineage compared to established Cerrena taxa (i.e., C. albocinnamomea, C. cystidiata, C. gilbertsonii, C. multipileata, C. unicolor, and C. zonata), this strain might represent a novel species and warrants further investigation through sequencing of its protein-coding regions. In aquatic environments, the genus Cerrena was isolated from decaying mangrove wood[57]. This is the first report of Cerrena isolated from brackish water. Growth was only observed on 25 ppm cadmium and not on higher concentrations (50 and 100 ppm).

      Cerrena is a white rot fungus that possesses the same enzyme as Peroneutypa scoparia, namely the laccase enzyme, and has been reported to play a role in environmental contaminant biodegradation[58]. This enzyme has been associated with the tolerance of fungal species to heavy metals, as it exhibits broad-spectrum activity against various contaminants, including heavy metals[59]. Furthermore, this enzyme is known for its ability to decolorize and detoxify industrial effluents, aiding in wastewater treatment[60]. It has been observed to degrade a multitude of contaminants, including heavy metals[61]. Interestingly, the promoter regions of the laccase genes harbor distinct recognition sites tailored for xenobiotics and heavy metals, making heavy metals significant activators of this enzyme[62]. When present in the medium, xenobiotics and heavy metals can attach to the gene's recognition sites, thereby triggering the production of laccase[60]. For example, when Cerrena sp. is exposed to Cu2+ and Zn2+, increased laccase production is observed[63]. Cadmium has also been observed to specifically increase laccase activity and is considered an inducer of the enzyme[6466].

      Irpicaceae Spirin & Zmitr., in Spirin, Mycena 3: 48 (2003)

      Irpex Fr., Syst. orb. veg. (Lundae): 81 (1825)

      Irpex laceratus (N. Maek., Suhara & R. Kondo) C.C. Chen & Sheng H. Wu, Fungal Diversity 111: 411 (2021), Fig. 5df

      Description: Colony on PDA has a diameter that ranges 45–48 mm after 7 d of incubation at 25 °C, white (surface and obverse), velvety, evenly flat, and translucent with some areas having moderate fluffiness while others seem to be waxy. Hyphae 2.94–4.23 μm ($\bar {\rm x} $ = 3.71 μm), hyaline, rough, monomitic, coenocytic.

      Material examined: PHILIPPINES, Iloilo, Iloilo City, Iloilo Ferry Terminal Port, on water samples, 31 October 2023, F.M. Dela Cruz, WHD-622, culture UPVMICC 23-007.

      Distribution: Irpex laceratus is typically found on angiosperms and exhibits a vast distribution across Southeast and East Asia[6769]. Additionally, this fungus contributes to white rot in wood and has a global presence, with its primary concentration in the boreal region of the Northern Hemisphere[70]. The new isolate clustered with other I. laceratus strains (BRPET15, ATCC42010, Y.H. Yeh l1215, J5-2, hlb5_4_1) (Fig. 3) with strong statistical support (98% ML, 1.00 BYPP). Interestingly, it also grouped with I. latemarginatus with moderate support (93% ML, 1.00 BYPP). While traditionally known as a terrestrial fungus, I. laceratus has been isolated from aquatic environments in Acanthaster planci[71] and Ostrea denselamellosa[72], demonstrating bioactivity. It is considered a marine fungus by Jones et al.[73] and Calabon et al.[74]. This is the first report of I. laceratus isolated from brackish water in the Philippines. Growth was observed on 25 ppm cadmium, while no growth was observed on both 50 and 100 ppm.

      The only mention of I. laceratus in relation to bioremediation pertains to its wood-degrading capabilities, with no references to its effectiveness in heavy metal remediation[75].

      Meruliaceae Rea, Brit. basidiomyc. (Cambridge): 620 (1922)

      Phlebia Fr., Syst. mycol. (Lundae) 1: 426 (1821)

      Phlebia sp., Fig. 5gi

      Description: The colony on PDA reaches a diameter of 48–49 mm after 7 d at 25 °C, white, dense, and cottony center on the obverse side, as the colony grows outwards, the cottony texture becomes more flat and yellow coloration of the agar occurs. Hyphae 2.94– 4.23 μm ($\bar {\rm x} $ = 3.71 μm), hyaline, dimitic.

      Material examined: PHILIPPINES, Iloilo, Iloilo City, Iloilo Ferry Terminal Port, on sediment samples, 31 October 2023, F.M. Dela Cruz, SHD-632, culture UPVMICC 23-009.

      Notes: Species of the genus Phlebia are widely distributed across North America and Europe, with additional records from New Zealand[76]. The Phlebia sp. isolated in this study clustered with uncultured fungal clones (L042881-122-061-B10, L042881-122-061-B11, L042881-122-061-B12, and L042886-122-066-F07), two unidentified Phlebia species (MG60, MGEF50), and a fungal endophyte, GFM3, with robust statistical support (100% ML, 0.97 BYPP) (Fig. 3). Notably, this clade also contained Aspergillus versicolor, a clear case of misidentification. This taxon was included in the phylogenetic tree based on BLASTn results. Phylogenetic analysis revealed that Phlebia sp. UPVMICC 23-009 forms a distinct lineage separate from other known Phlebia species (P. floridensis and P. radiata), suggesting it may represent a novel taxon. Further investigation through sequencing of its protein-coding regions is warranted. While Phlebia species have been reported in aquatic environments, such as mangrove stands in Japan, with biodegradation capabilities[77], this study represents the first record of the genus in brackish waters within the Philippines. There was observed growth in all three concentrations of cadmium 25, 50, and 100 ppm and has the potential to be used for the bioremediation of cadmium.

      Phlebia species utilized as bioremediation agents have shown promising results, particularly in the remediation of heavy metals[78]. For instance, it has been observed that Phlebia species remove up to 97.1% of cadmium, 97.5% of lead, and 72.7% of nickel[79]. However, the same study also illustrated changes in the morphology of the fungus when exposed to heavy metals[79]. Similar to Cerrena sp., Phlebia sp. is another example of white rot fungi[80], which are generally renowned for their biodegradability and have recently gained attention for their remarkable results in heavy metal-specific bioremediation[81,82]. Phlebia species is also capable of producing the previously mentioned enzyme laccase[79]. Similarly, laccase production in Phlebia sp. is intensified by the presence of metals, acting as inducers. It is believed that the main mechanism for Phlebia sp. tolerance and potential bioremediation application involves intracellular compartmentalization of metals inside vacuoles and biosorption[79]. Known Phlebia species with capabilities for heavy metal bioremediation and tolerance include P. brevispora, P. floridensis, and P. radiata[79,81].

      Hymenochaetales Oberw., Beitr. Biol. Pfl.: 89 (1977)

      Hymenochaetaceae Donk, Bull. bot. Gdns Buitenz. 17(4): 474 (1948)

      Pyrrhoderma Imazeki, Trans. Mycol. Soc. Japan 7: 4 (1966)

      Pyrrhoderma noxium (Corner) L.W. Zhou & Y.C. Dai, Mycologia 110(5): 882 (2018), Fig. 6

      Description: The colony on PDA has a diameter that ranges 38–40 mm after 7 d of incubation at 25 °C, light greyish brown at the surface view, round, irregular, with a dense and cottony center; as the colony grows outwards, the cottony texture disappears, and the colony becomes flat and white which later develops brown pigmentation as time passes. Hyphal system 1.88–2.96 μ ($\bar {\rm x} $ =2.31 μm), hyaline, dimitic, septate. Arthrospores 3.08–5.06 μm ($\bar {\rm x} $ = 3.94 μm), cylindrical, hyaline (Fig. 6).

      Material examined: PHILIPPINES, Iloilo, Iloilo City, Iloilo Ferry Terminal Port, on sediment samples, 31 October 2023, F.M. Dela Cruz, WHD-621, culture UPVMICC 23-011; ibid, on water samples, 31 October 2023, F.M. Dela Cruz, SHD-411, culture UPVMICC 23-016.

      Notes: Pyrrhoderma noxium was formerly classified as Phellinidium noxium but was transferred to the genus Pyrrhoderma based on morphological and phylogenetic analyses by Zhou et al.[83]. As a significant pathogen, P. noxium infects a broad host range of over 200 plant species, with a particularly devastating impact on more than 100 tropical tree species, including economically and culturally important plants like breadfruit and mango[84,85]. This pathogen poses a substantial threat to Pacific Island ecosystems due to its widespread distribution across pan-tropical regions, encompassing Asia, Australia, Africa, and Oceania[8587]. Phylogenetic analysis based on the ITS gene region placed the isolated UPVMICC 23-016 within a clade of P. noxium strains (Fig. 3). In aquatic environments, P. noxium has been isolated from marine sediment in the South China Sea[88]. The isolated Pyrrhoderma noxium UPVMICC 23-016 was the first record of its species in the brackish waters of the Philippines. Growth was observed on 25, 50, and 100 ppm cadmium.

      No previous literature has reported on the bioremediation potential or applications of P. noxium. This study is the first to identify and demonstrate its tolerance to cadmium, highlighting its potential as a novel species for bioremediation.

      Agaricales Underw., Moulds, mildews, and mushrooms. A guide to the systematic study of the Fungi and Mycetozoa and their literature (New York): 97 (1899)

      Schizophyllaceae Quél., Fl. mycol. France (Paris): 365 (1888)

      Schizophyllum Fr. [as 'Schizophyllus'], Observ. mycol. (Havniae) 1: 103 (1815)

      Schizophyllum commune Fr. [as 'Schizophyllus communis'], Observ. mycol. (Havniae) 1: 103 (1815), Fig, 5jl

      Description: Colony on PDA reaches 48–49 mm in diameter after 7 d at 25 °C, white and woolly cottony colony in surface view; obverse, white. Mycelia density is compact. Hyphae 2.94–4.23 μm ($\bar {\rm x} $ = 3.71 μm) coenocytic, hyaline, spicules, and clamp connections present, branching nondichotomously.

      Material examined: PHILIPPINES, Iloilo, Iloilo City, Iloilo Ferry Terminal Port, on sediment samples, 31 October 2023, F.M. Dela Cruz, SHD-622, culture UPVMICC 23-012.

      Notes: Schizophyllum commune is widely distributed across temperate and tropical regions[89]. It typically exists as a weak parasite or saprophyte on various woody plants, although it can occasionally be found on herbaceous species[89]. This fungus has been isolated from diverse terrestrial habitats and exhibits an unexpected versatility in aquatic environments. In marine ecosystems, S. commune has been recovered from subseafloor and deep-sea sediments[9093], as well as associated with marine macroalgae, Flabellia petiolata[94] and Padina pavonica[95], and seagrass Posidonia oceanica[96]. In the phylogenetic analysis, the isolate UPVMICC 23-012 clustered with other strains of S. commune (CBS 124811; CBS 227.57; CBS 342.58; IHEM:25479; IHEM:27140; KoRLI047393; T28)[97]. Notably, the isolation of Schizophyllum commune UPVMICC 23-012 represents the first record of this species in the brackish waters of the Philippines. The isolate was only able to grow on 25 ppm cadmium with no growth in 50 and 100 ppm.

      This fungal species has been previously proven to withstand high heavy metal concentrations and effectively bioremediate them[81,98]. The species' tolerance can be attributed to the up-regulation of glutathione S-transferase[99,100]. Another mechanism for tolerance observed in this species is their ability to bind heavy metals in their cell walls. The protein β-glucans found in the walls of S. commune can bind heavy metals such as Cu2+[101]. It has also been revealed that binding, rather than uptake, of heavy metals, is perhaps the main mechanism for tolerance. Binding to the cell wall can serve as a storage mechanism, limiting the availability of these molecules to competitors, or preventing toxic influx into the cytoplasm[102]. Additionally, S. commune is another fungal species capable of producing the enzyme laccase, with overexpression of this enzyme observed alongside the accumulation of heavy metals such as Zn, Pb, and Cd inside its organelles. Overexpression of this enzyme has also been associated with higher heavy metal tolerance[103].

      The effectiveness of S. commune as biosorbents is attributed to their rapid kinetics, impressive biosorption capacity, and ability to selectively remove metal ions from electroplating industrial effluents[104]. It has also been demonstrated that S. commune can transport heavy metals along their hyphae[105].

      Tremellomycetes Doweld, Prosyllabus Tracheophytorum, Tentamen Systematis Plantarum Vascularium (Tracheophyta) (Moscow): LXXVIII (2001)

      Trichosporonales Boekhout & Fell, Int. J. Syst. Evol. Microbiol. 50: 1363 (2000)

      Trichosporonaceae Nann. [as 'Trichosporaceae'], Repert. mic. uomo: 285 (1934)

      Trichosporon Behrend, Berliner Klin. Wochenschr. 21: 464 (1890)

      Trichosporon asahii Akagi ex Sugita, A. Nishikawa & Shinoda, J. gen. appl. Microbiol., Tokyo 40(5): 405 (1994), Fig. 7

      Description: Colony on PDA has a diameter that ranges from 29–31 mm after 7 d at 25 °C, white, creamy to dry, powdery on the surface and obverse view, the center is farinose and agar fissures on its reverse side. Blastoconidia 2.58–5.04 μm ($\bar {\rm x} $ = 3.60 μm), cylindrical, subglobose to ovoidal, hyaline (Fig. 7).

      Material examined: PHILIPPINES, Iloilo, Iloilo City, Iloilo Ferry Terminal Port, on sediment samples, 31 October 2023, F.M. Dela Cruz, SHD-022, culture UPVMICC 23-013.

      Notes: Trichosporon, a fungus resembling yeast, is widely distributed in nature and found in soil, water, plants, animals, and even within the natural flora of the human body[106]. Although typically harmless, it can cause illness in humans under certain circumstances. Research spanning several decades has documented instances of Trichosporon asahii infection, totaling 140 cases between 1996 and 2019, with reports from various regions worldwide. Asia accounted for most of these cases, comprising 77.1% of the total[107]. The isolate's ITS gene region gave 93% ML support as its clade with several T. asahii strains. Together with the other type strains, they grouped with strains of Trichosporon aquatile, T. japonicum and T. asteroides with 100 % ML and 1.00 BYPP support (Fig. 3). In aquatic environments, T. asahi was isolated from sand[108], and sediments[109113]. The isolation of T. asahii UPVMICC 23-013 represents the first record of this species in the brackish waters of the Philippines. The isolate has potential for use in cadmium bioremediation, as it grew in all tested concentrations of cadmium (25, 50, and 100 ppm).

      This species has been observed to tolerate not only cadmium but also other heavy metals, including Pb, Cu, and As. The resistant strains exhibited multiple heavy metal binding sites in their cell walls[114,115]. However, it has also been observed that the involvement of the cell wall in the heavy metal stress response caused phenotypic changes, particularly in the cell wall structure[116]. Additionally, the glutathione (GSH)-glutathione disulfide (GSSG) system, the most prevalent redox system in fungi and yeasts, was downregulated in T. asahi when exposed to heavy metals, except for cadmium, where it was increased[111,114]. GSH is thought to be a tolerance mechanism of the yeast against high levels of oxidative stress caused by metal exposure. Yeasts are also observed to have efficient heavy metal accumulation methods[114]. This suggests that cadmium has a specific and positive impact in the tolerance and bioremediation capability of T. asahii.

      Figure 6. 

      Pyrrohoderma noxium UPVMICC 23-016. Colony on PDA with 25 ppm Cd(II) from (a) surface, and (b) reverse. (c) Septate hyphae beginning to fragment. (d)–(f) Arthrospores. Scale bars: (c)–(f) 25 μm.

      Figure 7. 

      Trichosporon asahii UPVMICC 23-013. Colony on PDA with 25 ppm Cd(II) from (a) surface, and (b) reverse. (c)–(f) Blastoconidia. Scale bars: (c)–(f) = 25 μm.

    • The Iloilo River, particularly the Iloilo-Guimaras Ferry Port, serves as a significant sample site, revealing the presence of heavy metal-tolerant fungi. This observation suggests potential heavy metal contamination in the area, as such fungi are typically found in polluted environments[14,117122]. The level of heavy metal tolerance displayed by these fungi reflects the conditions of their isolation site. Fungi isolated from metal-contaminated areas tend to exhibit higher resistance compared to those from uncontaminated sites, indicating that heavy metals exert selective pressure favoring resistant microorganisms[123]. Studies on the Iloilo River have consistently shown that heavy metal concentrations exceed the guidelines set by the Philippine Department of Environment and Natural Resources (DENR)[124127]. The contamination is likely driven by activities at the ferry port, including leachates from reclamation areas, port operations, untreated industrial effluents, and domestic sewage[128130].

      Sediment samples yielded more fungal isolates than water samples, highlighting the critical role of sediments as a source for identifying diverse fungal species. This can be attributed to sediment's vital function in river ecosystems, where it acts both as a sink and source of heavy metals[131,132]. As a dynamic component of the river basin, sediment shapes diverse habitats and environments[133]. Heavy metals tend to accumulate in sediments at significantly higher concentrations than in the water column, particularly in river mouths, which are often more polluted than other zones[134137]. The proximity of the ferry port further supports the sampling choice, as port infrastructure and operations are linked to elevated heavy metal concentrations[138]. Sediments not only collect heavy metals from river systems but can also emit them back into the environment. Elevated metal levels in sediments often reflect the geochemical composition of the parent material[139]. Heavy metals in sediments and water commonly originate from effluents released by metal processing, recreational, municipal, and agricultural industries. These pollutants undergo chemical and microbial transformations, impacting environmental quality and public health[140,141]. Furthermore, sediments provide a more stable and nutrient-rich environment for fungal growth compared to water. The stability of the substrate supports the development of vegetative hyphae in filamentous fungi while offering refuge from environmental fluctuations, contributing to the higher fungal diversity observed in sediments. In contrast, aquatic environments are more dynamic and less conducive to fungal stability, leading to lower isolation rates[142146].

      Phylogenetic analysis of the fungal isolates identified two phyla: Ascomycota and Basidiomycota. Basidiomycota accounted for 66.7% (six isolates), and Ascomycota for 33.3% (three isolates). All isolates demonstrated growth on PDA plates supplemented with 25 ppm of cadmium during the initial screening. Within the Basidiomycota phylum, all isolates belonged to the subphylum Agaricomycotina, while all Ascomycota isolates were classified under Pezizomycotina. These isolates were further categorized into three classes: Agaricomycetes (five isolates; 55.6%), Sordariomycetes (three isolates; 33.3%), and Tremellomycetes (one isolate; 11.1%). The subclasses represented were Xylariomycetidae (two isolates; 22.2%), Incertae sedis (four isolates; 44.4%), Agaricomycetidae (two isolates; 22.2%), and Diaporthomycetidae (one isolate; 11.1%). The orders identified included Polyporales (three isolates; 33.3%), and Trichosporonales, Agaricales, Hymenochaetales, Xylariales, Amphisphaeriales, and Diaporthales, each with one isolate (11.1%). At the family level, Diaporthaceae, Trichosporonaceae, Schizophyllaceae, Hymenochaetaceae, Meruliaceae, Irpicaceae, Cerrenaceae, Diatrypaceae, and Pestalotiopsidaceae were each represented by one isolate (11.1%).

      These findings align with a metagenomic study[147,148], investigating fungal communities in heavy metal-contaminated sites, which also identified Ascomycota and Basidiomycota as dominant phyla. Similar to this study, Sordariomycetes was the most prominent class, supporting the assertion that it includes heavy metal-tolerant fungi. Within Basidiomycota, Agaricomycetes was the most abundant class, followed by Exobasidiomycetes and Tremellomycetes, consistent with the results of this study, where five of six Basidiomycota isolates belonged to Agaricomycetes, and one was classified as Tremellomycetes. Metagenomic analysis further revealed that isolates from these phyla possessed a higher number of genes associated with metabolic pathways, gene information processing, biosorption processes, and DNA repair, which were absent in isolates from non-contaminated sites. These genetic adaptations likely underpin the heavy metal tolerance of these fungal communities. Greater fungal diversity was also observed in less-contaminated sites, as reported in related studies[148].

      Ascomycota and Basidiomycota exhibit remarkable potential for heavy metal bioremediation and are often the dominant fungal groups in contaminated sites[80,148150]. The heavy metal tolerance of Basidiomycota members is attributed to their ability to produce a variety of degradative enzymes, such as peroxidases, H2O2-generating oxidases, and laccases, commonly found in white rot fungi species[81,82,151155]. These enzymes enhance survival and metabolic activities while supporting other fungal species in the consortium[80,145]. Similarly, members of the Ascomycota phylum demonstrate heavy metal tolerance through biosorption and metabolic capabilities, alongside their robust survival across diverse habitats, allowing them to adapt rapidly to environmental changes[8082,147157].

    • This study elucidates the considerable fungal diversity within the Iloilo River, revealing nine taxa with tolerance to cadmium exposure. The identification of taxa from both Ascomycota and Basidiomycota emphasizes the ecological resilience of fungi in response to anthropogenic pollution. These findings not only augment our understanding of fungal biodiversity within this aquatic environment but also provide a foundational basis for subsequent investigations into the bioremediation potential of these isolates. By harnessing the inherent metabolic capabilities of these fungi, we can develop innovative and sustainable strategies for mitigating heavy metal contamination. This approach holds promise for the restoration and protection of aquatic ecosystems, ultimately enhancing public health outcomes by reducing environmental toxicity.

      • The authors would like to thank the Department of Agriculture (DA) - Bureau of Fisheries and Aquatic Resources (BFAR) for the issuance of the gratuitous permit. MS Calabon is grateful to the UP System Balik PhD Program (OVPAA-BPhD-2022-02) entitled, 'Unraveling the hidden diversity of aquatic fungi from Panay Island, Philippines'.

      • The authors confirm contribution to the paper as follows: study conception and design: Dela Cruz FM, Calabon MS, Bermeo-Capunong MRA; data collection: Dela Cruz FM, Bagacay JFE; analysis and interpretation of results: Dela Cruz FM, Bagacay JFE, Calabon MS, Canto CM; draft manuscript preparation: Dela Cruz FM, Calabon MS. All authors reviewed the results and approved the final version of the manuscript.

      • All data generated during the study is available within the article.

      • The authors declare that they have no conflict of interest.

      • Copyright: © 2025 by the author(s). Published by Maximum Academic Press, Fayetteville, GA. This article is an open access article distributed under Creative Commons Attribution License (CC BY 4.0), visit https://creativecommons.org/licenses/by/4.0/.
    Figure (7)  Table (1) References (157)
  • About this article
    Cite this article
    Dela Cruz FM, Bermeo-Capunong MRA, Bagacay JFE, Canto CM, Calabon MS. 2025. Taxonomy, phylogeny, and preliminary screening of fungal isolates for cadmium tolerance from Iloilo Ferry Terminal, Iloilo, Philippines. Studies in Fungi 10: e001 doi: 10.48130/sif-0025-0001
    Dela Cruz FM, Bermeo-Capunong MRA, Bagacay JFE, Canto CM, Calabon MS. 2025. Taxonomy, phylogeny, and preliminary screening of fungal isolates for cadmium tolerance from Iloilo Ferry Terminal, Iloilo, Philippines. Studies in Fungi 10: e001 doi: 10.48130/sif-0025-0001

Catalog

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return