ArticlePDF AvailableLiterature Review

Abstract and Figures

Biofilms are highly-organized microbial communities attached to a biotic or an abiotic surface, surrounded by an extracellular matrix secreted by the biofilm cells. The majority of fungal pathogens contribute to biofilm formation within tissues or biomedical devices, leading to serious and persistent infections. The clinical significance of biofilms relies on the increased resistance to conventional antifungal therapies and suppression of the host immune system, which leads to invasive and recurrent fungal infections. While different features of yeast biofilms are well-described in the literature, the structural and molecular basis of biofilm formation of clinically related filamentous fungi has not been fully addressed. This review aimed to address biofilm formation in clinically relevant filamentous fungi.
Content may be subject to copyright.
Full Terms & Conditions of access and use can be found at
https://www.tandfonline.com/action/journalInformation?journalCode=imby20
Critical Reviews in Microbiology
ISSN: (Print) (Online) Journal homepage: https://www.tandfonline.com/loi/imby20
Biofilm formation in clinically relevant filamentous
fungi: a therapeutic challenge
Maryam Roudbary, Roya Vahedi-Shahandashti, André Luis Souza dos
Santos, Shahla Roudbar Mohammadi, Peyman Aslani, Cornelia Lass-Flörl &
Célia F. Rodrigues
To cite this article: Maryam Roudbary, Roya Vahedi-Shahandashti, André Luis Souza dos Santos,
Shahla Roudbar Mohammadi, Peyman Aslani, Cornelia Lass-Flörl & Célia F. Rodrigues (2021):
Biofilm formation in clinically relevant filamentous fungi: a therapeutic challenge, Critical Reviews in
Microbiology, DOI: 10.1080/1040841X.2021.1950121
To link to this article: https://doi.org/10.1080/1040841X.2021.1950121
Published online: 06 Aug 2021.
Submit your article to this journal
View related articles
View Crossmark data
REVIEW ARTICLE
Biofilm formation in clinically relevant filamentous fungi: a
therapeutic challenge
Maryam Roudbary
a
, Roya Vahedi-Shahandashti
b
, Andr
e Luis Souza dos Santos
c
,
Shahla Roudbar Mohammadi
d
, Peyman Aslani
e
, Cornelia Lass-Fl
orl
b
and C
elia F. Rodrigues
f
a
Department of Parasitology and Mycology, School of Medicine, Iran University of Medical Sciences, Tehran, Iran;
b
Institute of
Hygiene and Medical Microbiology, Medical University Innsbruck, Innsbruck, Austria;
c
Department of General Microbiology,
Microbiology Institute Paulo de G
oes, Federal University of Rio de Janeiro (UFRJ), Brazil;
d
Department of Mycology, Faculty of Medical
Science, Tarbiat Modares University, Tehran, Iran;
e
Department of Parasitology and Mycology, Faculty of Medicine, Aja University of
Medical Sciences, Tehran, Iran;
f
LEPABELaboratory for Process Engineering, Environment, Biotechnology and Energy, Faculty of
Engineering, University of Porto, Porto, Portugal
ABSTRACT
Biofilms are highly-organized microbial communities attached to a biotic or an abiotic surface,
surrounded by an extracellular matrix secreted by the biofilm-forming cells. The majority of fun-
gal pathogens contribute to biofilm formation within tissues or biomedical devices, leading to
serious and persistent infections. The clinical significance of biofilms relies on the increased
resistance to conventional antifungal therapies and suppression of the host immune system,
which leads to invasive and recurrent fungal infections. While different features of yeast biofilms
are well-described in the literature, the structural and molecular basis of biofilm formation of
clinically related filamentous fungi has not been fully addressed. This review aimed to address
biofilm formation in clinically relevant filamentous fungi.
ARTICLE HISTORY
Received 2 May 2021
Revised 17 June 2021
Accepted 19 June 2021
Published online 5 August
2021
KEYWORDS
Biofilm; filamentous fungi;
antifungal resistance
mechanisms; quo-
rum sensing
1. Introduction
Invasive fungal infections are one of the major causes
of morbidity and mortality, being attributed to enor-
mous healthcare-related costs, worldwide (Brown et al.
2012; Arastehfar et al. 2020). Species belonging to
yeasts, like Candida and Cryptococcus, and filamentous
fungi, such as Aspergillus, are considered as the most
clinically relevant fungal species, being able to cause an
array of infection in humans, ranging from superficial to
deep-seated systemic infections (Brown et al. 2012).
Unlike bacterial infections, for which there are numer-
ous classes of antibiotics with a wide range of mecha-
nisms of action, the number of antifungal agents
available in clinical settings is limited to four major
classes, namely azoles, polyenes, echinocandins, and
flucytosine (Gintjee et al. 2020). Echinocandins and
mould-active triazoles are mainly used to treat invasive
yeast and mould infections, respectively, while the use
of amphotericin B, a polyene agent, is limited due to its
nephrotoxicity (Arastehfar et al. 2020).
Similarly to bacteria, fungal species inherently pro-
duce complex structures, biofilms. These are a
community of cells embedded with extracellular matrix
(ECM) (Ramage et al. 2009; Pitangui et al. 2012). This
ECM is mostly composed of extracellular polysacchar-
ides, proteins, DNA, and lipids (e.g. ergosterol), which
shields the embedded cells from hostile host conditions
and renders them to be extremely resistant to various
chemical compounds, such as antifungals, disinfectants
used in the hospital settings, and other agents found in
natural niches (Ramage et al. 2009; Bridier et al. 2011;
Pitangui et al. 2012; Rodrigues et al. 2017). Biofilms are
also seen as a key virulence factor, which, after colon-
isation, allow the pathogen to stabilise and conse-
quently resulting in persistent infections (Desai et al.
2014). Finally, these features let the pathogenic fungi to
survive in hospital environment and seed future out-
breaks (Schelenz et al. 2016; Sherry et al. 2017).
Although an accurate estimation is lacking in clinic-
ally important fungi, it has been calculated that
approximately 80% of the recurrent and chronic bacter-
ial infections (Davies 2003) and 500,000 annual deaths
(Sharifi et al. 2018) are attributed to biofilms. Since bio-
films can be formed on both biotic and abiotic surfaces,
CONTACT C
elia F. Rodrigues c.fortunae@gmail.com LEPABELaboratory for Process Engineering, Environment, Biotechnology and Energy, Faculty
of Engineering, University of Porto, Porto, Portugal
ß2021 Informa UK Limited, trading as Taylor & Francis Group
CRITICAL REVIEWS IN MICROBIOLOGY
https://doi.org/10.1080/1040841X.2021.1950121
and the fact that the number of implanted devices
(Raad 1998; Costerton et al. 1999; Donlan 2002;
Vlamakis 2011), such as central venous catheters,
indwelling bladder catheters, joint prostheses, are
increasingly being used in clinical settings, it would be
plausible to assume a growing trend towards biofilm-
related infections. The clinical relevance of biofilms is
further compounded by the fact that recalcitrant bio-
film-related infections mostly require removal of the
implant devices and affected tissues (Desai et al. 2014),
which not only increases the economic burden, but
also complicate the patients outcomes.
Apart from the knowledge gap on the clinical impact
of fungi, studying biofilms structures, and the dynamics
and complex circuits regulating biofilm formation have
been mainly focussed on a major yeast, Candida albi-
cans. On the other side, there is a very limited informa-
tion on other fungal species, such as Aspergillus,
Zygomycetes,Scedosporium,Histoplasma,Fusarium, and
black-yeast like fungi. These fungi can successfully col-
onise surfaces and produce strong biofilms. Therefore,
the scope of this review is to discuss the biofilm forma-
tion of clinically relevant filamentous fungi that have
been stagnant, in comparison with bacteria and other
fungi. Additionally, we highlight the structural and
molecular basis of biofilms and their role in the devel-
opment of resistance to antifungal agents.
2. Biofilms in the medically prevalent
filamentous fungi
2.1. Aspergillus spp.
Aspergillus fumigatus, the causative agent of invasive
respiratory infections in immunocompromised patients
and colonisation in the airways of cystic fibrosis (CF)
patients (Oxford Academic 2020a), has demonstrated to
be involved in biofilm formation (Chandrasekar and
Manavathu 2008; Loussert et al. 2010; Morisse et al.
2013; Beauvais and Latg
e2015). The primary establish-
ment of chronic aspergillosis involves the germination
of conidia into mycelia, following the invasion into pul-
monary cells (Filler and Sheppard 2006). Scanning elec-
tron microscopy (SEM) has indicated that the aerial
hyphae of the mycelial colony developed over pulmon-
ary cells are embedded in a dense hydrophobic ECM
(Beauvais et al. 2007). The resistance of colonies
embedded in ECM to amphotericin B, compared with
the planktonic state, supports the evidence of biofilm
formation during pulmonary aspergillosis (Beauvais
et al. 2007; Ader et al. 2009; Escande et al. 2011).
Subsequently, in vivo, the presence of mature A. fumi-
gatus biofilms has been confirmed in both aspergilloma
and during the development of disseminated aspergil-
losis (Loussert et al. 2010). Bronchopulmonary lavage
(BAL) histopathologic examination of patients with
aspergillosis has revealed a complex multicellular myce-
toma structure (Microbiology Society 2020). The high
percentage of failure of antifungal therapy in the treat-
ment of invasive aspergillosis may be associated with
Aspergillus spp. pulmonary biofilms. In several studies,
the biofilm formation of A. fumigatus occurs under in
vitro conditions on polystyrene microtiter plates
(Mowat et al. 2007; Seidler et al. 2008; Singhal et al.
2011) and biomaterial devices such as catheters, cardiac
pacemakers, joint replacements, heart valves, and
breast augmentation implants have been extensively
reported (Escande et al. 2011; Jeloka et al. 2011; Sardi
et al. 2014). Despite the lack of clinical in vivo studies,
the ability of biofilm formation in vitro by other
Aspergillus spp., of Aspergillus flavus,Aspergillus terreus,
and Aspergillus niger has been reported (Fiori et al.
2011; Chatterjee and Das 2020).
Remarkably, a high percentage of biofilms consist of
mixed-species communities of bacteria and fungi that
are profoundly common and clinically important (Amin
et al. 2010; Peleg et al. 2010). Aspergillus fumigatus and
non-fumigatus Aspergillus are frequently co-colonising
with Pseudomonas aeruginosa, a common Gram-nega-
tive bacterium, in the airways of CF patients, open skin
wounds in burn patients, and cardiac implants (Amin et
al. 2010; Peleg et al. 2010; Manavathu et al. 2014). The
structural and functional in vitro properties of mixed
biofilms of P. aeruginosa and A. fumigatus vary from
those of monomicrobial biofilms, occurring in a severe
clinical problem in CF patients and other patient groups
at risk of airway infection with these two species
(Manavathu et al. 2014). For example, phenazine-
derived metabolites, redox-active small molecule sec-
ondary metabolites, produced by P. aeruginosa, can
serve as interspecies signals in co-culture biofilm with
A. fumigatus (Zheng et al. 2015). Biofilm formation of A.
fumigatus and Aspergillus nidulans has been differen-
tially modulated by P. aeruginosa phenazine-derived
metabolites, shifting from weak vegetative growth to
induced asexual sporulation (Zheng et al. 2015). A wide
range of diseases are related to the co-infection of fungi
and bacteria at various tissue have been described in
numerous studies. As an example, we note P. aerugi-
nosa,S. aureus,Burkholderia cepacia,Acinetobacter bau-
manii,Haemophilus influenzae mixed with Candida
albicans,A. fumigatus, and Scedosporium spp. (Dixon
and Hall 2015), and in burn wounds and trauma sites,
Candida,Aspergillus,Mucorales, and Fusarium spp. co-
colonized with Pseudomonas and Staphylococcus spp.
2 M. ROUDBARY ET AL.
(Peleg et al. 2010). The increasing co-infections of fungi
and bacteria explain the necessity of understanding the
mechanisms and to outcomes of polymicrobial bio-
film formation.
2.2. Zygomycetes
Zygomycosis, a fulminant opportunistic fungal infec-
tion, is caused by fungi belonging to the class
Zygomycetes and the order Mucorales, including
Rhizopus,Rhizomucor,Mucor,Lichtheimia,
Apophysomyces,Cunninghamella, and Saksenaea genus
(Ribes et al. 2000). Opportunistic infections caused by
Zygomycetes, can be characterised by angioinvasion
and fungal-ball formation (Chakrabarti et al. 2006).
Mucormycosis is categorised into rhino-orbito-cerebral
(ROC), pulmonary, gastrointestinal, cutaneous, and dis-
seminated types, depending upon the clinical presenta-
tion and the anatomical sites involved (Chakrabarti et
al. 2006). The biofilm-formation potential of common
pathogenic zygomycetes including Rhizopus oryzae
(Almeida et al. 2018), Lichtheimia corymbifera,
Rhizomucor pusillus, and Apophysomyces elegans has
been previously addressed (Singh et al. 2011). Phase-
contrast and confocal laser scanning micrographs dem-
onstrated the existence of highly compact adherent
hyphal structures encased in an ECM composed primar-
ily of N-acetylglucosamine and glucosamine in all the
examined species, excluding A. elegans (Singh et al.
2011). It has also been shown that the biofilm forma-
tion may be involved in angio-invasive infections
caused by Mucorales, including paranasal fungal balls,
chronic rhinosinusitis, and catheter-based infections
(Singh et al. 2011; Ezzat et al. 2020).
2.3. Black yeast-like fungi
Black fungi members include melanized species with
either slimy conidia or dry conidia (Chowdhary et al.
2014). Melanized species with wet conidia are nearly
common etiologic agents of human infections, in con-
trast, to dry conidia (Wang et al. 2018). Black fungi spe-
cies belonging to the genera Exophiala,Aureobasidium,
Ochroconis,Coelomycete and Pyrenochaeta are associ-
ated with human infections (Lugauskas et al. 2004; Lian
and de Hoog 2010). Humid conditions such as bath-
rooms, swimming pools, and sauna serve as favoured
environments for black fungi including from the
Exophiala genus (e.g. Exophiala equina,Exophiala leca-
nii-corni), which are the most common causative agents
of superficial mycoses in humans (Hilmarsdottir et al.
2005; Heinrichs et al., 2013a). Exophiala castellanii,
Exophiala equina,Exophiala lecanii-corni,Knufia epider-
midis, and Ochroconis spp. have been detached from
black pigmented biofilms happening in indoor wet cells
(Heinrichs et al., 2013a,2013b). A comparative metage-
nomic analysis of the dark pigmented fungal biofilms of
the water-air interphase of water taps, revealed that E.
lecanii-corni is a dominant component of these biofilms
(Heinrichs et al., 2013b). Of note, the biofilm of
Exophiala dermatitidis (caused severe phaeohyphomy-
cosis), has also exhibited higher resistance to the tested
antimycotic agents than the planktonic form (Gao et al.
2016;2017; Kirchhoff et al. 2017). Further investigation
showed that the biofilm formation and virulence of E.
dermatitidis were influenced by co-cultivation with P.
aeruginosa. For example, N-acyl-L-homoserine lactone
(AHL), P. aeruginousa quorum sensing molecule, inhibits
biofilm and filaments formation of E. dermatitidis
(Kirchhoff et al. 2020). Therefore, further research is
required to evaluate the potential pathogenesis of the
darkly pigmented biofilms.
2.4. Histoplasma capsulatum
Histoplasma capsulatum, a thermally dimorphic fungus
with yeast and filamentous phases, causes a respiratory
and systemic mycosis termed histoplasmosis (Sil and
Andrianopoulos 2014). As H. capsulatum has a high dis-
persion ability to adapt to various conditions by
expressing specific genes, the filamentous form of that
might present in medical settings (L
opez 2006; Carreto-
Binaghi et al. 2015). The adherence pattern to pneumo-
cytes as well as biofilm formation ability of H. capsula-
tum yeasts has been assessed (Pitangui et al. 2012). The
results revealed efficient adherence to host cells and
biofilm formation, which could be a possible adapted
survival strategy to avoid the immune cells (Pitangui et
al. 2012). Furthermore, a reduced activity of amphoteri-
cin B and itraconazole against biofilm of H. capsulatum
has also been shown (Brilhante et al. 2015). Farnesol
showed promising outcomes, as an adjuvant, by
enhancing the susceptibility to antifungal drugs in both
planktonic and biofilm structure (Brilhante et al. 2015).
Additionally, the development of H. capsulatum biofilms
in different nutritional conditions and oxygen concen-
trations, clarified that lower oxygen concentrations and
medium components are associated with a denser bio-
film and higher polysaccharide content (Gonc¸alves et
al. 2020). Currently, the filamentation of H. capsulatum
in the biofilm is unknown. Nevertheless, analysing the
reports, it can be deduced that the limitation of crucial
nutrients might be a stressor that stimulates the fila-
mentation transition from yeast (Gonc¸alves et al. 2020).
CRITICAL REVIEWS IN MICROBIOLOGY 3
2.5. Scedosporium spp.
Scedosporium spp.,cause a wide spectrum of life-threat-
ening infections with broad clinical manifestation, rang-
ing from colonisation of the respiratory tract,
sinopulmonary, extrapulmonary, and disseminated
infections (Cortez et al. 2008). Biofilm formation and
adherence ability of Scedosporium apiospermum,
Scedosporium aurantiacum, and Scedosporium minutis-
porum have been described in both polystyrene and
lung epithelial cells (Mello et al. 2016; Kitisin et al.
2020). In the biofilms, the dense compact of the myce-
lial biomass with internal channels explains the biofilm
formation ability by these fungal pathogens (Mello et
al. 2016). Further examination on the matured
Scedosporium spp. biofilms proved that presence of
ECM consisted of carbohydrate-rich molecules and
eDNA (Mello et al. 2020). Scedosporium minutisporum
and S. aurantiacum manifested a slight higher amount
of ECM composition than S. apiospermum (Mello et al.
2020). This genus has shown adherence to epithelial
lung cells with different avidity (de Mello et al. 2020).
The forming monolayer of mycelia over the cells have
demonstrated the ability to differentiate into hyphae,
resulting in irreversible damage to the epithelial cell
and cell death (de Mello et al. 2020).
2.6. Fusarium spp.
Fusarium spp. have a high genetic diversity and resist-
ance to systemic antifungals, causing a broad spectrum
of infections dependent on the immune status of the
host (Sav et al. 2018). The biofilm-forming capacity of
clinical isolated Fusarium spp. showed that Fusarium
solani,Fusarium petroliphilum, and Fusarium keratoplas-
ticum can produce biofilms (Sav et al. 2018). As it hap-
pens with other genus, the increased antifungal
minimum inhibitory concentrations (MICs) of biofilm-
forming Fusarium isolates, compared with their plank-
tonic counterparts, indicate the possible correlation of
biofilm formation and increased resistance (Sav et al.
2018). Furthermore, the compacted architecture of
Fusarium oxysporum f. sp. cucumerinum with extracellu-
lar polysaccharide materials revealed less susceptibility
to fungicides and environmental stress, including tem-
perature, UV light (Peiqian et al. 2014). Fusarium kerato-
plasticum, a common species of human infections can
develop a dense biofilm, with an increased matrix with
the presence of extracellular DNA, RNA, polysacchar-
ides, and proteins (Kischkel et al. 2020). The anti-biofilm
activity of farnesol on the preformed biofilm resulted in
the destruction of hyphae and the ECM (Kischkel et al.
2020), preventing the adhesion of conidia,
filamentation, and the formation of biofilm (Kischkel et
al. 2020). Elucidating the different stages of biofilm for-
mation in F. solani showed a profound structure with
increased resistance to natamycin, voriconazole, caspo-
fungin, and amphotericin B (C
ordova-Alc
antara et al.
2019). On contact lenses and fingernails (Burkhart et al.
2002), the biofilm formation of Fusarium (Mukherjee et
al. 2012) showed to be a strong contributor to patho-
genicity, following the high virulence and mortality rate
in invasive infections.
2.7. Dermatophytes
Dermatophytes are fungi responsible for dermatophyt-
osis, which include superficial fungal infections of the
skin and its appendages (Danielli et al. 2017).
Dermatophytosis is the most common mycosis world-
wide, affecting 2025% of the general population.
Azole and allylamines are the most common drugs
used as treatment; however dermatophytes are
described to be resistant to most of them, particularly
biofilms, in addition to their considerable toxicity
(Costa-Orlandi et al. 2020). Trichophyton rubrum, C. albi-
cans, and Candida parapsilosis are some of the most
prevalent species related to this disease. Numerous
reports indicate a variable predominance of Candida
spp. in relation to dermatophytes, especially in onycho-
mycosis and the possibility of isolating both from the
same site (Garcia et al. 2020).
Costa-Orlandi et al. (2014) described, for the first
time, the characteristics of T. rubrum and Trichophyton
mentagrophytes biofilms. The authors indicated that
both species can form a mature biofilm within 72 h, and
T. rubrum biofilm produces more biomass and exopoly-
saccharide (EPS) and is denser than T. mentagrophytes
biofilm. Additionally, the results demonstrated a coordi-
nated network of hyphae in all directions, embedded
within EPS in some areas (Costa-Orlandi et al. 2014).
There is presently no evidence on the involvement
of filamentous fungi in multi-species biofilms. A recent
report revealed that, notwithstanding the predomin-
ance of Candida spp., the presence of T. rubrum seems
to inhibit C. albicans filamentation and C. parapsilosis
development, which confirmed an antagonistic inter-
action, and susceptibility to terbinafine and efinacona-
zole (Garcia et al. 2020).
Unprecedentedly, it was described the biofilm forma-
tion by Microsporum canis, responsible for tinea capitis
and categorised by a reduced therapeutic response
(Danielli et al. 2017). The study indicated a highly struc-
tured network of hyphae growing in all directions, with
compacted EPS regions, but typically with a porous and
4 M. ROUDBARY ET AL.
thin consistency, concentrated in the superficial region
of the mycelium (Danielli et al. 2017). An ex vivo model
for dermatophyte biofilm growth, using hair from dogs
and cats, was developed by Brilhante et al. (2019). With
some variations among species, that work showed
the appropriateness of this model to develop
structured biofilms with good biomass of M. canis
and Microsporum gypseum,T. mentagrophytes and
Trichophyton tonsurans. Furthermore, the microscopic
analysis of the biofilms showed hyphae colonising and
perforating the hair shaft, profuse fungal conidia, bio-
film EPS and biofilm water channels (Brilhante et
al. 2019).
Several recent studies have showed the potential anti-
fungal effect of novel drugs, against dermatophytes. A
recent work incorporated nonyl 3,4-dihydroxybenzoate
(fungicide compound against planktonic cells) and derm-
atophyte biofilms in nanostructured lipid systems, to
decrease toxicity in high concentrations, increase its
solubility and maintain its effectiveness against T. rubrum
and T. mentagrophytes biofilms. The formulation showed
a good efficacy against biofilms and a lower toxicity
(Costa-Orlandi et al. 2020). Similarly, the anti-dermato-
phytic and anti-biofilm activity of 2-hydroxychalcone
(2-chalcone) in the dark and photodynamic therapy
(PDT)-mediated against T. rubrum and T. mentagrophytes
biofilms. It was showed that the drug targets ergosterol
in the cell and promotes the generation of ROS, leading
to cell death by apoptosis and necrosis. Early-stage bio-
film and mature biofilms were inhibited by 2-chalcone,
indicating the potential therapeutical use of this com-
pound (Bila et al. 2021).
Formation of biofilm by dermatophytes is a key fac-
tor of fungal virulence, which whitethorn be tangled to
the persistence of infections. The characterisation of
biofilms formed by dermatophytes can surely contrib-
ute to the search of new drugs, doses and duration of
the treatment of these mycoses.
2.8. Paracoccidioides,Sporothrix and
Coccidioides
Paracoccidioides species are dimorphic fungi that firstly
infect the lungs, but can also spread through the body
(e.g. oral cavity, systemic colonisation, vascular pros-
thesis), most likely due to the formation of a biofilm
that makes it hard for the host to eradicate the infec-
tion (Sardi et al. 2015). These species are associated to
paracoccidioidomycosis, an endemic mycosis limited to
Latin America arising with the chronic form in 90% of
cases (Cattana et al. 2017).
The characterisation of Paracoccidioides brasiliensis
biofilms can help understanding the pathogenesis of
paracoccidioidomycosis, and the search for new and
effective drugs. Paracoccidioides brasiliensis biofilm and
gene expression of adhesins and hydrolytic enzymes
was reported by Sardi et al. (2015). The authors indi-
cated that the presence of a biofilm is associated with a
growth in the expression of adhesins and enzymes (e.g.
GP43, enolase, GAPDH and aspartyl proteinase). On
contrary, phospholipase was down-regulated in biofilm
(Sardi et al. 2015). Moreover, the dynamics of mono-
and dual-species biofilm growth of Paracoccidioides bra-
siliensis and C. albicans and their in vivo infectivity was
evaluated by Oliveira et al. (2020). The results indicated
that dual-species biofilm with P. brasiliensis plus C. albi-
cans presented a higher biomass, higher metabolic
activity and CFUs (colony forming unit) than their
mono-species biofilms. Galleria mellonella larvae
showed a reduction in the survival rate when infected
with P. brasiliensis and C. albicans, when compared with
those infected with single species, which finally indi-
cated that P. brasiliensis and C. albicans co-existence is
likely to occur on oral mucosal biofilms (Oliveira et
al. 2020).
Observing other species, it is known that Sporothrix
schenckii complex can origin a benign subcutaneous
mycosis, and biofilm formation may be one of the sig-
nificant elements associated to its virulence. S
anchez-
Herrera et al. (2021) reported that throughout the
development, S. schenckii biofilms are surrounded by
EPS, particularly glycoprotein (mannose rich), carbohy-
drates, lipids, and nucleic acids. The later was consid-
ered as a key component to structural integrity and
antifungal resistance (S
anchez-Herrera et al. 2021).
Finally, recurrent coccidioidal meningitis has also been
related to biofilms (on the tip of ventriculo-peritoneal
shunt tubing). The formed biofilm was the probable
responsible for a 4-year persistence of Coccidioides
immitis, notwithstanding the fact that the patient took
an adequate dosage of fluconazole (Davis et al. 2002).
Studies fungal species biofilm contributes to a better
understanding of growth biofilm and physiology, add-
ing new insights into the mechanisms of virulence and
persistence of pathogenic microorganisms.
An overview of clinically relevant filamentous fungi
that can form biofilm is given in Table 3.
3. Biofilm structures in filamentous fungi
Biofilms produced by budding yeasts are more compar-
able to bacteria rather than those formed by filament-
ous fungi, with hyphal tip growth (Beauvais et al. 2007;
CRITICAL REVIEWS IN MICROBIOLOGY 5
Harding et al. 2009; Mowat et al. 2009). Filamentous
fungal biofilms are thought to have unique morpho-
logical traits, since they lack the binary fission and the
growth dynamics observed in the asexual phase in
yeasts (Biesebeke et al. 2002; Beauvais et al. 2007; Gulis
et al. 2008; Harding et al. 2009). These distinct morpho-
logical properties may have contributed to a specialised
liquid-air interface structures in filamentous fungi,
which thought to be involved in host penetration and
nutrient acquisition (Harding et al. 2009).
Some studies have explored the biofilm properties of
the most common filamentous fungi, namely A. fumiga-
tus (Beauvais et al. 2007; Mowat et al. 2008,2009), A.
niger (Villena and Gutierrez-Correa 2006), F. solani, and
F. oxysporum (Imamura et al. 2008; Smith and Mcginnis
2005), through microscopic and antifungal susceptibil-
ity testing approaches.
Based on all the descriptions, a step-wise model for
filamentous fungal biofilms has been proposed
(Harding et al. 2009; Costa-Orlandi et al. 2017). The sug-
gested model provides a better understanding of the
nature of filamentous fungal biofilms development
(Figure 1): 1. adsorption: contacting of propagules such
as spores, sporangia, or hyphal fragments of the organ-
ism to a biotic or abiotic surface; 2. fixed adhesion:
attachment of spores or other reproductive structures
to surface via secreting adhesins during germination or
hydrophobins production, involved proteins in the
adhesion of hyphae to hydrophobic surfaces (W
osten
2001); 3. initial microcolony colonisation: formation
microcolonies through apical elongation and hyphal
branching. This step following by monolayer formation
and ECM production, resulting in tenacious adherence
to the substrate; 4. preliminary maturation: formation
condensed hyphal networks or hypha-hypha adhesion,
resulting in layering covered by ECM, and water chan-
nel formation via hydrophobic repulsion among
hyphae; 5. maturation: formation survival structures
(fruiting bodies, sporogenous cells, sclerotia, and other)
depending on the fungi; 6. dispersion or planktonic
phase: releasing conidia or hyphae fragments, resulting
in a new cycle.
4. Molecular basis governing filamentous
fungal biofilm development
The application of omicstechnologies have revealed a
distinct underlying regulatory system and gene expres-
sion patterns during biofilm phases related to plank-
tonic counterparts, which reflect their unique
phenotypes (Bruns et al. 2010; Muszkieta et al. 2013). It
is known that biofilm and planktonic phases have dis-
tinct growth demands and this is reflected by the differ-
entially regulation of many genes and proteins such as
secondary metabolism, cell wall structure, transcription
factors, stress responses, and the translation machinery
(Gibbons et al. 2012).
Proteomic and transcriptomic analyses of A. fumiga-
tus biofilm have showed a stage dependent pattern
with different energy requirements, depending on the
age of the biofilm (Bruns et al. 2010). Young biofilm
(24 h), has higher demands of energy through upregu-
lation of primary metabolism genes (Bruns et al. 2010),
which are required processes for growth, maintenance,
Figure 1. The schematic stepwise phases of the development of filamentous fungi biofilms on biotic and abiotic surfaces. The
Figure has been made by Biorender.com.
6 M. ROUDBARY ET AL.
and survival of the organism (Drew and Demain 1977).
Comparatively, the metabolic activity of matured bio-
film (48 h), has been diminished (Bruns et al. 2010). In
contrast, the hydrophobins family, including RodAp and
RodBp (Paris et al. 2003), and secondary metabolites
such as gliotoxin, an immunosuppressive toxin (Sugui
et al. 2007), have been upregulated in the matured bio-
film (Bruns et al. 2010). The enhanced production of
gliotoxin in the matured biofilm stage may protect A.
fumigatus from the host immune system and enable its
survival and persistence in chronic lung infections such
as aspergilloma (Bruns et al. 2010). The substantial up-
regulation of secondary metabolism genes in A. fumiga-
tus biofilms could be associated with the upregulation
of LAEA, a secondary metabolism regulator (Gibbons et
al. 2012) which has also been reported to regulate the
synthesis of gliotoxin (Sugui et al. 2007). Two types of
proteins, hydrophobins and adhesins,confer a hydro-
phobic character to fungal morphotypes (Sunde et al.
2008; Gibbons et al. 2012). Hindering the immune rec-
ognition is another role attributed to hydrophobins,
RodAp, which absence is associated with a reduction of
virulence in A. fumigatus (Aimanianda et al. 2009; Bruns
et al. 2010). The upregulated hydrophobin expression
in biofilm cells, including several Rod genes (RodA,
RodB, and RodD) (Beauvais et al. 2007; Gibbons et al.
2012) supports the need for further studies to elucidate
the roles of hydrophobins in biofilm.
Adhesins have not been deeply investigated bio-
chemically in filamentous fungi; however, it is reported
an upregulation of 11 of the 25 genes with the highest
adhesion probability score (Afu1g09510, Afu1g14430,
Afu3g00420, Afu3g01150, Afu3g09690, Afu4g03240,
Afu6g13720, Afu7g00580, Afu7g02460, Afu7g05340, and
Afu8g01970)inA. fumigatus genome (Upadhyay et al.
2009; Gibbons et al. 2012). Since adhesion is the first
and one of the most crucial steps in the colonisation of
a substrate by a fungus (Sheppard 2011), this ability
should be further explored. On the other side, the role
of mitogen-activated protein kinases (MAPK) and phos-
phatases in adhesion and biofilm formation of A. fumi-
gatus, as well as the cell wall and ECM production, has
been explained by the loss of the MAPK genes (MpkA,
MpkC, and SakA), and phosphatase genes (sitA and
ptcB) (Manfiolli et al. 2018). Indeed, the loss of particular
MAP kinases and phosphatases cascades, which operate
different responses such as cell growth, differentiation,
and environmental stress adaptation based on changes
in certain environmental conditions (Chen and Thorner
2007), influences the cell wall carbohydrate exposure
and the ECM during biofilm formation and, subse-
quently, diminished adherence of A. fumigatus to
polystyrene and fibronectin-coated plates. Considering
the role of MpkC and SakA in A. fumigatus, the associ-
ation of MAPK and the high-osmolarity glycerol (HOG)
pathway involved in osmotic and oxidative stress
response, a biofilm regulation could be proposed (Chen
and Thorner 2007; Nascimento et al. 2016). It has also
been showed that the deletion of ptcB (one of the
phosphatases which promotes Hog1 phosphorylation),
in A. fumigatus, results in the increased B-1,3-glucan
and chitin content but more sensitivity to cell-wall dam-
aging drugs, reduced adhesion, biofilm formation, and
ECM production, which directly implies the association
of the HOG pathway in the regulation of biofilm forma-
tion (Ribes et al. 2000).
In A. niger, the time-dependent regulation manner of
HOG1 pathways genes, revealed the role of HOG path
in upregulation of genes of at different times in biofilm
state (Sun et al. 2020). Besides, biofilm cells and free
mycelia of A. niger demonstrated an enhanced activity
regarding the extracellular polysaccharide synthesis
pathway, and amino acid metabolisms (Sun et al. 2020).
It was also indicated the downregulation of the tricarb-
oxylic acid cycle (TCA cycle) at 24 h, which resulted in
the acceleration of amino acid accumulation, also
reflects the higher activity of biofilm (Sun et al. 2020).
The role of the Ca
2þ
mediated calcineurin signalling
pathway (CSP) on the biofilm formation of A. niger has
also been investigated (Liu et al. 2020). CSP is an essen-
tial signalling pathway regulating Ca
2þ
balance in the
mycelium of filamentous fungi. Knocking out several
genes involved in CSP, including the Ca
2þ
channel
MidA and CchA, calcineurin catalytic subunit CnaA, and
transcription factor CrzA, led to the reduction of both
Ca
2þ
levels in the mycelium and biofilm formation. The
deletion of the mentioned genes affected the biofilm
formation by decreasing the expression of hydrophobin
RodA and, consequently, the hydrophobicity and adhe-
sion of conidia. Actually, the hydrophobic ability of the
conidia influences the adsorption, which is a critical
step in the early stage of formation, immobilisation and
development of biofilm (Liu et al. 2020). Additionally,
those authors revealed that there is an increased sensi-
tivity of mutant strains to cell wall disrupters, compared
to wild-types. This study brought up the possibility of
different extracellular polysaccharides and cell surface
components, such as b-1,3-glucan and chitin, and their
influence on flocculation ability of hyphae, cell-cell
adhesion, and aggregation ability (Liu et al. 2020). The
lost mycelial flocculation ability of mutant strains, rela-
tive to the wild-types, provided proof for changes in
mycelium adhesion and aggregation, proposing an
impact of CSP on biofilm formation by altering the
CRITICAL REVIEWS IN MICROBIOLOGY 7
composition of mycelial cell carbohydrates in the mid-
dle stage of biofilm formation (Liu et al. 2020). In the
following, the altered expression of more than 50% of
cell wall genes in A. fumigatus biofilm (Fanning and
Mitchell 2012), and the absence of cell wall in human
hosts suggest the cell wall biogenesis pathways as a
potential target for specific disrupter. Moreover, the
role of the cspA gene, encoding a glycophosphatidyli-
nositol-anchored protein in the cell wall, in biofilm
development, and antifungal drug resistance in A. fumi-
gatus has been described (Fan et al. 2015). It was indi-
cated that the cspA deletion resulted in a reduced
biofilm formation, decreased resistance to antifungal
agents, and increased internalisation by A549 human
lung epithelial cells. Also, the disruption of cspA
resulted in an irregular arrangement of hyphae, circle-
like connections instead of the classic bundles of paral-
lel filaments forming acute angles, which might affect
biofilm construction. Comparing with the wild-type
cells, the absence of ECM in the mutant strain, sug-
gested that cspA is linked to the ECM production, and,
subsequently, with drug resistance (Fan et al. 2015).
Recently, the influence of oxygen on the biofilm
morphotypes of A. fumigatus exhibited a hypoxia-typic
morphotype (Kowalski et al. 2019), which has been gen-
erated by the expressed sub-telomeric gene cluster,
containing genes that modify the hyphal surface and
disturb inter-hyphal interactions. Hypoxia-morphotype
leads to increased host inflammation, rapid disease pro-
gression, and mortality in a murine model of invasive
aspergillosis, suggesting a direct impact of filamentous
fungal biofilm morphology on fungal-host interactions
(Kowalski et al. 2019).
Besides, on specific pathways, biofilms display a
reduced susceptibility to antifungal drugs that can also
be involved in the poor in vivo drug efficacy, which
seem to be explicitly expressed (Bruns et al. 2010;
M
uller et al. 2011; Rajendran et al. 2011; Gibbons et al.
2012; Muszkieta et al. 2013). For instance, MDR4, a mul-
tidrug resistance transporter gene in A. fumigatus,
showed to have a phase-dependent expression manner
in biofilm, upregulating in both 12- and 24-h, with max-
imal upregulation at 12-h. Furthermore, after treatment
with voriconzaole, A. fumigatus has shown an increased
resistance in the 12-h (16- to 128-fold) and 24-h (>512-
fold) phases, when compared with 8-h germlings, which
is in agreement with the phase expression pattern
(Rajendran et al. 2011). Given the expression profile of
MDR4, the efflux pump is likely to play an early function
in biofilm by regulating the internal environment
through removing antifungals. Since ECM is produced
during the maturation phase of biofilm, it is therefore
likely that A. fumigatus might use the ECM to decrease
the drug penetration, suggesting it as a reason for the
phase-dependent expression of MDR4. Additionally,
several mechanisms might be responsible for a high
resistance of biofilms (compared with the planktonic
state), as discussed later in section 5.
Biofilm formation is thought to be a single process
but it has become clear that biofilm development
involves a series of sequential molecular events.
Shedding more light on the underlying regulatory net-
work of intracellular signalling pathways involved in the
biofilm, will provide new avenues for research into the
identification of potential therapeutic targets.
5. Quorum-sensing molecules and their actions
in filamentous fungi
Apart from survival and reproduction competition,
microorganisms are no longer considered independent
cells (Kalia 2015; Barriuso et al. 2018; Padder et al.
2018). They have developed a communication manner -
quorum sensing (QS) - that allows them to orchestrate
an appropriate response to local environmental
changes (Kalia 2015; Barriuso et al. 2018; Padder et al.
2018). QS is a process that links gene expression to the
cell density of microbial populations by secretion of
small metabolites (Hogan 2006; Srinivasan et al. 2016).
The cell-cell communication in QS relies on the produc-
tion and release of small diffusible chemical signalling
molecules in the extracellular environment, referred to
as autoinducers or quorum-sensing molecules (QSMs)
(Ramage et al. 2009; Wongsuk et al. 2016). QSM serve
as transcriptional regulators; while their concentration
reaches a critical threshold, corresponding to an appro-
priate cell density, they bind to the receptors (Waters
and Bassler 2005; Padder et al. 2018; Zhang et al. 2019).
Subsequently, the target genes of QS and the genes
encoding QS signal synthesis are prompted by tran-
scriptional regulators, thereby generating self-inducing
feedback, rising the production of the corresponding
signalling molecules (Hogan 2006; Williams et al. 2013).
The existence of QS systems in fungi has been recog-
nised by farnesol, QSM of C. albicans (Hornby et al.
2001; Rodrigues and
Cern
akov
a2020). Many QSMs are
actively involved in fungal QS, contributing to various
vital mechanisms in microorganisms, including regula-
tion of virulence genes of pathogenesis and dissemin-
ation during the infection establishment, adaptation of
microorganisms to host, biofilm development, morpho-
genesis, secondary metabolite production, stress resist-
ance, and suppression of the cell population (Ramage
et al. 2005; Dixon and Hall 2015; Wongsuk et al. 2016;
8 M. ROUDBARY ET AL.
Barriuso et al. 2018; Bandara et al. 2020). However,
much research has focussed on QS in bacteria and
yeast, particularly C. albicans, but QS in non-unicellular
filamentous fungi remains controversial due to the
nature of these fungi that develop mycelial growth.
There are several QSMs in filamentous fungi, particu-
larly in species from the genera Aspergillus and
Penicillium. Among them, we can find oxylipins, butyro-
lactone I, linoleic acid, c-heptalactone, multicolic acid,
farnesol (Table 1).
Oxylipins are oxygenated polyunsaturated fatty
acids, made by oxygenase enzymes (Tsitsigiannis and
Keller 2007; Brodhun and Feussner 2011). In A. flavus,
oxylipins serve as a signal that regulates secondary
metabolisms (mycotoxin production), and morpho-
logical differentiation and spore development either
asexual spores or sclerotia via a density-dependent
mechanism resembling QS (Horowitz Brown et al. 2008;
Brown et al. 2009; Affeldt et al. 2012). Increased cell
density resulted in the lowest numbers of sclerotia, and
the highest numbers of conidia, suggesting density-
dependent phenomena (Horowitz Brown et al. 2008).
Moreover, deletion of Aflox, encoding an oxylipin-gen-
erating lipoxygenase (LOX), showed that there is a cor-
relation between LOX activity and density-dependent
development of both sclerotia and conidia, suggesting
the possible existence of a dense dependent factor,
maybe a LOX-derived metabolite stimulates the conid-
ium-to-sclerotium switch, depending on the high or
low density of cells (Horowitz Brown et al. 2008). In
order to elucidate how oxylipins, as signals, are
detected and transmitted by A. flavus and involved in
QS mechanisms, the role of G protein-coupled recep-
tors (GPCRs) has been investigated (Affeldt et al. 2012).
Depletion of GprC and GprD in A. flavus resulted in lock-
ing of fungus in a low-density state, which revealed the
necessity of these receptors to respond and transit to a
high-density development state (Affeldt et al. 2012). It
also has been revealed that oxylipin acts as develop-
mental signals through GPCRs and induces fungal cellu-
lar differentiation, including lateral branching (Niu et al.
2020), mediating the interhyphal communications by
fusion to other hyphae, providing more substrate inva-
sion by mechanical strength, and more molecular and
nutrient exchanges within the network (Fischer and
Glass 2019; Harris 2019). These findings propose a cas-
cade-like regulation process, which leads to the morph-
ology and metabolic differentiation, more branching
and nutrient exchange, that might be arranged for
effective competition and accessing multiple nutrients
in complex environments, such as biofilms.
Butyrolactone I, a secondary metabolite produced by
A. terreus, has been considered as a QSM, according to
its auto-stimulatory effect (Schimmel et al. 1998; Raina
et al. 2012). Butyrolactone I influences its own produc-
tion and induces morphological changes, increases
hyphal branching, spore formation (Schimmel et al.
1998; Raina et al. 2012). Also, further analysis uncovered
the regulatory role of butyrolactone I in the expression
level of lovastatin (Palonen et al. 2017). LaeA, a global
Table 1. QSMs and their known roles in clinically relevant filamentous fungi.
QSMs Organism Role References
Oxylipins Aspergillus flavus -Regulation of secondary metabolism (mycotoxin
production), morphological differentiation and
spore development.
-Cellular differentiation, like lateral branching.
(Horowitz Brown et al. 2008; Brown et
al. 2009; Affeldt et al. 2012; Niu et
al. 2020)
Butyrolactone I Aspergillus terreus -Auto-inducer effect on Butyrolactone I
production.
-Induction of morphological changes, increase in
hyphal branching, spore formation, and
regulation of lovastatin production.
(Schimmel et al. 1998; Raina et
al. 2012)
c-heptalactone Aspergillus nidulans -Shortening of the lag phase, earlier
deceleration phase, and enhanced
penicillin production.
(Williams et al. 2012)
Multicolic acid Penicillium sclerotiorum -Increase of sclerotiorin production. (Raina et al. 2010)
Farnesol (exogenous) Aspergillus fumigatus -Inhibition of the cell wall integrity pathway by
direct or indirect production of cell wall
stressor.
-Disorganisation of the apical actin cytoskeleton.
(Dichtl et al. 2010)
Farnesol (exogenous) Aspergillus nidulans - Induction of apoptosis. (Semighini et al. 2006)
Farnesol (exogenous) Aspergillus niger -Abolishion of conidiation.
-Increase of aerial hyphae formation.
-Decrease of the intracellular level of cyclic
adenosine monophosphate (cAMP).
(Lorek et al. 2008)
Farnesol (exogenous) Aspergillus falvus - Induction of apoptosis. (Wang et al. 2014)
Farnesol (exogenous) Fusarium graminearum - Induction of apoptosis. (Semighini et al. 2008)
Farnesol (exogenous) Histoplasma capsulatum -Synergistic effect with itraconazole on biofilm
and planktonic phases.
(Brilhante et al. 2015)
CRITICAL REVIEWS IN MICROBIOLOGY 9
regulator, is involved in conidiation and secondary
metabolism along with the velvet family.
The overexpression of LaeA, induced by butyrolactone I,
increased the biogenesis of lovastatin, supporting the
suggested role of butyrolactone I as a QS molecule in A.
terreus (Palonen et al. 2017). Moreover, lovastatin pro-
duction and the expression levels of lovastatin biosyn-
thetic genes, lovB and lovF,inA. terreus have been
increased by exogenous linoleic acid supplementation,
as an oxylipin precursors (Sorrentino et al. 2010).
However, supplementing linoleic acid to the low cell
density of A. terreus culture enhanced lovastatin pro-
duction to 1.8-fold. Nevertheless, lovastatin yield has
been impaired by adding at a later stage of growth,
suggesting the regulatory role of cell density in lovasta-
tin production at the time of linoleic acid supplementa-
tion (Sorrentino et al. 2010).
c-heptalactone is a lactone-containing QSM that reg-
ulates the A. nidulanss behaviour at a low cell density
(Williams et al. 2012). The supplementation of low-dens-
ity cell cultures of A. nidulans with nanomolar concen-
trations of c-heptalactone resulted in the shortening of
the lag phase, earlier deceleration phase, and enhanced
penicillin production via increased ipnA::lacZ reporter
activity, a penicillin production marker (Williams et al.
2012). Further analysis in A. nidulans showed that the
oleic and linoleic acid-derived oxylipins production
depends on the regulation of three conserved oxylipin
biosynthetic enzymes, PpoA, PpoB, and PpoC. A. nidu-
lans oxylipin mutation (DppoA; DppoB; DppoC A)
resulted in the defective colonisation of seeds, reduced
lipase production, altered secondary metabolite profiles
and impaired mycotoxin sterigmatocystin production
(Tsitsigiannis and Keller 2006). The silenced mutation in
ppo genes, ppoA,ppoB, and ppoC,inA. fumigatus,
revealed that they are involved in hypervirulence in the
invasive pulmonary aspergillosis murine model and
increased tolerance to H
2
O
2
stress (Tsitsigiannis et al.
2005). These findings imply that Ppo products, oxylipins
(already mentioned above), serve as activators of host
defense mechanisms by hindering the development of
pulmonary and invasive aspergillosis (Tsitsigiannis et al.
2005). c-butyrolactone-containing compounds (multi-
colic acid and related derivatives) have been suggested
as the QSMs in Penicillium sclerotiorum, by increasing
sclerotiorin production, a secondary metabolite (Raina
et al. 2010).
The QSM farnesol, a precursor of sterol biosynthesis
in C. albicans, prevents the biofilm development and
morphological transition in Candida spp. (Hornby et al.
2001; Hogan 2006; Rodrigues and
Cern
akov
a2020).
Exogenous farnesol affects filamentous fungi such as A.
fumigatus (Dichtl et al. 2010), A. nidulans (Semighini et
al. 2006), A. niger (Lorek et al. 2008), A. flavus (Wang et
al. 2014), Fusarium graminearum (Semighini et al. 2008)
and Histoplasma capsulatum (Brilhante et al. 2015).
In A. fumigatus, exogenous farnesol has been impli-
cated as a cell wall perturbing agent that inhibits the
cell wall integrity (CWI) pathway, originating direct or
indirect cell wall stress (Dichtl et al. 2010). Mutants lack-
ing two essential kinases of CWI signalling, AfMkk2 or
AfMpkA, namely the MAP kinase kinase AfMkk2 or its
target MAP kinase AfMpkA, revealed high susceptibility
to farnesol (Dichtl et al. 2010). The higher susceptibility
of mutants compared with the wild type suggested
that the mode of action of farnesol is not only restricted
to CWI signalling inhibition and could be related to
interfering with prenylated proteins. The loss of hyphal
tip localisation of AfRho1 and AfRho3, two prenylated
Rho family GTPases involved in CWI and cytoskeleton
controlling, implied disorganisation of the apical actin
cytoskeleton, as complementary part in higher suscepti-
bility of mutant to farnesol (Dichtl et al. 2010). In A.
nidulans, farnesol has been shown to induce apoptosis,
a programmed cell death, through a mechanism that
demands functional mitochondria, ROS production, and
mediating a heterotrimeric G protein complex (GTP-
binding proteins), FadA (Semighini et al. 2006). Farnesyl
pyrophosphate (FPP), a protein prenylation precursor,
plays a role in the post-translation modification of pro-
teins. Farnesylated proteins include Ras and Ras-related
G proteins, control cell growth, differentiation, prolifer-
ation, and survival (Edwards and Ericsson 1999).
Additionally, co-cultivation of A. nidulans with C. albi-
cans revealed an impaired hyphal growth and colony
development in a farnesol-dependent manner
(Semighini et al. 2006). Since A. nidulans does not
secrete detectable amounts of farnesol, this may sug-
gest that it reacts to farnesol produced by other fungi
(Semighini et al. 2006). In addition to its QS function
(morphogenesis regulation), farnesol seems to be
employed by C. albicans to reduce the competition
with other microorganisms in a competitive situation.
Exogenous farnesol produced by C. albicans serves as a
conidiation inhibitor in A. niger (Lorek et al. 2008). The
high concentration of farnesol resulted in the abolished
conidiation and increased aerial hyphae formation with
a cotton-like appearance, and the diminished intracellu-
lar level of cyclic adenosine monophosphate (cAMP)
(Lorek et al. 2008). Moreover, A. flavus treated with far-
nesol had inhibition of germination and growth, and
induction of apoptosis markers, including nuclear con-
densation, phosphatidylserine (PS) externalisation, DNA
fragmentation, and ROS generation, metacaspase
10 M. ROUDBARY ET AL.
activation, and abnormal cellular ultrastructure (Wang
et al. 2014). Another study demonstrated that farnesol
in Fusarium graminearum triggers apoptosis through
impairing development and inhibition of macroconidial
germination and viability (Semighini et al. 2008).
Farnesol also affects filamentous fungus, H. capsulatum,
in biofilm and planktonic phases, particularly in com-
bination with itraconazole (Brilhante et al. 2015), sug-
gesting farnesol as a promising antifungal.
QSMs can affects the host immune system by several
means (Joo and Jetten 2008). For example, farnesol
induces the expression of some immune responses and
inflammatory genes in lung cells through the nuclear
factor kappa-B (NF-jB) signalling pathway, including IL-
6,CXCL3,IL-1a, and COX-2 (Joo and Jetten 2008). The
NF-jB pathway stimulation via MEK1/2-ERK1/2-MSK1-
dependent phosphorylation of p65 l leads to the pro-
duction of cytokines (interleukin-6 and interleukin-1a)
(Joo and Jetten 2008). Also, farnesol with yeast cell wall
components (zymosan), intensifies the expression of
proinflammatory cytokines in the murine macrophage
cell line but inhibits activation of cellular immunity, sug-
gesting farnesol as an immune-modulatory molecule
(Oxford Academic 2020b). On the other hand, the mat-
uration of monocytes to dendritic cells has been
affected by farnesol (Leonhardt et al. 2015). Matching
with the control, in the presence of farnesol, the cell
surface markers of immature dendritic cells were modi-
fied, including increased CD86 and reduced CD1a, low-
ering expression of multiple genes involved in cell
adhesion and migration (Leonhardt et al. 2015).
Therefore, there was a diminished ability of dendritic
cells to activate T cells, which eventually dampens the
adaptive immune response (Leonhardt et al. 2015).
Additionally, it has been concluded that QSMs
impact the stress resistance in fungi, by preserving the
fungus from ROS. In C. albicans, resistance to ROS has
been enhanced by the increased expression of protect-
ive catalase Cat1, due to the inhibition of the Ras1-
cAMP pathway and cross-talk with Hog1 regulators
(Westwater et al. 2005; Deveau et al. 2010).
Taken together, QS process seems to have been
adapted by fungi to regulate a range of developmental
processes gene expression and coordinate cell-to-cell
interactions, to optimise a successful cellular morph-
ology and metabolic differentiation for effective compe-
tition in a complex environment. Therefore, further
elucidation of the QS mechanisms in filamentous fungal
biofilm, as a tangled network, will lead to a better
understanding of fungal pathogenesis and facilitating
the development of novel antifungal tar-
get approaches.
6. Antifungal resistance phenomenon in
biofilm structure and possible mechanisms
Fungal biofilm resistance is a multifactorial phenom-
enon, including fundamental physical barriers and com-
plex regulatory processes (Ramage et al. 2012;
Shishodia et al. 2019; Kowalski et al. 2020). Different
resistance mechanisms in the fungal biofilm might be
developed to decrease the efficacy of antifungal treat-
ment, including ECM production, modifications or over-
expression of target molecules, compound expulsion by
efflux pumps, restricted diffusion, tolerance, and cell
density (Niimi et al. 2010; Ramage et al. 2012). The fol-
lowing sections will describe the mechanisms involved
in the increased antifungal resistance of the filamentous
fungal biofilms (Table 2).
6.1. Major cases of drug resistance
Biofilm structures are highly resistant to antimicrobial
agents and hence giving rise to persistent infections in
patients. Treating infections caused by biofilms require
a higher dosage of antifungal agents and different
therapeutic schemes, since structural attributes, such as
ECM, will dramatically reduce the efficacy of the drug
used (e.g. diffusion issues) (Rodrigues et al. 2017;
Shishodia et al. 2019). The inefficacy of several antifun-
gal agents against fungal biofilms has been frequently
reported in several studies (Mowat et al. 2007;2008;
Seidler et al. 2008; Fiori et al. 2011; Liu et al. 2012;
Zhang et al. 2012; Pierce et al. 2013; Wuren et al. 2014;
Kirchhoff et al. 2020; Nazir et al. 2020). Compared to
planktonic cells, biofilm or sessile cells of Aspergillus
spp. (A. fumigatus,A. flavus,A. terreus, and A. niger)
exhibited higher sessile minimum inhibitory concentra-
tions 90 (SMIC90) of amphotericin B, voriconazole, and
echinocandins, including anidulafungin and caspofun-
gin (Fiori et al. 2011). Further investigation on the
developed biofilm of A. fumigatus on bronchial epithe-
lial cells revealed a decreased antifungal drug suscepti-
bility towards azoles, polyenes, and echinocandins
(Seidler et al. 2008). For instance, the efficacy of itracon-
azole, voriconazole, amphotericin B, and caspofungin
against biofilm of A. fumigatus, was 1000 times higher
than that of the planktonic cells (Mowat et al. 2007).
Besides, exposure of Aspergillus spp. to serum proteins,
such as foetal bovine serum (FBS), fetuin-A, and bovine
serum albumin (BSA), led to the formation of a thick
ECM and increased resistance to antifungals (Wuren et
al. 2014). Similarly, F. solani and F. oxysporum biofilms
also displayed reduced susceptibility to all tested anti-
fungal agents, including amphotericin B, voriconazole,
CRITICAL REVIEWS IN MICROBIOLOGY 11
itraconazole and fluconazole (Zhang et al. 2012;
Machado Vila et al. 2015).
6.2. Physiological state and cell density
The physiological state of cells in biofilm communities
has also been associated with the susceptibility profiles
of biofilms. Different studies explained how environ-
mental stresses modify biofilm architecture and likely
antifungal sensitivity, including glucose and iron limita-
tion (Baillie and Douglas 1998), elastin addition
(Brand~
ao et al. 2018), growth rate (Beauvais and Latg
e
2015), oxygen availability (Dumitru et al. 2004; Kowalski
et al. 2019), pH (Kuchar
ıkov
a et al. 2011), and tempera-
ture (Pettit et al. 2010). It was shown that self-induced
hypoxic microenvironments in A. fumigatus during bio-
film maturation facilitate fungal survival in the presence
of antifungal treatments (Kowalski et al. 2020).
Decreasing oxygen levels in the biofilm of A. fumigatus
increase antifungal drug resistance (Kowalski et al.
2020). Therefore, it is more likely that multiple factors
are implicated in the physiological state of the cell,
Table 2. Biofilm associated resistance mechanisms of filamentous fungi.
Mechanism of resistance Effect References
Altering the physiological state of cells Biofilm architecture modification (Baillie and Douglas 1998; Dumitru et al. 2004;
Kuchar
ıkov
a et al. 2011; Beauvais and Latg
e
2015; Brand~
ao et al. 2018; Kowalski et al.
2019,2020)
Cell density Quorum sensing (Lawrence et al. 1991; Beer et al. 1994; Lass-Fl
orl
et al. 2003; Perumal et al. 2007; Mowat et al.
2008; Costa-Orlandi et al. 2017)
Extracellular matrix (ECM) Preventing penetration of antifungals into the
biofilm and reaching the targets
(Beauvais et al. 2007; Fontaine et al. 2010;
Loussert et al. 2010; Rajendran et al. 2013;
Bom et al. 2015; Sheppard and Howell 2016)
Up-regulation efflux pumps and ATP-binding-
cassette transporter proteins
Reducing the intracellular performance of
antifungals through pumping out
(Ferreira et al. 2006; Bueid et al. 2010; Rajendran
et al. 2011; Gibbons et al. 2012; Ramage et al.
2012; Muszkieta et al. 2013)
Heat shock proteins 90 (HSP90) Increased drug resistance through possible
enhanced reactive oxygen species capacity
(Robbins et al. 2011; Delattin et al. 2014;Tuet
al. 2020)
Drug tolerance and persister cells Inactivation antifungal drug targets by entering
the dormancy phase
(Schumacher et al. 2009; Lafleur et al. 2010;
Lewis 2010; Wuyts et al. 2018)
Table 3. Summary of filamentous fungal biofilm and their clinical relevance.
Biofilm former Biofilm morphology Matrix component
Devices associated
with biofilm References
Aspergillus spp. Hyphal cells embedded
in ECM
Galactomannan, a-1,3-glucans,
monosaccharides, polyols,
melanins, proteins such as
hydrophobins and DNA
Catheters, cardiac
pacemakers, joint
replacements, heart
valves, and breast
augmentation implants
(Beauvais et al. 2007; Loussert et
al. 2010; Escande et al. 2011;
Jeloka et al. 2011; Sardi et al.
2014; Chatterjee and Das 2020)
Zygomycetes spp. Compact hyphal structures
encased in ECM
ECM contained
N-acetylglucosamine and
glucosamine
Catheters (Singh et al. 2011; Almeida et al.
2018; Ezzat et al. 2020)
Black yeast-like fungi Species-depending high
biomass biofilm
Water distribution and
supplies systems
(Hilmarsdottir et al. 2005;
Heinrichs et al., 2013a,2013b;
Gao et al. 2016; Kirchhoff et
al. 2017)
Histoplasma capsulatum Dense biofilm of yeast, as
an active biofilm
structural contributor
ECM with
polysaccharide content
Probably healthcare-
associated
biomaterial devices
(Pitangui et al. 2012; Carreto-
Binaghi et al. 2015; Gonc¸alves
et al. 2020)
Scedosporium spp. Dense compact of the
mycelial biomass with
internal channels
enclosed in ECM
ECM consisted of carbohydrate-
rich molecules and
extracellular DNA
Central venous catheter (Mello et al. 2016; Kitisin et al.
2020; Mello et al. 2020)
Fusarium spp. The compacted architecture
of hyphal surrounded
by ECM
ECM composed of extracellular
DNA, RNA, polysaccharides,
and proteins
Contact lenses (Mukherjee et al. 2012; Sav et al.
2018;C
ordova-Alc
antara et al.
2019; Kischkel et al. 2020)
Dermatophyte The dense and thick
network of hyphae,
embedded in EPS
Water channel and
host tissue(hair)
(Costa-Orlandi et al. 2014; Danielli
et al. 2017; Brilhante et al.
2019; Costa-Orlandi et al. 2020)
Sporothrix schenckii A Network of hyphae,
embedded in EPS
EPS composed of glycoprotein
(mannose rich),
carbohydrates, lipids, and
nucleic acid
(S
anchez-Herrera et al. 2021)
in A. fumigatus
12 M. ROUDBARY ET AL.
suggesting a possible role in resistance to antifungal
agents of biofilm.
Cell density is also considered as one of the resist-
ance mechanisms within the complex fungal biofilms.
Previous investigations support the idea that the high
density of the cells, which can be regulated by QS, pro-
duces resistance to antifungal agents (Lass-Fl
orl et al.
2003; Perumal et al. 2007; Mowat et al. 2008). Relatively,
the low cell density (10
3
cells/mL) of both planktonic
and resuspended biofilm cells showed susceptibility to
azoles. However, ten folds increased cell density leads
to the enhanced resistance to azoles (Perumal et
al. 2007).
6.3. ECM
ECM is an essential component of all fungal biofilms,
rendering physical barrier function against environ-
mental insults, like host defences and antimicrobial
therapies (Mitchell et al. 2016; Pierce et al. 2017;Le
Mauff 2020). The antifungal drug permeation into
the biofilm or binding to the target can be ham-
pered by the presence of a dense ECM (Nett and
Andes 2017; Rodrigues et al. 2017; Singh et al.
2018), suggesting the possible role of the chemical
composition of the ECM and its regulation in the
resistance mechanisms. A typical ECM consists of
polysaccharides (e.g. galactomannan, a1,3-glucans,
monosaccharides), polyols, melanin, eDNA, lipids (e.g.
ergosterol) and proteins (including antigens and
hydrophobins), however, its composition varies sig-
nificantly based on the growth conditions and the
species (Beauvais et al. 2007; Loussert et al. 2010;
Fiori et al. 2011). In A. fumigatus biofilms, ECM is a
heterogeneous substance that consists of extracellu-
lar DNA, proteins, lipids and polyols, and exopolysac-
charides (e.g. a-glucans, galactomannan, and the
glycan galactosaminogalactan (GAG)) (Beauvais et al.
2007; Fontaine et al. 2010; Loussert et al. 2010;
Rajendran et al. 2013; Sheppard and Howell 2016).
GAG is a partially deacetylated heteropolymer of
a-1,4-linked galactose and N-acetyl galactosamine
(Lee et al. 2014) and is absent in A. fumigatus
spores but exists in growing hyphae. It mediates
protection against host immune defences and also
accelerates the adhesion between both fungi and to
other surfaces (Fontaine et al. 2011; Gravelat et al.
2013; Lee et al. 2016). Deficiency in GAG production
in A. fumigatus fails to produce ECM (the basis of
biofilm formations), indicating a critical role of GAG
in the maintenance of the ECM of A. fumigatus bio-
films (Gravelat et al. 2013; Bamford et al. 2015; Lee
et al. 2015). Therefore, the attenuated virulence of
GAG deficient A. fumigatus strains in the invasive
aspergillosis mouse model, explains the functional
role of this component in evading host defences
(Gravelat et al. 2013; Lee et al. 2015).
Galactofuranose-deficient mutants showed a higher
ability to form biofilms, as a result of enhanced pro-
duction of GAG, confirming the insignificance role of
galactomannan in biofilm formation (Lamarre et al.
2009; Gravelat et al. 2013; Lee et al. 2014). GAG pro-
duction is regulated by the developmental regulatory
proteins MedA (Gravelat et al. 2010; Al Abdallah et
al. 2012), StuA (Gravelat et al. 2013), and SomA (Lin
et al. 2015; Chen et al. 2020). The deletion of the
phosphatases SitA or PtcB through the activation of
cell wall integrity has been linked to the destruction
of biofilm formation and extracellular matrix produc-
tion (Bom et al. 2015; Winkelstr
oter et al. 2015).
Additionally, A. fumigatus biofilm cells attached to
epithelial cells exhibited raised levels of ECM, with a
coincidental decreased sensitivity to antifungal drugs
(Seidler et al. 2008). Whereas the precise role of the
ECM is not completely understood, it is thought that
the matrix of Aspergillus biofilms contributes to the
higher resistance, by either slowing drug penetration
or direct drug sequestration (Beauvais et al. 2007;
Bugli et al. 2013; Rajendran et al. 2013; Rodrigues et
al. 2017). Moreover, the increased ECM in matured
A. fumigatus biofilm suggests it as a drug resistance
promoter during the later phases of biofilm growth
(Beauvais et al. 2007). Biofilm of A. niger has also
shown an increased thickness of ECM at the matur-
ation stage, proving ECM as the principal constituent
(Bandara et al. 2020).
Apart from exopolysaccharides, eDNA, which is
another key component of the ECM, also appears to
have an impact on drug resistance through providing
structural integrity (Rajendran et al. 2013). The
increased content of eDNA in matured A. fumigatus
biofilms (8 to 48 h) has been confirmed as a genomic
DNA (Rajendran et al. 2013). The association between
eDNA and chitinase activity is a probable conse-
quence of autolysis, conferring the structural support
for the development and maintenance of the com-
plex biofilm architecture (Rajendran et al. 2013).
Furthermore, degradation of eDNA by DNase destabil-
ises the biomass and the integrity of the biofilm and
renders biofilms to become more susceptible to both
amphotericin B and caspofungin (Shopova et al.
2013). However, the mechanisms responsible for the
secretion of DNA are unknown but the integrated
CRITICAL REVIEWS IN MICROBIOLOGY 13
DNA in the ECM of biofilms may develop a more
resistant structural biofilm.
6.4. Efflux pumps and ATP-binding-cassette
transporter proteins
The reduced susceptibility of fungal biofilm to drugs
could be the result of the increased activity of multi-
drug resistance (MDR) pumps, including ATP-binding-
cassette (ABC) and the major facilitator superfamily
(MFS). Frequent exposure to antifungal drugs in the
clinical settings, such as aspergilloma cases, may
enhance the levels of efflux pump expression, thereby
either contributing to or inducing clinical resistance
(Bueid et al. 2010). Genome analysis of A. fumigatus
revealed MDR pumps have been associated with
increased resistance to azoles (Ramage et al. 2012;
Muszkieta et al. 2013). In biofilm conditions, 140 efflux
pump genes, including the ABC transporters MDR1,
MDR2, and MDR4 were upregulated (Ramage et al.
2012; Muszkieta et al. 2013). The highest increase in
resistance in A. fumigatus mature biofilms correlated
with the maximum expression of the MDR4 family
(Ramage et al. et al. 2012; Muszkieta et al. 2013). In a
subcutaneous A. fumigatus biofilm-related mouse
model, the in vivo expression of MDR4 was upregulated
in response to voriconazole treatment (Rajendran et al.
2011). The MFS, azole resistance-related genes, and the
members of the MDR1 family have been significantly
upregulated in A. fumigatus biofilms (Ferreira et
al. 2006).
6.5. Heat shock proteins 90 (HSP90)
The molecular chaperone heat shock proteins 90
(Hsp90) has been implicated as a principal regulator of
biofilm dispersion and drug resistance (Robbins et al.
2011). Inhibition of Hsp90 by geldanamycin, Hsp90
inhibitor, led to the reduction of resistance of A. fumiga-
tus biofilms to echinocandins (Robbins et al. 2011).
Moreover, vorinostat (suberoylanilide hydroxamic acid,
SAHA), a novel histone deacetylase inhibitor, showed
synergistic effects with azoles, including itraconazole,
voriconazole, and posaconazole, against both biofilm
and planktonic cells of Aspergillus spp. (Tu et al. 2020).
The synergistic effect might be partly associated with
dampened expression of the efflux pump genes and
largely via the inhibition of Hsp90 (Tu et al. 2020).
Given that biofilm formation in C. albicans is associated
with an increased capacity against oxidative stress
(Seneviratne et al. 2008; Delattin et al. 2014), the
increased ROS scavenging capacity might be respon-
sible for increased resistance of biofilm to ROS inducer
antifungal agents. Therefore, more investigation on ROS
scavenging capacity in filamentous fungi and targeting
the oxidative defense system could be a good strategy
to combat resistant fungal biofilms.
6.6. Drug tolerance and persister cells
Persister cells or dormant variants are a subpopulation
of general cells inside a microbial population that are
highly tolerant to drugs (Lewis 2008,2010; Ramage et
al. 2012; Rosenberg et al. 2018). These cells are an
important mechanism of tolerance particularly in
chronic infections. The key feature of a persister cell is
to tolerate the disruptive effect of antifungal agents by
reducing its metabolism and cell division (Lewis 2010).
Since antifungal agents interfere with active metabolic
processes, dormant cells, which are not metabolising
substrates and dividing, are no longer active targets for
the antifungals (Lewis 2010). Within fungal biofilms,
Box 1. An overview of the substitute therapeutic strategies against resistant biofilms.
14 M. ROUDBARY ET AL.
persister cells can resist high concentrations of antifun-
gals by a phenotypic variation that enter dormancy
(Schumacher et al. 2009; Lafleur et al. 2010). As a result,
they are most likely contribute to the resistance and
recalcitrance of biofilm infections (Wuyts et al. 2018).
Candida albicans biofilms have been shown to produce
multidrug tolerant persisters (LaFleur et al. 2006).
Relative to wild-types, persisters are not mutants but
rather phenotypic variants (LaFleur et al. 2006).
Remarkably, C. albicans persisters have only been
observed in the biofilm but not in a planktonic state
(LaFleur et al. 2006). Identification of possible genes
and mechanisms involved in persister, particularly in
filamentous fungi, may uncover a way to design an
effective anti-biofilm therapy, like the conventional anti-
fungal combination with a compound inhibiting per-
sister formation or maintenance.
7. Concepts in anti-biofilm approaches
Given the resistance of fungal biofilm-related infections
towards conventional antifungals, alternative anti-bio-
film agents and strategies are urgently needed. Since
the MICs of sessile cells are often notably higher than
the MICs of planktonic cells (Mowat et al. 2007; Seidler
et al. 2008), the higher values of agents are required to
reach the MIC in vivo, which is not possible due to the
toxicity and side effects. Consequently, the treatment
with conventional drugs that are currently used may
only decrease the biofilms, but might not eliminate the
entire biofilm, resulting in chronic and persistent infec-
tions (Wuyts et al. 2018). Therefore, to overcome the
drug resistance of biofilm communities, alternative
approaches are required (Box 1). Biofilm-associated
infections can be treated in two manners; either by
inhibition of biofilm formation or by disruption of
established biofilms (Ahmad Khan et al. 2020; Mishra et
al. 2020).
Combination therapy is a novel and relevant antibio-
film strategy. It could be based on either antifungal
drug combinations or an antifungal with a non-antifun-
gal or a potentiator to accelerate the antifungal efficacy
(Livengood et al. 2020; Tits et al. 2020). Combination of
caspofungin with amphotericin B or voriconazole
against Aspergillus biofilms demonstrated a synergistic
effect with a significant reduction of the metabolic
activity of A. fumigatus biofilm, in comparison to caspo-
fungin, amphotericin B, or voriconazole alone at the
same concentrations (Liu et al. 2012). Furthermore, the
combination of vorinostat as a potentiator with azoles,
revealed a synergistic effect against both biofilm and
planktonic cells of Aspergillus spp., including A.
fumigatus,A. flavus,A. terreus (Tu et al. 2020). Besides,
some enzymes also showed potential roles in combin-
ation with conventional antifungal agents (Bugli et al.
2013; Lohse et al. 2020). For instance, A. fumigatus bio-
films exposed to alginate lyase, an enzyme degrading
the EPSs of biofilm, combined with amphotericin B,
indicated more susceptibility, proposing EPS-degrading
enzymes as a promising combination therapy to
improve the management of biofilm-associated infec-
tions (Bugli et al. 2013).
Apart from conventional antifungal agents, many
novel antifungals have already been described (Vahedi-
Shahandashti and Lass-Fl
orl 2020), which could be
promising either alone or in combination. The antibio-
film activity of olorofim and voriconazole against
Lomentospora prolificans showed promising efficacy,
while amphotericin B and micafungin did not exhibit a
desirable antibiofilm activity (Kirchhoff et al. 2020),
which emphasises the importance of more investigation
on antibiofilm activity of novel agents.
Many studies have been focused on inhibiting fungal
adhesion to surfaces through a range of surface coating
manners, including coating with metal nanoparticles,
antimicrobials, and lock solutions (Cateau et al. 2008;
Redding et al. 2009; Barad et al. 2017; Paulone et al.
2017; Lagree et al. 2018; Naderi et al. 2019; Kitisin et al.
2020; Vera-Gonz
alez and Shukla 2020; Vladkova et al.
2020). For instance, in Candida spp., nanoparticles such
as silver, zinc oxide, and titanium dioxide, graphen
oxide (Asadi Shahi et al. 2019), and gold nanoparticles
(Rahimi et al. 2019) showed potent activity against fun-
gal biofilm by different mechanisms. These included
ROS production, lipid peroxidation, cell wall disruption,
and decreased gene expression profiles mainly involved
in biofilm formation (Haghighi et al. 2013; Lara et al.
2015; Barad et al. 2017; Nikoomanesh et al. 2019).
Additionally, synthesis, or function inhibition of GAG, as
adhesive features, in A. fumigatus, has been introduced
as a potential therapeutic target (Lee et al. 2016). This
adhesion activity demands deacetylation by the carbo-
hydrate esterase type 4 Agd3, which has a crucial role
in virulence and extracellular localisation, proposing
Agd3 to be a promising therapeutic target (Lee et
al. 2016).
Some research has highlighted the anti-biofilm effi-
cacy of natural products (Khan and Ahmad 2012; Sardi
et al. 2013; Roemer and Krysan 2014) and metabolites
(Klausmeyer et al. 2005; Estrela and Abraham 2016),
that could lead to the development of unique alterna-
tive therapeutics. Several structural parts of the biofilm
can be potential targets for the development of
enzyme-based therapies, such as matrix and eDNA
CRITICAL REVIEWS IN MICROBIOLOGY 15
(Martins et al. 2012; Rajendran et al. 2013). This is the
case of glycoside hydrolases Sph3 and Ega3, enzymes
involved in the GAG biosynthetic pathway, that have
been repurposed as anti-GAG therapeutics which can
disrupt biofilm (Bamford et al. 2019; Le Mauff et
al. 2019).
Furthermore, interrupting signalling pathways that
are involved in biofilm formation is also a compelling
strategy, such as QS signalling (Pierce and Lopez-Ribot
2013;Mishraetal.2020). Quorum-quenching molecules
(QQMs) or QS inhibitors (Sharma et al. 2015), are an
interesting alternative therapy. They typically interfere in
the process of QS, using three main tactics: (1) prevent-
ing QSMs production, (2) degrading QSMs, and (3) block-
ing the QSMs receptors. Currently, some QQMs have
been applied against bacterial and fungal (e.g. Candida
spp.), biofilm; for instance, thiazolidinedione-8, QQ-5,
and QQ-7 (Feldman et al. 2014; Weiland-Br
auer et al.
2019). Further screening, either on new QQMs in natural
products libraries or QSMs secreted by other microor-
ganisms in polymicrobial communities (such as filament-
ous fungi), may play a QQ role against other species and
may provide more therapeutic options.
Additionally, probiotics with different features have
revealed a distinct horizon to combat biofilms. They pos-
sess less cytotoxicity than QS-inhibitors and less induc-
tion of selective pressure on resistant isolates than
conventional agents (Sadiq et al. 2019; Barzegari et al.
2020). Probiotics, particularly Lactobacillus spp., produce
different anti-biofilm metabolites, which have been
widely studied in bacteria, such as organic acids, perox-
ides, fatty acids, bacteriocins, and biosurfactants
(Shokouhfard et al. 2015; Vahedi Shahandashti et al.
2016; Barzegari et al. 2020). Some studies have shown
the beneficial effects of probiotics, either cells or cell-free
supernatants, on fungal biofilm, mostly C. albicans (Vilela
et al. 2015; Matsubara et al. 2016;Tanetal.2017;
Cern
akov
aetal.2019) and bacterial-fungal polymicrobial
biofilms (Hager et al. 2019). However, the influence of
probiotics on filamentous fungal biofilm remains
unknown, which implies opening new research paths to
provide more treatment or/and prevention strategies.
Another plausible alternative therapy is photo-
dynamic therapy (PDT), which has been applied in bac-
teria and Candida spp.. Antimicrobial PDT uses different
photosensitizers or non-toxic dyes, various light sour-
ces, and wavelengths, which produce ROS and attack
proteins, lipids, and nucleic acids present within the
biofilm matrix (Plaetzer et al. 2009; Hu et al. 2018). In
Candida spp. biofilms, antimicrobial PDT showed prom-
ising results, featuring the potential use of this method
in other fungal biofilm eradication (Lopes et al. 2014).
PDT might also be suitable to be adopted for filament-
ous fungal biofilms eradication (Biel and Gomer 2010;
Hu et al. 2018; Sharma et al. 2019).
Finally, the CRISPR-based tool is one of the most
powerful methods in underlying genetic interactions
determination (Shapiro et al. 2018; Sharma et al. 2019).
In C. albicans, single or double deletions of 12 adhesin
genes by the CRISPR system, led to a combination
adhesin mutant with 144 single- and double-knockout
(Shapiro et al. 2018). A combined adhesin mutant dem-
onstrated a set of adhesins that can be targeted in bio-
film impairment, highlighting the contribution of
CRISPR in the feasibility of generating a combination of
mutations. The potential applications of CRISPR-based
methods look promising in uncovering the distinctive
pathways and genes involved in filamentous fun-
gal biofilm.
8. Perspective
The biofilm formation of saprophytic and pathogenic
fungi is a worldwide health concern. Invasive fungal
infections directly linked to biofilms are associated with
severe clinical conditions, and their treatment remains
incredibly challenging due to the high resistance to the
available antifungals. Despite recent progress in the
understanding of biofilms formed by bacteria and
yeasts, a comprehensive understanding of the molecu-
lar mechanisms and crucial components involved in
establishing filamentous fungal biofilm infections is still
required. Because of the distinctive biofilm growth and
planktonic growth, developing biofilm-based treatment
alternatives to manage or eradicate these microbial
communities seems an urgent necessity. Moreover,
future studies are needed to understand the scale of
biofilm formation and its clinical impact by patho-
genic fungi.
Acknowledgments
C.F.R. would like to acknowledge the UID/EQU/00511/2020
ProjectLaboratory of Process Engineering, Environment,
Biotechnology and Energy (LEPABE), financed by national
funds through FCT/MCTES (PIDDAC). The authors warmly
thank Azin Kheirkhah for the scientific linguistic revision.
Author contributions
Maryam Roudbary contributed to data curation, the review
design and writing. Roya Vahedi-Shahandashti contributed to
review design, writing, Figure preparation and editing. Andr
e
Luis Souza dos Santos contributed to interpretation/selection
and editing. Shahla Roudbar Mohammadi and Peyman Aslani
contributed to review draft and revision. Cornelia Lass-Fl
orl
16 M. ROUDBARY ET AL.
and C
elia F. Rodrigues contributed to the review design,
editing and supervision. They are all accountable for their
own contributions and ensure the accuracy of any part of
the work.
Disclosure statement
The authors declare no conflict of interest.
ORCID
C
elia F. Rodrigues http://orcid.org/0000-0001-8633-2230
References
Ader F, Bienvenu A-L, Rammaert B, Nseir S. 2009.
Management of invasive aspergillosis in patients with
COPD: rational use of voriconazole. Int J Chron Obstruct
Pulmon Dis. 4:279287.
Affeldt KJ, Brodhagen M, Keller NP. 2012. Aspergillus oxylipin
signaling and quorum sensing pathways depend on g
protein-coupled receptors. Toxins. 4(9):695717.
Ahmad Khan MS, Alshehrei F, Al-Ghamdi SB, Bamaga MA, Al-
Thubiani AS, Alam MZ. 2020. Virulence and biofilms as
promising targets in developing antipathogenic drugs
against candidiasis. Future Sci Oa. 6(2):FSO440.
Aimanianda V, Bayry J, Bozza S, Kniemeyer O, Perruccio K,
Elluru SR, Clavaud C, Paris S, Brakhage AA, Kaveri SV, et al.
2009. Surface hydrophobin prevents immune recognition
of airborne fungal spores. Nature. 460(7259):11171121.
Al Abdallah Q, Choe S-I, Campoli P, Baptista S, Gravelat FN,
Lee MJ, Sheppard DC. 2012. A conserved C-Terminal
domain of the Aspergillus fumigatus developmental regula-
tor meda is required for nuclear localization, adhesion and
virulence. PLoS One. 7(11):e49959.
Almeida CA, Azevedo MMB, Chaves FCM, Roseo de Oliveira
M, Rodrigues IA, Bizzo HR, Gama PE, Alviano DS, Alviano
CS. 2018. Piper essential oils inhibit Rhizopus oryzae
growth, biofilm formation, and rhizopuspepsin activity.
Can J Infect Dis Med Microbiol. 2018:5295619.
Amin R, Dupuis A, Aaron SD, Ratjen F. 2010. The effect of
chronic infection with Aspergillus fumigatus on lung func-
tion and hospitalization in patients with cystic fibrosis.
Chest. 137(1):171176.
Arastehfar A, Gabald
on T, Garcia-Rubio R, Jenks JD, Hoenigl
M, Salzer HJF, Ilkit M, Lass-Fl
orl C, Perlin DS. 2020. Drug-
resistant fungi: an emerging challenge threatening our
limited antifungal armamentarium. Antibiotics. 9(12):877.
Arastehfar A, Lass-Fl
orl C, Garcia-Rubio R, Daneshnia F, Ilkit
M, Boekhout T, Gabaldon T, Perlin DS. 2020. The quiet and
underappreciated rise of drug-resistant invasive fungal
pathogens. JoF. 6(3):138.
Asadi Shahi S, Roudbar Mohammadi S, Roudbary M, Delavari
H. 2019. Formulation of graphene oxide/fluconazole com-
pound as a promising agent against Candida albicans.
Prog Biomater. 8(1):4350.
Baillie GS, Douglas LJ. 1998. Iron-Limited biofilms of Candida
albicans and their susceptibility to Amphotericin B.
Antimicrob Agents Chemother. 42(8):21462149.
Bamford NC, Le Mauff F, Subramanian AS, Yip P, Mill
an C,
Zhang Y, Zacharias C, Forman A, Nitz M, Cod
ee JDC, et al.
2019. Ega3 from the fungal pathogen Aspergillus fumigatus
is an endo-a-1,4-galactosaminidase that disrupts microbial
biofilms. J Biol Chem. 294(37):1383313849.
Bamford NC, Snarr BD, Gravelat FN, Little DJ, Lee MJ,
Zacharias CA, Chabot JC, Geller AM, Baptista SD, Baker P,
et al. 2015. Sph3 is a glycoside hydrolase required for the
biosynthesis of galactosaminogalactan in Aspergillus
fumigatus. J Biol Chem. 290(46):2743827450.
Bandara HMHN, Wood DLA, Vanwonterghem I, Hugenholtz P,
Cheung BPK, Samaranayake LP. 2020. Fluconazole resist-
ance in Candida albicans is induced by Pseudomonas aeru-
ginosa quorum sensing. Sci Rep. 10(1):7769.
Barad S, Roudbary M, Omran AN, Daryasari MP. 2017.
Preparation and characterization of ZnO nanoparticles
coated by chitosan-linoleic acid; fungal growth and biofilm
assay. Bratisl Lek Listy. 118(3):169174.
Barriuso J, Hogan DA, Keshavarz T, Mart
ınez MJ. 2018. Role
of quorum sensing and chemical communication in fungal
biotechnology and pathogenesis. FEMS Microbiol Rev.
42(5):627638.
Barzegari A, Kheyrolahzadeh K, Hosseiniyan Khatibi SM,
Sharifi S, Memar MY, Zununi Vahed S. 2020. Zununi
Vahed, S. The battle of probiotics and their derivatives
against biofilms. Infect Drug Resist. 13:659672.
Beauvais A, Latg
e J-P. 2015. Aspergillus biofilm in vitro and in
vivo. Microbiol Spectr. 3(4). DOI:10.1128/microbiolspec.MB-
0017-2015.
Beauvais A, Schmidt C, Guadagnini S, Roux P, Perret E, Henry
C, Paris S, Mallet A, Pr
evost M-C, Latg
e JP. 2007. An extra-
cellular matrix glues together the aerial-grown hyphae of
Aspergillus fumigatus. Cell Microbiol. 9(6):15881600.
Beer D, de Stoodley P, Lewandowski Z. 1994. Liquid flow in
heterogeneous biofilms. Biotechnol Bioeng. 44(5):636641.
Biel MA, 2010. Photodynamic therapy of bacterial and fungal
biofilm infections. In: Gomer CJ. editor. Photodynamic
therapy: methods and protocols, Totowa (NJ): Humana
Press. p. 175194. https://doi.org/10.1007/978-1-60761-
697-9_13.
Biesebeke R, Ruijter G, Rahardjo YSP, Hoogschagen MJ,
Heerikhuisen M, Levin A, Driel KGA, Schutyser MAI,
Dijksterhuis J, Zhu Y, et al. 2002. Aspergillus oryzae in
solid-state and submerged fermentations. progress report
on a multi-disciplinary project. FEMS Yeast Res. 2(2):
245248.
Bila NM, Costa-Orlandi CB, Vaso CO, Bonatti JLC, de Assis LR,
Regasini LO, Fontana CR, Fusco-Almeida AM, Mendes-
Giannini MJS. 2021. 2-hydroxychalcone as a potent com-
pound and photosensitizer against dermatophyte biofilms.
Front Cell Infect Microbiol. 11:679470.
Bom VLP, de Castro PA, Winkelstr
oter LK, Marine M, Hori JI,
Ramalho LNZ, dos Reis TF, Goldman MHS, Brown NA,
Rajendran R, et al. 2015. The Aspergillus fumigatus SitA
phosphatase homologue is important for adhesion, cell
wall integrity, biofilm formation, and virulence. Eukaryot
Cell. 14(8):728744.
Brand~
ao I, de SL, Oliveira-Moraes HM, da S. Souza Motta CM,
de Oliveira NT, de, Magalh~
aes OMC. 2018. Elastin increases
biofilm and extracellular matrix production of Aspergillus
fumigatus. Braz J Microbiol. 49(3):675682.
CRITICAL REVIEWS IN MICROBIOLOGY 17
Bridier A, Briandet R, Thomas V, Dubois-Brissonnet F. 2011.
Resistance of bacterial biofilms to disinfectants: a review.
Biofouling. 27(9):10171032.
Brilhante RSN, Aguiar L, de Sales JA, Ara
ujo G. d S, Pereira
VS, Pereira-Neto W, de A, Pinheiro A, de Q, Paix~
ao GC,
Cordeiro R, de A, Sidrim JJC, et al. 2019. Ex vivo biofilm-
forming ability of dermatophytes using dog and cat hair:
an ethically viable approach for an infection model.
Biofouling. 35(4):392400.
Brilhante RSN, de Lima RAC, Marques FJ, de F, Silva NF,
Caetano
EP, Castelo-Branco D, de SCM, Bandeira T, de JPG,
Moreira JLB, Cordeiro R, de A, Monteiro AJ, et al. 2015.
Histoplasma capsulatum. in planktonic and biofilm forms:
in vitro susceptibility to amphotericin b, itraconazole and
farnesol. J Med Microbiol. 64(Pt 4):394399.
Brodhun F, Feussner I. 2011. Feussner, I. oxylipins in fungi.
Febs J. 278(7):10471063.
Brown GD, Denning DW, Gow NAR, Levitz SM, Netea MG,
White TC. 2012. Hidden killers: human fungal infections.
Sci Transl Med. 4(165):165rv13.
Brown SH, Scott JB, Bhaheetharan J, Sharpee WC, Milde L,
Wilson RA, Keller NP. 2009. Oxygenase coordination is
required for morphological transition and the host-fungus
interaction of Aspergillus flavus. Mol Plant Microbe Interact.
22(7):882894.
Bruns S, Kniemeyer O, Hasenberg M, Aimanianda V,
Nietzsche S, Thywissen A, Jeron A, Latg
e J-P, Brakhage AA,
Gunzer M. 2010. Production of extracellular traps against
Aspergillus fumigatus in vitro and in infected lung tissue is
dependent on invading neutrophils and influenced by
hydrophobin RodA. PLoS Pathog. 6(4):e1000873.
Bruns S, Seidler M, Albrecht D, Salvenmoser S, Remme N,
Hertweck C, Brakhage AA, Kniemeyer O, M
uller F-MC.
2010. Functional genomic profiling of Aspergillus fumigatus
biofilm reveals enhanced production of the mycotoxin
gliotoxin. Proteomics. 10(17):30973107.
Bueid A, Howard SJ, Moore CB, Richardson MD, Harrison E,
Bowyer P, Denning DW. 2010. Azole Antifungal resistance
in Aspergillus fumigatus: 2008 and 2009. J Antimicrob
Chemother. 65(10):21162118.
Bugli F, Posteraro B, Papi M, Torelli R, Maiorana A, Paroni
Sterbini F, Posteraro P, Sanguinetti M, De Spirito M. 2013.
In Vitro interaction between alginate lyase and amphoteri-
cin B against Aspergillus fumigatus biofilm determined by
different methods. Antimicrob Agents Chemother. 57(3):
12751282.
Burkhart CN, Burkhart CG, Gupta AK. 2002. Dermatophytoma:
recalcitrance to treatment because of existence of fungal
biofilm. J Am Acad Dermatol. 47(4):629631.
Carreto-Binaghi LE, Damasceno LS, Pitangui N, de S, Fusco-
Almeida AM, Mendes-Giannini MJS, Zancop
e-Oliveira RM,
Taylor ML. 2015. Could Histoplasma capsulatum be related
to healthcare-associated infections? Biomed Res Int. 2015:
982429982411.
Cateau E, Rodier M-H, Imbert C. 2008. In vitro efficacies of
caspofungin or micafungin catheter lock solutions on
Candida albicans biofilm growth. J Antimicrob Chemother.
62(1):153155.
Cattana ME, Tracogna MF, Marques I, Rojas F, Fern
andez M,
Sosa M, de los
A, Mussin J, Giusiano G. 2017. In vivo
Paracoccidiodes Sp. Biofilm on vascular prosthesis.
Mycopathologia. 182(7-8):747749.
Cern
akov
a L, Light C, Salehi B, Rogel-Castillo C, Victoriano M,
Martorell M, Sharifi-Rad J, Martins N, Rodrigues CF. 2019.
Novel therapies for biofilm-based Candida spp. infections.
Adv Exp Med Biol 1214:93123.
Chakrabarti A, Das A, Mandal J, Shivaprakash MR, George VK,
Tarai B, Rao P, Panda N, Verma SC, Sakhuja V. 2006. The
rising trend of invasive zygomycosis in patients with
uncontrolled diabetes mellitus. Med Mycol. 44(4):335342.
Chandrasekar PH, Manavathu EK. 2008. Do Aspergillus spe-
cies produce biofilm? Future Microbiol. 3(1):1921.
Chatterjee S, Das S. 2020. Developmental stages of biofilm
and characterization of extracellular matrix of manglico-
lous fungus Aspergillus niger BSC-1. J Basic Microbiol.
60(3):231242.
Chen RE, Thorner J. 2007. Function and regulation in MAPK
signaling pathways: lessons learned from the yeast
Saccharomyces cerevisiae. Biochim Biophys Acta. 1773(8):
13111340.
Chen Y, Mauff FL, Wang Y, Lu R, Sheppard DC, Lu L, Zhang
S. 2020. The transcription factor SomA synchronously reg-
ulates biofilm formation and cell wall homeostasis in
Aspergillus fumigatus. mBio. 11(6). DOI:10.1128/mBio.
02329-20.
Chowdhary A, Perfect J, de Hoog GS. 2014. Black molds and
melanized yeasts pathogenic to humans. Cold Spring Harb
Perspect Med. 5(8):a019570.
C
ordova-Alc
antara IM, Venegas-Cort
es DL, Mart
ınez-Rivera
M
A, P
erez NO, Rodriguez-Tovar AV. 2019. Biofilm charac-
terization of Fusarium solani keratitis isolate: increased
resistance to antifungals and UV light. J Microbiol. 57(6):
485497.
Cortez KJ, Roilides E, Quiroz-Telles F, Meletiadis J,
Antachopoulos C, Knudsen T, Buchanan W, Milanovich J,
Sutton DA, Fothergill A, et al. 2008. Infections Caused by
Scedosporium Spp. Clin Microbiol Rev. 21(1):157197.
Costa-Orlandi C, Sardi J, Pitangui N, de Oliveira H, Scorzoni L,
Galeane M, Medina-Alarc
on K, Melo W, Marcelino M, Braz
J, et al. 2017. Fungal Biofilms and Polymicrobial Diseases. J
Fungi (Basel). 3(2):22.
Costa-Orlandi CB, Sardi JCO, Santos CT, Fusco-Almeida AM,
Mendes-Giannini MJS. 2014. In vitro characterization of
Trichophyton rubrum and Trichophyton mentagrophytes
Biofilms. Biofouling. 30(6):719727.
Costa-Orlandi CB, Serafim-Pinto A, da Silva PB, Bila NM,
Bonatti JL, de C, Scorzoni L, Singulani J, de L, Santos CT,
dos Nazar
e AC, Chorilli M, et al. 2020. Incorporation of
Nonyl 3,4-dihydroxybenzoate into nanostructured lipid
systems: effective alternative for maintaining anti-dermato-
phytic and antibiofilm activities and reducing toxicity at
high concentrations. Front Microbiol. 11:1154.
Costerton JW, Stewart PS, Greenberg EP. 1999. Bacterial bio-
films: a common cause of persistent infections. Science.
284(5418):13181322.
Danielli LJ, Lopes W, Vainstein MH, Fuentefria AM, Apel MA.
2017. Biofilm formation by Microsporum canis. Clin
Microbiol Infect. 23(12):941942.
Davies D. 2003. Understanding biofilm resistance to antibac-
terial agents. Nat Rev Drug Discov. 2(2):114122.
Davis LE, Cook G, Costerton J. 2002. Biofilm on ventriculo-
peritoneal shunt tubing as a cause of treatment failure in
Coccidioidal meningitis. Emerg Infect Dis. 8(4):376379.
18 M. ROUDBARY ET AL.
de Mello TP, Aor AC, Branquinha MH, Dos Santos ALS. 2020.
Insights into the interaction of Scedosporium apiospermum,
Scedosporium aurantiacum, Scedosporium minutisporum,
and Lomentospora prolificans with lung epithelial cells.
Braz J Microbiol. 51(2):427436.
Delattin N, Cammue BPA, Thevissen K. 2014. Reactive oxygen
species-inducing antifungal agents and their activity
against fungal biofilms. Future Med Chem. 6(1):7790.
Desai JV, Mitchell AP, Andes DR. 2014. Fungal biofilms, drug
resistance, and recurrent infection. Cold Spring Harbor
Perspect Med. 4(10):a019729a019729.
Deveau A, Piispanen AE, Jackson AA, Hogan DA. 2010.
Farnesol induces hydrogen peroxide resistance in Candida
albicans yeast by inhibiting the ras-cyclic AMP signaling
pathway. Eukaryot Cell. 9(4):569577.
Dichtl K, Ebel F, Dirr F, Routier FH, Heesemann J, Wagener J.
2010. Farnesol misplaces tip-localized rho proteins and
inhibits cell wall integrity signalling in Aspergillus fumiga-
tus farnesol impairs cell wall integrity in A. fumigatus. Mol
Microbiol. 76(5):11911204.
Dixon EF, Hall RA. 2015. Noisy neighbourhoods: quorum
sensing in fungal-polymicrobial infections. Cell Microbiol.
17(10):14311441.
Donlan RM. 2002. Biofilms: microbial life on surfaces. Emerg
Infect Dis. 8(9):881890.
Drew SW, Demain AL. 1977. Effect of primary metabolites on
secondary metabolism. Annu Rev Microbiol. 31:343356.
Dumitru R, Hornby JM, Nickerson KW. 2004. Defined anaer-
obic growth medium for studying Candida albicans basic
biology and resistance to eight antifungal drugs.
Antimicrob Agents Chemother. 48(7):23502354.
Edwards PA, Ericsson J. 1999. Sterols and isoprenoids: signal-
ing molecules derived from the cholesterol biosynthetic
pathway. Annu Rev Biochem. 68(1):157185.
Escande W, Fayad G, Modine T, Verbrugge E, Koussa M,
Senneville E, Leroy O. 2011. Culture of a prosthetic valve
excised for streptococcal endocarditis positive for
Aspergillus fumigatus 20 years after previous a fumigatus
endocarditis. Ann Thorac Surg. 91(6):e92e93.
Estrela A, Abraham W-R. 2016. Fungal metabolites for the
control of biofilm infections. Agriculture. 6(3):37.
Ezzat WF, Askoura AM, Elshazly AN, Falah HY. 2020.
Pathological role of biofilm in fungal rhinosinusitis. QJM.
1(Supplement):113.
Fan Z, Li Z, Xu Z, Li H, Li L, Ning C, Ma L, Xie X, Wang G, Yu
H. 2015. CspA influences biofilm formation and drug resist-
ance in pathogenic fungus Aspergillus fumigatus. Biomed
Res Int. 2015:960357960359.
Fanning S, Mitchell AP. 2012. Fungal biofilms. PLoS Pathog.
8(4):e1002585.
Feldman M, Al-Quntar A, Polacheck I, Friedman M, Steinberg
D. 2014. Therapeutic potential of thiazolidinedione-8 as an
antibiofilm agent against Candida albicans. PLoS One. 9(5):
e93225.
Ferreira ME, da S, Malavazi I, Savoldi M, Brakhage AA,
Goldman MHS, Kim HS, Nierman WC, Goldman GH. 2006.
Transcriptome analysis of Aspergillus fumigatus exposed to
voriconazole. Curr Genet. 50(1):3244.
Filler SG, Sheppard DC. 2006. Fungal invasion of normally
non-phagocytic host cells. PLoS Pathog. 2(12):e129.
Fiori B, Posteraro B, Torelli R, Tumbarello M, Perlin DS, Fadda
G, Sanguinetti M. 2011. In vitro activities of anidulafungin
and other antifungal agents against biofilms formed by
clinical isolates of different Candida and Aspergillus spe-
cies. Antimicrob Agents Chemother. 55(6):30313035.
Fischer MS, Glass NL. 2019. Communicate and fuse: how fila-
mentous fungi establish and maintain an interconnected
mycelial network. Front Microbiol. 10:619.
Fontaine T, Beauvais A, Loussert C, Thevenard B, Fulgsang
CC, Ohno N, Clavaud C, Prevost M-C, Latg
e J-P. 2010. Cell
wall alpha1-3glucans induce the aggregation of germinat-
ing conidia of Aspergillus fumigatus. Fungal Genet Biol.
47(8):707712.
Fontaine T, Delangle A, Simenel C, Coddeville B, van Vliet SJ,
van Kooyk Y, Bozza S, Moretti S, Schwarz F, Trichot C, et
al. 2011. Galactosaminogalactan, a new immunosuppres-
sive polysaccharide of Aspergillus fumigatus. PLoS Pathog.
7(11):e1002372.
Gao L, Jiang S, Sun Y, Deng M, Wu Q, Li M, Zeng T. 2016.
Evaluation of the effects of photodynamic therapy alone
and combined with standard antifungal therapy on plank-
tonic cells and biofilms of Fusarium Spp. and Exophiala
Spp. Front Microbiol. 7:617.
Gao L, Sun Y, He C, Zeng T, Li M. 2017. Synergistic effects of
tacrolimus and azoles against Exophiala dermatitis.
Antimicrob Agents Chemother. 61(12):e00948-17.
Garcia LM, Costa-Orlandi CB, Bila NM, Vaso CO, Gonc¸alves
LNC, Fusco-Almeida AM, Mendes-Giannini MJS. 2020. A
two-way road: antagonistic interaction between dual-spe-
cies biofilms formed by Candida albicans/Candida parapsi-
losis and Trichophyton rubrum. Front Microbiol. 11:1980.
Gibbons JG, Beauvais A, Beau R, McGary KL, Latg
e J-P, Rokas
A. 2012. Global transcriptome changes underlying colony
growth in the opportunistic human pathogen Aspergillus
fumigatus. Eukaryot Cell. 11(1):6878.
Gintjee TJ, Donnelley MA, Thompson GR. 2020. Aspiring anti-
fungals: review of current antifungal pipeline develop-
ments. JoF. 6(1):28.
Gonc¸alves LNC, Costa-Orlandi CB, Bila NM, Vaso CO, Da Silva
RAM, Mendes-Giannini MJS, Taylor ML, Fusco-Almeida AM.
2020. Biofilm formation by Histoplasma capsulatum in dif-
ferent culture media and oxygen atmospheres. Front
Microbiol. 11:1455.
Gravelat FN, Beauvais A, Liu H, Lee MJ, Snarr BD, Chen D, Xu
W, Kravtsov I, Hoareau CMQ, Vanier G, et al. 2013.
Aspergillus galactosaminogalactan mediates adherence to
host constituents and conceals hyphal b-glucan from the
immune system. PLoS Pathog. 9(8):e1003575.
Gravelat FN, Ejzykowicz DE, Chiang LY, Chabot JC, Urb M,
Macdonald KD, Al-Bader N, Filler SG, Sheppard DC. 2010.
Aspergillus fumigatus MedA governs adherence, host cell
interactions and virulence. Cell Microbiol. 12(4):473488.
Gulis V, Suberkropp K, Rosemond AD. 2008. Comparison of
fungal activities on wood and leaf litter in unaltered and
nutrient-enriched headwater streams. Appl Environ
Microbiol. 74(4):10941101.
Hager CL, Isham N, Schrom KP, Chandra J, McCormick T,
Miyagi M, Ghannoum MA. 2019. Effects of a novel pro-
biotic combination on pathogenic bacterial-fungal polymi-
crobial biofilms. mBio. 10(2):e00338e00319.
Haghighi F, Mohammadi S, Mohammadi P, Hosseinkhani S,
Shidpour R. 2013. Antifungal activity of TiO 2 nanopar-
ticles and EDTA on Candida albicans biofilms. 1:3338.
CRITICAL REVIEWS IN MICROBIOLOGY 19
Harding MW, Marques LLR, Howard RJ, Olson ME. 2009. Can
filamentous fungi form biofilms? Trends Microbiol. 17(11):
475480.
Harris SD. 2019. Hyphal branching in filamentous fungi. Dev
Biol. 451(1):3539.
Heinrichs G, H
ubner I, Schmidt CK, de Hoog GS, Haase G.
2013a. Analysis of black fungal biofilms occurring at
domestic water taps. i: compositional analysis using tag-
encoded FLX amplicon pyrosequencing. Mycopathologia.
175(5-6):387397.
Heinrichs G, H
ubner I, Schmidt CK, de Hoog GS, Haase G.
2013b. Analysis of black fungal biofilms occurring at
domestic water taps (II): potential routes of entry.
Mycopathologia. 175(5-6):399412.
Hilmarsdottir I, Haraldsson H, Sigurdardottir A, Sigurgeirsson
B. 2005. Dermatophytes in a swimming pool facility: differ-
ence in dermatophyte load in mens and womens dress-
ing rooms. Acta Derm Venereol. 85(3):267268.
Hogan DA. 2006. Talking to themselves: autoregulation and
quorum sensing in fungi. Eukaryot Cell. 5(4):613619.
Hornby JM, Jensen EC, Lisec AD, Tasto JJ, Jahnke B,
Shoemaker R, Dussault P, Nickerson KW. 2001. Quorum
sensing in the dimorphic fungus Candida albicans is medi-
ated by farnesol. Appl Environ Microbiol. 67(7):29822992.
Horowitz Brown S, Zarnowski R, Sharpee WC, Keller NP. 2008.
Morphological transitions governed by density depend-
ence and lipoxygenase activity in Aspergillus fumigatus.
Appl Environ Microbiol. 74(18):56745685.
Hu X, Huang Y-Y, Wang Y, Wang X, Hamblin MR. 2018.
Antimicrobial photodynamic therapy to control clinically
relevant biofilm infections. Front Microbiol. 9:01299.
Imamura Y, Chandra J, Mukherjee PK, Lattif AA, Szczotka-
Flynn LB, Pearlman E, Lass JH, ODonnell K, Ghannoum
MA. 2008. Fusarium and Candida albicans biofilms on soft
contact lenses: model development, influence of lens
type, and susceptibility to lens care solutions. Antimicrob
Agents Chemother. 52(1):171182.
Jeloka TK, Shrividya S, Wagholikar G. 2011. Catheter outflow
obstruction due to an aspergilloma. Perit Dial Int. 31(2):
211212.
Joo JH, Jetten AM. 2008. NF-kappaB-dependent transcrip-
tional activation in lung carcinoma cells by farnesol
involves p65/RelA(Ser276) phosphorylation via the MEK-
MSK1 signaling pathway. J Biol Chem. 283(24):
1639116399.
Kalia VC, editor. 2015. Quorum sensing vs quorum quench-
ing: a battle with no end in sight. India: Springer.
Khan MSA, Ahmad I. 2012. Antibiofilm activity of certain phy-
tocompounds and their synergy with fluconazole against
Candida albicans biofilms. J Antimicrob Chemother. 67(3):
618621.
Kirchhoff L, Dittmer S, Weisner A-K, Buer J, Rath P-M,
Steinmann J. 2020. Antibiofilm activity of antifungal drugs,
including the novel drug olorofim, against Lomentospora
prolificans. J Antimicrob Chemother. 75(8):21332140.
Kirchhoff L, Olsowski M, Zilmans K, Dittmer S, Haase G,
Sedlacek L, Steinmann E, Buer J, Rath P-M, Steinmann J.
2017. Biofilm formation of the black yeast-like fungus
Exophiala dermatitis and its susceptibility to antiinfective
agents. Sci Rep. 7(1):42886.
Kirchhoff L, Weisner A-K, Schrepffer M, Hain A, Scharmann U,
Buer J, Rath P-M, Steinmann J. 2020. Phenotypical
characteristics of the black yeast Exophiala dermatitis are
affected by Pseudomonas aeruginosa in an artificial spu-
tum medium mimicking cystic fibrosis-like conditions.
Front Microbiol. 11:471.
Kischkel B, Souza GK, Chiavelli LUR, Pomini AM, Svidzinski
TIE, Negri M. 2020. The ability of farnesol to prevent adhe-
sion and disrupt Fusarium keratoplasticum biofilm. Appl
Microbiol Biotechnol. 104(1):377389.
Kitisin T, Muangkaew W, Ampawong S, Sukphopetch P. 2020.
Tryptophol coating reduces catheter-related cerebral and
pulmonary infections by Scedosporium apiospermum. Infect
Drug Resist. 13:24952508.
Klausmeyer P, McCloud TG, Tucker KD, Cardellina JH,
Shoemaker RH. 2005. Aspirochlorine class compounds
from Aspergillus flavus inhibit azole-resistant Candida albi-
cans. J Nat Prod. 68(8):13001302.
Kowalski CH, Kerkaert JD, Liu K-W, Bond MC, Hartmann R,
Nadell CD, Stajich JE, Cramer RA. 2019. Fungal biofilm
morphology impacts hypoxia fitness and disease progres-
sion. Nat Microbiol. 4(12):24302441.
Kowalski CH, Morelli KA, Schultz D, Nadell CD, Cramer RA.
2020. Fungal biofilm architecture produces hypoxic micro-
environments that drive antifungal resistance. Proc Natl
Acad Sci USA. 117(36):2247322483.
Kuchar
ıkov
a S, Tournu H, Lagrou K, Van Dijck P, Bujd
akov
aH.
2011. Detailed comparison of Candida albicans and
Candida glabrata biofilms under different conditions and
their susceptibility to caspofungin and anidulafungin. J
Med Microbiol. 60(Pt 9):12611269.
LaFleur MD, Kumamoto CA, Lewis K. 2006. Candida albicans
biofilms produce antifungal-tolerant persister cells.
Antimicrob Agents Chemother. 50(11):38393846.
Lafleur MD, Qi Q, Lewis K. 2010. Patients with long-term oral
carriage harbor high-persister mutants of Candida albicans.
Antimicrob Agents Chemother. 54(1):3944.
Lagree K, Mon HH, Mitchell AP, Ducker WA. 2018. Impact of
surface topography on biofilm formation by Candida albi-
cans. PLoS One. 13(6):e0197925.
Lamarre C, Beau R, Balloy V, Fontaine T, Hoi JWS, Guadagnini
S, Berkova N, Chignard M, Beauvais A, Latg
e J-P. 2009.
Galactofuranose attenuates cellular adhesion of Aspergillus
fumigatus. Cell Microbiol. 11(11):16121623.
Lara HH, Romero-Urbina DG, Pierce C, Lopez-Ribot JL,
Arellano-Jim
enez MJ, Jose-Yacaman M. 2015. Effect of sil-
ver nanoparticles on Candida albicans biofilms: an ultra-
structural study. J Nanobiotechnology. 13(1):91.
Lass-Fl
orl C, Speth C, Kofler G, Dierch MP, Gunsilius E,
W
urzner R. 2003. Effect of increasing inoculum sizes of
Aspergillus hyphae on mics and mfcs of antifungal agents
by broth microdilution method. Int J Antimicrob Agents .
21(3):229233.
Lawrence JR, Korber DR, Hoyle BD, Costerton JW, Caldwell
DE. 1991. Optical sectioning of microbial biofilms. J
Bacteriol. 173(20):65586567.
Le Mauff F, Bamford NC, Alnabelseya N, Zhang Y, Baker P,
Robinson H, Cod
ee JDC, Howell PL, Sheppard DC. 2019.
Molecular mechanism of Aspergillus fumigatus biofilm dis-
ruption by fungal and bacterial glycoside hydrolases. J
Biol Chem. 294(28):1076010772.
Le Mauff F. 2020. Exopolysaccharides and biofilms. In: Latg
e,
J.-P. editor. The fungal cell wall. vol. 425, Cham: Current
20 M. ROUDBARY ET AL.
Topics in Microbiology and Immunology; Springer
International Publishing. p. 225254.
Lee MJ, Geller AM, Bamford NC, Liu H, Gravelat FN, Snarr BD,
Le Mauff F, Chabot J, Ralph B, Ostapska H, et al. 2016.
Deacetylation of fungal exopolysaccharide mediates adhe-
sion and biofilm formation. mBio. 7(2):e00252e00216.
Lee MJ, Gravelat FN, Cerone RP, Baptista SD, Campoli PV,
Choe S-I, Kravtsov I, Vinogradov E, Creuzenet C, Liu H, et
al. 2014. Overlapping and distinct roles of Aspergillus fumi-
gatus UDP-glucose 4-epimerases in galactose metabolism
and the synthesis of galactose-containing cell wall poly-
saccharides. J Biol Chem. 289(3):12431256.
Lee MJ, Liu H, Barker BM, Snarr BD, Gravelat FN, Al Abdallah
Q, Gavino C, Baistrocchi SR, Ostapska H, Xiao T, et al.
2015. The fungal exopolysaccharide galactosaminogalac-
tan mediates virulence by enhancing resistance to neutro-
phil extracellular traps. PLoS Pathog. 11(10):e1005187.
Leonhardt I, Spielberg S, Weber M, Albrecht-Eckardt D, Bl
ass
M, Claus R, Barz D, Scherlach K, Hertweck C, L
offler J, et al.
2015. The fungal quorum-sensing molecule farnesol acti-
vates innate immune cells but suppresses cellular adaptive
immunity. mBio. 6(2):e00143.
Lewis K. 2008. Multidrug tolerance of biofilms and persister
cells. In: Romeo T, editor. Bacterial biofilms. Berlin,
Heidelberg: Springer. p. 107131.
Lewis K. 2010. Persister cells. Annu Rev Microbiol. 64(1):
357372.
Lian X, de Hoog GS. 2010. Indoor wet cells harbour melan-
ized agents of cutaneous infection. Med Mycol. 48(4):
622628.
Lin C-J, Sasse C, Gerke J, Valerius O, Irmer H, Frauendorf H,
Heinekamp T, Straßburger M, Tran VT, Herzog B, et al.
2015. Transcription factor SomA is required for adhesion,
development and virulence of the human pathogen
Aspergillus fumigatus. PLoS Pathog. 11(11):e1005205.
Liu L, Yu B, Sun W, Liang C, Ying H, Zhou S, Niu H, Wang Y,
Liu D, Chen Y. 2020. Calcineurin signaling pathway influ-
ences Aspergillus niger Biofilm formation by affecting
hydrophobicity and cell wall integrity. Biotechnol Biofuels.
13(1):54.
Liu W, Li L, Sun Y, Chen W, Wan Z, Li R, Liu W. 2012.
Interaction of the Echinocandin caspofungin with
Amphotericin B or Voriconazole against Aspergillus biofilms
in vitro. Antimicrob Agents Chemother. 56(12):64146416.
Livengood SJ, Drew RH, Perfect JR. 2020. Combination ther-
apy for invasive fungal infections. Curr Fungal Infect Rep.
14(1):4049.
Lohse MB, Gulati M, Craik CS, Johnson AD, Nobile CJ. 2020.
Combination of antifungal drugs and protease inhibitors
prevent Candida albicans biofilm formation and disrupt
mature biofilms. Front Microbiol. 11:1027.
Lopes M, Alves CT, Rama Raju B, Gonc¸alves MST, Coutinho
PJG, Henriques M, Belo I. 2014. Application of
Benzo[a]phenoxazinium chlorides in antimicrobial photo-
dynamic therapy of Candida albicans biofilms. J
Photochem Photobiol B. 141:9399.
L
opez CE. [ 2006. Dimorphism and pathogenesis of
Histoplasma capsulatum. Rev Argent Microbiol. 38(4):
235242. ].
Lorek J, P
oggeler S, Weide MR, Breves R, Bockm
uhl DP. 2008.
Influence of farnesol on the morphogenesis of Aspergillus
niger. J Basic Microbiol. 48(2):99103.
Loussert C, Schmitt C, Prevost M-C, Balloy V, Fadel E, Philippe
B, Kauffmann-Lacroix C, Latg
e JP, Beauvais A. 2010. In Vivo
Biofilm Composition of Aspergillus fumigatus. Cell
Microbiol. 12(3):405410.
Lugauskas A, Krikstaponis A, Sveistyte L. 2004. Airborne fungi
in industrial environmentspotential agents of respiratory
diseases. Ann Agric Environ Med. 11(1):1925.
Machado Vila TV, Sousa Quintanilha N, Rozental S. 2015.
Miltefosine is effective against Candida albicans and
Fusarium oxyporum nail biofilms in vitro. J Med Microbiol.
64(11):14361449.
Manavathu EK, Vager DL, Vazquez JA. 2014. Development
and antimicrobial susceptibility studies of in vitro mono-
microbial and polymicrobial biofilm models with
Aspergillus fumigatus and Pseudomonas aeruginosa. BMC
Microbiol. 14:53.
Manfiolli AO, dos Reis TF, de Assis LJ, de Castro PA, Silva LP,
Hori JI, Walker LA, Munro CA, Rajendran R, Ramage G, et
al. 2018. Mitogen Activated Protein Kinases (MAPK) and
protein phosphatases are involved in Aspergillus fumigatus
adhesion and biofilm formation. The Cell Surface. 1:4356.
Martins M, Henriques M, Lopez-Ribot JL, Oliveira R. 2012.
Addition of DNase improves the in vitro activity of antifun-
gal drugs against Candida albicans biofilms: DNase-anti-
fungal combination on candida biofilms. Mycoses. 55(1):
8085.
Matsubara VH, Wang Y, Bandara HMHN, Mayer MPA,
Samaranayake LP. 2016. Probiotic lactobacilli inhibit early
stages of Candida albicans biofilm development by reduc-
ing their growth, cell adhesion, and filamentation. Appl
Microbiol Biotechnol. 100(14):64156426.
Mello TP, Aor AC, Gonc¸alves DS, Seabra SH, Branquinha MH,
Santos ALS. 2016. Assessment of Biofilm Formation by
Scedosporium apiospermum,S. aurantiacum,S. minutispo-
rum and Lomentospora prolificans. Biofouling. 32(7):
737749.
Mello TP, Branquinha MH, Santos ALS. 2020. Biofilms formed
by Scedosporium and Lomentospora Species: focus on the
extracellular matrix. Biofouling. 36(3):308318.
Microbiology Society 2020. Microscopic, cultural and molecu-
lar evidence of disseminated invasive aspergillosis involv-
ing the lungs and the gastrointestinal tract. [Accessed 28
Nov 2020]. https://www.microbiologyresearch.org/content/
journal/jmm/10.1099/jmm.0.46394-0
Mishra R, Panda AK, De Mandal S, Shakeel M, Bisht SS, Khan
J. 2020. Natural anti-biofilm agents: strategies to control
biofilm-forming pathogens. Front Microbiol. 11:566325.
Mitchell KF, Zarnowski R, Andes DR. 2016. Fungal super glue:
the biofilm matrix and its composition, assembly, and
functions. PLoS Pathog. 12(9):e1005828.
Morisse H, Heyman L, Sala
un M, Favennec L, Picquenot JM,
Bohn P, Thiberville L. 2013. In vivo molecular microimag-
ing of pulmonary aspergillosis. Med Mycol. 51(4):352360.
Mowat E, Butcher J, Lang S, Williams C, Ramage G. 2007.
Development of a simple model for studying the effects
of antifungal agents on multicellular communities of
Aspergillus fumigatus. J Med Microbiol. 56(Pt 9):12051212.
Mowat E, Lang S, Williams C, McCulloch E, Jones B, Ramage
G. 2008. Phase-dependent antifungal activity against
Aspergillus niger developing multicellular filamentous bio-
films. J Antimicrob Chemother. 62(6):12811284.
CRITICAL REVIEWS IN MICROBIOLOGY 21
Mowat E, Williams C, Jones B, Mcchlery S, Ramage G. 2009.
The characteristics of Aspergillus fumigatus mycetoma
development: is this a biofilm? Med Mycol. 47(s1):
S120S126.
Mukherjee PK, Chandra J, Yu C, Sun Y, Pearlman E,
Ghannoum MA. 2012. Characterization of Fusarium kera-
titis outbreak isolates: contribution of biofilms to anti-
microbial resistance and pathogenesis. Invest Ophthalmol
Vis Sci. 53(8):44504457.
M
uller F-MC, Seidler M, Beauvais A. 2011. Aspergillus fumiga-
tus biofilms in the clinical setting. Med Mycol. 49(S1):
S96S100.
Muszkieta L, Beauvais A, P
ahtz V, Gibbons JG, Anton Leberre
V, Beau R, Shibuya K, Rokas A, Francois JM, Kniemeyer O,
et al. 2013. Investigation of Aspergillus fumigatus biofilm
formation by various omicsapproaches. Front Microbiol.
4:13.
Naderi J, Giles C, Saboohi S, Griesser HJ, Coad BR. 2019.
Surface coatings with covalently attached anidulafungin
and micafungin prevent Candida albicans biofilm forma-
tion. J Antimicrob Chemother. 74(2):360364.
Nazir N, Siddique A, Khan NA. 2020. Inhibitory effects of anti-
fungal drugs on biofilm producing Aspergillus spp. recov-
ered from drinking water system; preprint.
Nett JE, Andes D. 2017. The role of biofilm matrix in media-
ting antifungal resistance. In: Berghuis A, Matlashewski G,
Wainberg MA, Sheppard D, editors. Handbook of anti-
microbial resistance. New York (NY): Springer. p. 369384.
Niimi M, Firth NA, Cannon RD. 2010. Antifungal drug resist-
ance of oral fungi. Odontology. 98(1):1525.
Nikoomanesh F, Roudbarmohammadi S, Khoobi M, Haghighi
F, Roudbary M. 2019. Design and Synthesis of
Mucoadhesive Nanogel Containing Farnesol: Investigation
of the Effect on HWP1, SAP6 and Rim101 Genes
Expression of Candida albicans in Vitro. Artif Cells
Nanomed Biotechnol. 47(1):6472.
Niu M, Steffan BN, Fischer GJ, Venkatesh N, Raffa NL,
Wettstein MA, Bok JW, Greco C, Zhao C, Berthier E, et
al. 2020. Fungal oxylipins direct programmed developmen-
tal switches in filamentous fungi. Nat Commun. 11(1):
5158.
Nascimento ACMOB, Dos Reis TF, de Castro PA, Hori JI, Bom
VLP, de Assis LJ, Ramalho LNZ, Rocha MC, Malavazi I,
Brown NA, et al. 2016. Mitogen activated protein kinases
SakAHOG1 and MpkC collaborate for Aspergillus fumigatus
virulence - Bruder Nascimento. Mol Microbiol. 100:
841859. https://onlinelibrary.wiley.com/doi/full/10.1111/
mmi.13354
Oliveira LT, Medina-Alarc
on KP, Singulani J, de L, Fregonezi
NF, Pires RH, Arthur RA, Fusco-Almeida AM, Mendes
Giannini MJS. 2020. Dynamics of mono- and dual-species
biofilm formation and interactions between
Paracoccidioides brasiliensis and Candida albicans. Front
Microbiol. 11:551256.
Oxford Academic 2020b. Candida albicans cell wall compo-
nents and farnesol stimulate the expression of both
inflammatory and regulatory cytokines in the murine
RAW264.7 macrophage cell line. Pathogens and Disease.
[Accessed 18 Oct 2020]. https://academic.oup.com/
femspd/article/60/1/63/537159
Oxford Academic. 2020a. Pathophysiological aspects of
Aspergillus colonization in disease. Medical Mycology.
[Accessed 28 Nov 2020]. https://academic.oup.com/mmy/
article/57/Supplement_2/S219/5098503.
Padder SA, Prasad R, Shah AH. 2018. Quorum sensing: a less
known mode of communication among Fungi. Microbiol
Res. 210:5158.
Palonen EK, Raina S, Brandt A, Meriluoto J, Keshavarz T, Soini
JT. 2017. Transcriptomic complexity of Aspergillus terreus
velvet gene family under the influence of butyrolactone I.
Microorganisms. 5(1):12.
Paris S, Debeaupuis J-P, Crameri R, Carey M, Charl
es F,
Pr
evost MC, Schmitt C, Philippe B, Latg
e JP. 2003. Conidial
hydrophobins of Aspergillus fumigatus. Appl Environ
Microbiol. 69(3):15811588.
Paulone S, Ardizzoni A, Tavanti A, Piccinelli S, Rizzato C,
Lupetti A, Colombari B, Pericolini E, Polonelli L, Magliani W,
et al. 2017. The synthetic killer peptide KP impairs Candida
albicans biofilm in vitro. PLoS One. 12(7):e0181278.
Peiqian L, Xiaoming P, Huifang S, Jingxin Z, Ning H, Birun L.
2014. Biofilm formation by Fusarium oxysporum F. Sp.
cucumerinum and susceptibility to environmental stress.
FEMS Microbiol Lett. 350(2):138145.
Peleg AY, Hogan DA, Mylonakis E. 2010. Medically important
bacterial-fungal interactions. Nat Rev Microbiol. 8(5):
340349.
Perumal P, Mekala S, Chaffin WL. 2007. Role for cell density
in antifungal drug resistance in Candida albicans biofilms.
Antimicrob Agents Chemother. 51(7):24542463.
Pettit RK, Repp KK, Hazen KC. 2010. Temperature affects the
susceptibility of cryptococcus neoformans biofilms to anti-
fungal agents. Med Mycol. 48(2):421426.
Pierce C, Vila T, Romo J, Montelongo-Jauregui D, Wall G,
Ramasubramanian A, Lopez-Ribot J. 2017. The Candida
albicans biofilm matrix: composition, structure and func-
tion. JoF. 3(1):14.
Pierce CG, Lopez-Ribot JL. 2013. Candidiasis drug discovery
and development: new approaches targeting virulence for
discovering and identifying new drugs. Expert Opin Drug
Discov. 8(9):11171126.
Pierce CG, Srinivasan A, Uppuluri P, Ramasubramanian AK,
L
opez-Ribot JL. 2013. Antifungal therapy with an emphasis
on biofilms. Curr Opin Pharmacol. 13(5):726730.
Pitangui NS, Sardi JCO, Silva JF, Benaducci T, Moraes da Silva
RA, Rodr
ıguez-Arellanes G, Taylor ML, Mendes-Giannini
MJS, Fusco-Almeida AM. 2012. Adhesion of Histoplasma
capsulatum to pneumocytes and biofilm formation on an
abiotic surface. Biofouling. 28(7):711718.
Plaetzer K, Krammer B, Berlanda J, Berr F, Kiesslich T. 2009.
Photophysics and photochemistry of photodynamic ther-
apy: fundamental aspects. Lasers Med Sci. 24(2):259268.
Raad I. 1998. Intravascular-catheter-related infections. Lancet.
351(9106):893898.
Rahimi H ,Roudbarmohammadi S, Delavari HH, Roudbary M.
2019. Antifungal effects of indolicidin-conjugated gold
nanoparticles against fluconazole-resistant strains of
Candida albicans isolated from patients with burn infec-
tion. Inter J Nanomed. 14:53235338.
Raina S, De Vizio D, Palonen EK, Odell M, Brandt AM, Soini
JT, Keshavarz T. 2012. Is quorum sensing involved in lova-
statin production in the filamentous fungus Aspergillus ter-
reus? Process Biochem. 47(5):843852.
22 M. ROUDBARY ET AL.
Raina S, Odell M, Keshavarz T. 2010. Quorum sensing as a
method for improving sclerotiorin production in penicil-
lium sclerotiorum. J Biotechnol. 148(2-3):9198.
Rajendran R, Mowat E, McCulloch E, Lappin DF, Jones B,
Lang S, Majithiya JB, Warn P, Williams C, Ramage G. 2011.
Azole resistance of Aspergillus fumigatus biofilms is partly
associated with efflux pump activity. Antimicrob Agents
Chemother. 55(5):20922097.
Rajendran R, Williams C, Lappin DF, Millington O, Martins M,
Ramage G. 2013. Extracellular DNA release acts as an anti-
fungal resistance mechanism in mature Aspergillus fumiga-
tus biofilms. Eukaryotic Cell. 12(3):420429.
Ramage G, Mowat E, Jones B, Williams C, Lopez-Ribot J.
2009. Our current understanding of fungal biofilms. Crit
Rev Microbiol. 35(4):340355.
Ramage G, Rajendran R, Sherry L, Williams C. 2012. Fungal
biofilm resistance. Int J Microbiol. 2012:528521528514.
Ramage G, Saville SP, Thomas DP, L
opez-Ribot JL. 2005.
Candida biofilms: an update. Eukaryot Cell. 4(4):633638.
Redding S, Bhatt B, Rawls HR, Siegel G, Scott K, Lopez-Ribot
J. 2009. Inhibition of Candida albicans biofilm formation
on denture material. Oral Surg Oral Med Oral Pathol Oral
Radiol Endodontol. 107(5):669672.
Ribes JA, Vanover-Sams CL, Baker DJ. 2000. Zygomycetes in
human disease. Clin Microbiol Rev. 13(2):236301.
Robbins N, Uppuluri P, Nett J, Rajendran R, Ramage G,
Lopez-Ribot JL, Andes D, Cowen LE. 2011. Hsp90 governs
dispersion and drug resistance of fungal biofilms. PLoS
Pathog. 7(9):e1002257.
Rodrigues CF,
Cern
akov
a L. 2020. Farnesol and tyrosol: sec-
ondary metabolites with a crucial quorum-sensing role in
Candida biofilm development. Genes. 11(4):444.
Rodrigues CF, Gonc¸alves B, Rodrigues ME, Silva S, Azeredo J,
Henriques M. 2017. The effectiveness of voriconazole in
therapy of Candida glabratasbiofilms oral infections and
its influence on the matrix composition and gene expres-
sion. Mycopathologia. 182(7-8):653664.
Rodrigues CF, Rodrigues ME, Silva S, Henriques M. 2017.
Candida glabrata biofilms: how far have we come? JoF.
3(1):11.
Roemer T, Krysan DJ. 2014. Antifungal Drug Development:
Challenges, Unmet Clinical Needs, and New Approaches.
Cold Spring Harbor Perspectives in Medicine. 4(5):
a019703a019703.
Rosenberg A, Ene IV, Bibi M, Zakin S, Segal ES, Ziv N, Dahan
AM, Colombo AL, Bennett RJ, Berman J. 2018. Antifungal
tolerance is a subpopulation effect distinct from resistance
and is associated with persistent candidemia. Nat
Commun. 9(1):2470.
Sadiq FA, Yan B, Tian F, Zhao J, Zhang H, Chen W. 2019.
Lactic acid bacteria as antifungal and anti-mycotoxigenic
agents: a comprehensive review. Compr Rev Food Sci
Food Saf. 18(5):14031436.
S
anchez-Herrera R, Flores-Villavicencio LL, Pichardo-Molina JL,
Castruita-Dom
ınguez JP, Aparicio-Fern
andez X, Sabanero
L
opez M, Villag
omez-Castro JC. 2021. Analysis of biofilm
formation by Sporothrix schenckii. Med Mycol. 59(1):3140.
Sardi J, de CO, Pitangui N, de S, Voltan AR, Braz JD, Machado
MP, Almeida AMF, Giannini MJSM. 2015. In vitro
Paracoccidiodes brasiliensis biofilm and gene expression of
adhesins and hydrolytic enzymes. Virulence. 6(6):642651.
Sardi JCO, Scorzoni L, Bernardi T, Fusco-Almeida AM, Mendes
Giannini MJS. 2013. Candida species: current epidemi-
ology, pathogenicity, biofilm formation, natural antifungal
products and new therapeutic options. J Med Microbiol.
62(Pt 1):1024.
Sardi JDCO, Pitangui NDS, Rodr
ıguez-Arellanes G, Taylor ML,
Fusco-Almeida AM, Mendes-Giannini MJS. 2014. Highlights
in pathogenic fungal biofilms. Rev Iberoam Micol. 31(1):
2229.
Sav H, Rafati H,
Oz Y, Dalyan-Cilo B, Ener B, Mohammadi F,
Ilkit M, van Diepeningen AD, Seyedmousavi S. 2018.
Biofilm formation and resistance to fungicides in clinically
relevant members of the fungal genus fusarium. JOF. 4(1):
16.
Schelenz S, Hagen F, Rhodes JL, Abdolrasouli A, Chowdhary
A, Hall A, Ryan L, Shackleton J, Trimlett R, Meis JF, et al.
2016. First hospital outbreak of the globally emerging
Candida auris in a European Hospital. Antimicrob Resist
Infect Control. 5(1). DOI:10.1186/s13756-016-0132-5.
Schimmel TG, Coffman AD, Parsons SJ. 1998. Effect of butyro-
lactone I on the producing fungus, Aspergillus terreus.
Appl Environ Microbiol. 64(10):37073712.
Schumacher MA, Piro KM, Xu W, Hansen S, Lewis K, Brennan
RG. 2009. Molecular mechanisms of HipA-mediated multi-
drug tolerance and its neutralization by HipB. Science.
323(5912):396401.
Seidler MJ, Salvenmoser S, M
uller F-MC. 2008. Aspergillus
fumigatus forms biofilms with reduced antifungal drug
susceptibility on bronchial epithelial cells. Antimicrob
Agents Chemother. 52(11):41304136.
Semighini CP, Hornby JM, Dumitru R, Nickerson KW, Harris
SD. 2006. Farnesol-induced apoptosis in Aspergillus nidu-
lans reveals a possible mechanism for antagonistic interac-
tions between fungi. Mol Microbiol. 59(3):753764.
Semighini CP, Murray N, Harris SD. 2008. Inhibition of
Fusarium graminearum growth and development by farne-
sol. FEMS Microbiol Lett. 279(2):259264.
Seneviratne CJ, Wang Y, Jin L, Abiko Y, Samaranayake LP.
2008. Candida albicans biofilm formation is associated
with increased anti-oxidative capacities. Proteomics. 8(14):
29362947.
Shapiro RS, Chavez A, Porter CBM, Hamblin M, Kaas CS,
DiCarlo JE, Zeng G, Xu X, Revtovich AV, Kirienko NV, et al.
2018. A CRISPR-Cas9-based gene drive platform for gen-
etic interaction analysis in Candida albicans. Nat Microbiol.
3(1):7382.
Sharma D, Misba L, Khan AU. 2019. Antibiotics versus biofilm:
an emerging battleground in microbial communities.
Antimicrob Resist Infect Control. 8(1):76.
Sharma R, Jangid K. 2015. Fungal quorum sensing inhibitors.
In: Kalia VC, editor. Quorum sensing vs quorum quench-
ing: a battle with no end in sight. New Delhi: Springer
India. p. 237257. https://doi.org/10.1007/978-81-322-
1982-8_20.
Sheppard DC, Howell PL. 2016. Biofilm exopolysaccharides of
pathogenic fungi: lessons from bacteria. J Biol Chem.
291(24):1252912537.
Sheppard DC. 2011. Molecular mechanism of Aspergillus fumi-
gatus adherence to host constituents. Curr Opin Microbiol.
14(4):375379.
Sherry L, Ramage G, Kean R, Borman A, Johnson EM,
Richardson MD, Rautemaa-Richardson R. 2017. Biofilm-
CRITICAL REVIEWS IN MICROBIOLOGY 23
forming capability of highly virulent, multidrug-resistant
Candida auris. Emerg Infect Dis. 23(2):328331.
Sharifi A, Mohammadzadeh A, Zahraei Salehi T, Mahmoodi P.
2018. Antibacterial, antibiofilm and antiquorum sensing
effects of Thymus daenensis and Satureja hortensis essen-
tial oils against Staphylococcus aureus isolates. J Appl
Microbiol. 124(2):379388. DOI:10.1111/jam.13639.
Shishodia SK, Tiwari S, Shankar J. 2019. Resistance mechan-
ism and proteins in Aspergillus species against antifungal
agents. Mycology. 10(3):151165.
Shokouhfard M, Kermanshahi RK, Shahandashti RV, Feizabadi
MM, Teimourian S. 2015. The Inhibitory Effect of A.
Lactobacillus acidophilus derived biosurfactant on biofilm
producer serratia marcescens. Iran J Basic Med Sci. 18(10):
10011007.
Shopova I, Bruns SM, Thywissen A, Kniemeyer O, Brakhage
AA, Hillmann F. 2013. Extrinsic extracellular DNA leads to
biofilm formation and colocalizes with matrix polysacchar-
ides in the human pathogenic fungus Aspergillus fumiga-
tus. Front Microbiol. 4:141.
Sil A, Andrianopoulos A. 2014. Thermally dimorphic human
fungal pathogens-polyphyletic pathogens with a conver-
gent pathogenicity trait. Cold Spring Harb Perspect Med.
5(8):a019794.
Singh R, Kumari A, Kaur K, Sethi P, Chakrabarti A. 2018.
Relevance of antifungal penetration in biofilm-associated
resistance of Candida albicans and non- albicans Candida
species. J Med Microbiol. 67(7):922926.
Singh R, Shivaprakash MR, Chakrabarti A. 2011. Biofilm for-
mation by zygomycetes: quantification, structure and
matrix composition. Microbiology. 157(Pt 9):26112618.
Singhal D, Baker L, Wormald P-J, Tan L. 2011. Aspergillus
fumigatus biofilm on primary human sinonasal epithelial
culture. Am J Rhinol Allergy. 25(4):219225.
Smith M, Mcginnis MR. 2005. Fusarium sporodochia on cuta-
neous wounds. Med Mycol. 43(1):8386.
Sorrentino F, Roy I, Keshavarz T. 2010. Impact of linoleic acid
supplementation on lovastatin production in Aspergillus
terreus cultures. Appl Microbiol Biotechnol. 88(1):6573.
Srinivasan R, Devi KR, Kannappan A, Pandian SK, Ravi AV.
2016. Piper betle and its bioactive metabolite phytol miti-
gates quorum sensing mediated virulence factors and bio-
film of nosocomial pathogen Serratia marcescens in vitro. J
Ethnopharmacol. 193:592603.
Sugui JA, Pardo J, Chang YC, M
ullbacher A, Zarember KA,
Galvez EM, Brinster L, Zerfas P, Gallin JI, Simon MM, et al.
2007. Role of LaeA in the regulation of Alb1, GliP, conidial
morphology, and virulence in Aspergillus fumigatus.
Eukaryot Cell. 6(9):15521561.
Sun W, Liu L, Yu Y, Yu B, Liang C, Ying H, Liu D, Chen Y.
2020. Biofilm-related, time-series transcriptome and gen-
ome sequencing in xylanase-producing Aspergillus niger
SJ1. ACS Omega. 5(31):1973719746.
Sunde M, Kwan AHY, Templeton MD, Beever RE, Mackay JP.
2008. Structural analysis of hydrophobins. Micron. 39(7):
773784.
Tan Y, Leonhard M, Moser D, Schneider-Stickler B. 2017.
Inhibition activity of lactobacilli supernatant against fun-
gal-bacterial multispecies biofilms on silicone. Microb
Pathog. 113:197201.
Tits J, Cammue BPA, Thevissen K. 2020. Combination therapy
to treat fungal biofilm-based infections. IJMS. 21(22):8873.
Tsitsigiannis DI, Bok J-W, Andes D, Nielsen KF, Frisvad JC,
Keller NP. 2005. Aspergillus cyclooxygenase-like enzymes
are associated with prostaglandin production and viru-
lence. Infect Immun. 73(8):45484559.
Tsitsigiannis DI, Keller NP. 2006. Oxylipins act as determi-
nants of natural product biosynthesis and seed coloniza-
tion in Aspergillus nidulans. Mol Microbiol. 59(3):882892.
Tsitsigiannis DI, Keller NP. 2007. Oxylipins as developmental
and host-fungal communication signals. Trends Microbiol.
15(3):109118.
Tu B, Yin G, Li H. 2020. Synergistic Effects of Vorinostat
(SAHA) and azoles against Aspergillus species and their
biofilms. BMC Microbiol. 20(1):28.
Upadhyay SK, Mahajan L, Ramjee S, Singh Y, Basir SF, Madan
T. 2009. Identification and characterization of a laminin-
binding protein of Aspergillus fumigatus: extracellular
thaumatin domain protein (AfCalAp). J Med Microbiol.
58(Pt 6):714722.
Vahedi Shahandashti R, Kasra Kermanshahi R, Ghadam P.
2016. The inhibitory effect of bacteriocin produced by
Lactobacillus acidophilus ATCC 4356 and Lactobacillus plan-
tarum ATCC 8014 on planktonic cells and biofilms of
Serratia marcescens. Turk J Med Sci. 46(4):11881196.
Vahedi-Shahandashti R, Lass-Fl
orl C. 2020. Lass-Fl
orl, C. novel
antifungal agents and their activity against Aspergillus spe-
cies. JoF. 6(4):213.
Vera-Gonz
alez N, Shukla A. 2020. Advances in biomaterials
for the prevention and disruption of Candida biofilms.
Front Microbiol. 11:538602.
Vilela SF, Barbosa JO, Rossoni RD, Santos JD, Prata MC,
Anbinder AL, Jorge AO, Junqueira JC. 2015. Lactobacillus
acidophilus ATCC 4356 Inhibits Biofilm Formation by C.
albicans and attenuates the experimental candidiasis in
Galleria mellonella. Virulence. 6(1):2939.
Villena GK, Gutierrez-Correa M. 2006. Production of cellulase
by Aspergillus niger biofilms developed on polyester cloth.
Lett Appl Microbiol. 43(3):262268.
Vladkova TG, Staneva AD, Gospodinova DN. 2020. Surface
engineered biomaterials and ureteral stents inhibiting bio-
film formation and encrustation. Surf Coat Technol . 404:
126424.
Vlamakis H. 2011. The world of biofilms. Virulence. 2(5):
431434.
Wang X, Cai W, van den Ende AHGG, Zhang J, Xie T, Xi L, Li
X, Sun J, de Hoog S. 2018. Indoor wet cells as a habitat
for melanized fungi, opportunistic pathogens on humans
and other vertebrates. Sci Rep. 8(1):7685.
Wang X, Wang Y, Zhou Y, Wei X. 2014. Farnesol induces
apoptosis-like cell death in the pathogenic fungus
Aspergillus flavus. Mycologia. 106(5):881888.
Waters CM, Bassler BL. 2005. Quorum sensing: cell-to-cell
communication in bacteria. Annu Rev Cell Dev Biol. 21:
319346.
Weiland-Br
auer N, Malek I, Schmitz RA. 2019. Metagenomic
quorum quenching enzymes affect biofilm formation of
Candida albicans and Staphylococcus epidermidis. PLoS
One. 14(1):e0211366.
Westwater C, Balish E, Schofield DA. 2005. Candida albicans-
conditioned medium protects yeast cells from oxidative
stress: a possible link between quorum sensing and oxida-
tive stress resistance. Eukaryot Cell. 4(10):16541661.
24 M. ROUDBARY ET AL.
Williams HE, Steele JCP, Clements MO, Keshavarz T. 2012.
Keshavarz, T. c-heptalactone is an endogenously produced
quorum-sensing molecule regulating growth and second-
ary metabolite production by Aspergillus nidulans. Appl
Microbiol Biotechnol. 96(3):773781.
Williams TC, Nielsen LK, Vickers CE, 2013. Engineered quorum
sensing using pheromone-mediated cell-to-cell communi-
cation in Saccharomyces cerevisiae. ACS Synth Biol. 2(3):
136149.
Winkelstr
oter LK, Bom VLP, de Castro PA, Ramalho LNZ,
Goldman MHS, Brown NA, Rajendran R, Ramage G, Bovier
E, dos Reis TF, et al. 2015. High osmolarity glycerol
response PtcB phosphatase is important for Aspergillus
fumigatus virulence. Mol Microbiol. 96(1):4254.
Wongsuk T, Pumeesat P, Luplertlop N. 2016. Fungal quorum
sensing molecules: role in fungal morphogenesis and
pathogenicity: quorum sensing in fungi. J Basic Microbiol.
56(5):440447.
W
osten HAB. 2001. Hydrophobins: multipurpose proteins.
Annu Rev Microbiol. 55(1):625646.
Wuren T, Toyotome T, Yamaguchi M, Takahashi-Nakaguchi A,
Muraosa Y, Yahiro M, Wang D-N, Watanabe A, Taguchi H,
Kamei K. 2014. Effect of serum components on biofilm for-
mation by aspergillus fumigatus and other Aspergillus spe-
cies. Jpn J Infect Dis. 67(3):172179.
Wuyts J, Dijck PV, Holtappels M. 2018. Fungal persister cells:
the basis for recalcitrant infections? PLoS Pathog. 14(10):
e1007301.
Zhang X, Sun X, Wang Z, Zhang Y, Hou W. 2012. Keratitis-
associated fungi form biofilms with reduced antifungal
drug susceptibility. Invest Ophthalmol Vis Sci. 53(12):
77747778.
Zhang Y, Li J, Liu F, Yan H, Li J, Zhang X, Jha AK. 2019.
Specific quorum sensing signal molecules inducing the
social behaviors of microbial populations in anaerobic
digestion. Bioresour Technol. 273:185195.
Zheng H, Kim J, Liew M, Yan JK, Herrera O, Bok J, Kelleher
NL, Keller NP, Wang Y. 2015. Redox metabolites signal pol-
ymicrobial biofilm development via the NapA oxidative
stress cascade in Aspergillus. Curr Biol. 25(1):2937.
CRITICAL REVIEWS IN MICROBIOLOGY 25
... However, taking into account the key characteristics of both bacterial and yeast biofilms -namely, that cells are embedded in an ECM, show increased stress tolerance and AMR, and altered gene expression when compared to pelagic cells, it is possible to identify filamentous fungal structures which are best described as biofilm [14,29]. As with other types of biofilm, those formed by filamentous fungi are best studied where they impinge upon human health, including biofilms formed by opportunistic human pathogens in genera such as Aspergillus and Fusarium [30]. ...
... In Aspergillus fumigatus, ECM is highly heterogeneous and predominantly composed of EPS such as α-glucans, galactosaminogalactan (GAG) and galactomannan. Other components include polyols, lipids, DNA and proteins [30]. Interestingly, GAG appears to function in many of the key biofilm characteristics, being involved in adhesion to surfaces, biofilm architecture and stability, AFR and host immune evasion [31]. ...
... Development of these filamentous biofilm structures can be characterised as a step-wise process beginning with adhesion, progressing through the formation of microcolonies, which may involve the development of hyphal bundles or layers, anastomosis and ECM secretion, and finally the maturation into a complex biofilm structure of multiple hyphal layers embedded in ECM [29,32,33]. Importantly, filamentous biofilms may be disrupted by the Candida quorum sensing molecule, farnesol, suggesting that cell-cell communication is important within these structures [30,35]. AFR and other forms of stress resistance are, again, associated with the ECM and with reduced metabolism at the centre of the biofilm, but there is also evidence of differential expression of ergosterol biosynthetic genes [31]. ...
Preprint
Full-text available
Zymoseptoria tritici is an economically damaging fungal pathogen of wheat which is able to survive for long periods on the leaf surface. In this environment, the fungus is exposed to many stresses, including fluctuations in temperature, UV radiation, drying, and foliar fungicide applications. We observed biofilm-like cell aggregations on the surface of wheat leaves infected with Z. tritici in both compatible and non-compatible interactions. A literature search revealed few examples of plant pathogenic fungi forming biofilms, but demonstrated that such biofilms have a number of key characteristics, which are shared with other biofilm forming microbes. These include adhesion of cells to the substrate, production of extracellular matrix, altered gene expression and increased tolerance to various stresses. We used a combination of microscopy, qRT-PCR and stress tolerance assays to determine whether putative Z. tritici biofilms possessed these diagnostic characteristics. We show that Z. tritici biofilms resemble in form and function those formed by other filamentous and dimorphic fungi, producing extra-cellular matrix in which cells become embedded, creating layers of hyphal bundles and blastospores. These biofilms show increased tolerance to drying and high temperature. There is also isolate-dependent resistance to reactive oxygen stress and the fungicide carboxin. Taken together, these findings demonstrate that Z. tritici is indeed able to produce biofilms.
... M. canis biofilms are composed of a multidirectionally expanded network of hyphae linked together by a polysaccharide extracellular matrix [16]. Biofilm reduces the penetrability of antifungal agents, thus contributing to treatment failure and recurrent infection [39,40]. The inhibitory effect of antifungal agents on biofilm formation was observed at concentrations higher than those required to inhibit the growth of planktonic cells [41]. ...
... The inhibitory effect of antifungal agents on biofilm formation was observed at concentrations higher than those required to inhibit the growth of planktonic cells [41]. Fungal biofilm formation is a key factor in fungal virulence, persistence, and invasion as well as recurrent fungal infections and conventional antifungal resistance [40,42]. The time-dependent adherence of arthroconidia was observed, starting at 2 h and up to 6 h after inoculation. ...
... During the early stage of C. albicans biofilm formation, genes encoding efflux pumps are upregulated, thereby mediating antifungal resistance [46]. Developing new compounds or alternative inhibitors to treat biofilm-related drug resistant fungal infections is essential to veterinary and human medicine [40,42,47]. In this study, fluconazole (4-16 µg/mL) had no effect on the mature biofilms of M. canis isolates. ...
Article
Full-text available
Microsporum canis is an important zoonotic fungus that causes dermatophytosis in domestic animals and their owners. Domestic cats are the primary reservoir for M. canis. Antifungal drugs frequently produce adverse effects on the host animal, increasing the demand for novel alternative treatments derived from nature. We evaluated the antifungal activity of Coleus amboinicus essential oil (CEO) and ethanolic extracts (CEE) against M. canis in planktonic and biofilm growth. Twelve clinical isolates of M. canis were identified in feline dermatophyte samples. Using GC-MS, 18 compounds were identified in CEO, with carvacrol being the major constituent. HPLC analysis of CEE revealed that it contained rosmarinic acid, apigenin, and caffeic acid. The planktonic growth of all M. canis isolates was inhibited by C. amboinicus extracts. The minimum inhibitory concentration at which ≥50% of the isolates were inhibited (MIC50) was 128 µg/mL (32–256 µg/mL) for both CEO and CEE. The MIC90 values of CEO and CEE were 128 and 256 µg/mL, respectively. CEO at MIC (128 µg/mL) and 2× MIC (256 µg/mL) significantly inhibited the biofilm formation of weak, moderate, and strong biofilm-producing M. canis. CEE at 2× MIC (256 µg/mL) significantly inhibited the biofilm formation of all isolates. Overall, C. amboinicus extracts inhibited planktonic growth and exhibited a significant antibiofilm effect against M. canis. Thus, C. amboinicus is a potential source of natural antifungal compounds.
... It is clear that invasive medical devices are susceptible to colonization by microorganisms present in the environment and in the human mycobiota. Knowledge about the formation of fungal biofilms on these devices is still relatively recent [47,55,69]. The biofilm formation can make the device an important source for the spread and development of a serious clinical infection. ...
Article
Full-text available
Most current protocols for the diagnosis of fungal infections are based on culture-dependent methods that allow the evaluation of fungal morphology and the identification of the etiologic agent of mycosis. Most current protocols for the diagnosis of fungal infections are based on culture-dependent methods that enable the examination of the fungi for further identification of the etiological agent of the mycosis. The isolation of fungi from pure cultures is typically recommended, as when more than one species is identified, the second agent is considered a contaminant. Fungi mostly survive in highly organized communities that provoke changes in phenotypic profile, increase resistance to antifungals and environmental stresses, and facilitate evasion from the immune system. Mixed fungal biofilms (MFB) harbor more than one fungal species, wherein exchange can occur that potentialize the effects of these virulence factors. However, little is known about MFB and their role in infectious processes, particularly in terms of how each species may synergistically contribute to the pathogenesis. Here, we review fungi present in MFB that are commensals of the human body, forming the mycobiota, and how their participation in MFB affects the maintenance of homeostasis. In addition, we discuss how MFB are formed on both biotic and abiotic surfaces, thus being a significant reservoir of microorganisms that have already been associated in infectious processes of high morbidity and mortality.
Article
Full-text available
Purpose of Review Farnesol (FOH) is a quorum-sensing molecule with potential as an antifungal drug. Given the growing concern about fungal drug resistance, exploring new solutions is crucial. Therefore, summarizing the antifungal activity of FOH is expected to be the basis for further FOH research and application. Herein, we reviewed the in vitro and in vivo antifungal efficacy of FOH alone and in combination with conventional antifungal drugs, as well as its mechanisms of action. Furthermore, we discussed the prospects and challenges of the FOH application in detail. Recent Findings Recent studies have revealed that FOH can target various aspects, such as reactive oxygen species production, induction of apoptosis, and modulation of virulence factors, to inhibit fungal growth and reduce fungal pathogenicity, thereby exerting its antifungal activity. Furthermore, FOH can suppress resistance-associated genes, such as those of biofilm and ergosterol, so as to enhance the fungicidal effectiveness of conventional antifungal drugs. However, the action mechanism of FOH on drug efflux pumps remains unclear and warrants further investigation. Summary FOH can prevent and treat fungal infections. It exerts significant antimicrobial effects on fungal planktonic and biofilm cells, enhances the antimicrobial efficacy of conventional antifungal drugs, and reverses and reduces fungal drug resistance. However, further in vitro and in vivo studies are needed to assess the safety of FOH due to potential adverse effects on immune cells and other factors.
Article
Full-text available
Histoplasmosis is a respiratory disease caused by Histoplasma capsulatum, a dimorphic fungus, with high mortality and morbidity rates, especially in immunocompromised patients. Considering the small existing therapeutic arsenal, new treatment approaches are still required. Chitosan, a linear polysaccharide obtained from partial chitin deacetylation, has anti-inflammatory, antimicrobial, biocompatibility, biodegradability, and non-toxicity properties. Chitosan with different deacetylation degrees and molecular weights has been explored as a potential agent against fungal pathogens. In this study, the chitosan antifungal activity against H. capsulatum was evaluated using the broth microdilution assay, obtaining minimum inhibitory concentrations (MIC) ranging from 32 to 128 µg/mL in the filamentous phase and 8 to 64 µg/mL in the yeast phase. Chitosan combined with classical antifungal drugs showed a synergic effect, reducing chitosan’s MICs by 32 times, demonstrating that there were no antagonistic interactions relating to any of the strains tested. A synergism between chitosan and amphotericin B or itraconazole was detected in the yeast-like form for all strains tested. For H. capsulatum biofilms, chitosan reduced biomass and metabolic activity by about 40% at 512 µg/mL. In conclusion, studying chitosan as a therapeutic strategy against Histoplasma capsulatum is promising, mainly considering its numerous possible applications, including its combination with other compounds.
Article
Aims This study aimed to investigate whether berberine (BBR) can inhibit the iron reduction mechanism of Candida albicans, lowering the iron uptake of the yeast and perhaps having antimicrobial effects. Methods and results We determined that BBR may cause extensive transcriptional remodeling in C. albicans and that iron permease Ftr1 played a crucial role in this process through eukaryotic transcriptome sequencing. Mechanistic research showed that BBR might selectively inhibit the iron reduction pathway to lower the uptake of exogenous iron ions, inhibiting C. albicans from growing and metabolizing. Subsequent research revealed that BBR caused significant mitochondrial dysfunction, which triggered the process of mitochondrial autophagy. Moreover, we discovered that C. albicans redox homeostasis, susceptibility to antifungal drugs, and hyphal growth are all impacted by the suppression of this mechanism by BBR. Conclusions The iron reduction mechanism in C. albicans is disrupted by BBR, which disrupts mitochondrial function and inhibits fungal growth. These findings highlight the potential promise of BBR in antifungal applications.
Article
In patients with compromised immunity, invasive fungal infections represent a significant cause of mortality. Given the limited availability and drawbacks of existing first-line antifungal drugs, there is a growing interest in exploring novel targets that could facilitate the development of new antifungal agents or enhance the effectiveness of conventional ones. While previous studies have extensively summarized new antifungal targets inherent in fungi for drug development purposes, the exploration of potential targets for novel antifungal drug delivery strategies has received less attention. In this review, we provide an overview of recent advancements in new antifungal drug delivery strategies that leverage novel targets, including those located in the physio-pathological barrier at the site of infection, the infection microenvironment, fungal-host interactions, and the fungal pathogen itself. The objective is to enhance therapeutic efficacy and mitigate toxic effects in fungal infections, particularly in challenging cases such as refractory, recurrent, and drug-resistant invasive fungal infections. We also discuss the current challenges and future prospects associated with target-driven antifungal drug delivery strategies, offering important insights into the clinical implementation of these innovative approaches.
Article
Microscopic filamentous fungi are ubiquitous microorganisms that adapt very easily to a variety of environmental conditions. Due to this adaptability, they can colonize a number of various surfaces where they are able to start forming biofilms. Life in the form of biofilms provides them with many benefits (increased resistance to desiccation, UV radiation, antimicrobial compounds, and host immune response). The aim of this study is to find a reliable and reproducible methodology to determine biofilm growth of selected microscopic filamentous fungi strains. Several methods (crystal violet staining, MTT assay, XTT assay, resazurin assay) for the determination of total biofilm biomass and its metabolic activity were tested on four fungi - Alternaria alternata, Aspergillus niger, Fusarium culmorum and Fusarium graminearum, and their biofilm was also imaged by spinning disc confocal microscopy using fluorescent dyes. A reproducible biofilm quantification method is essential for the subsequent testing of the biofilm growth suppression using antifungal agents or physical methods. Crystal violet staining was found to be a suitable method for the determination of total biofilm biomass of selected strains, and the MTT assay for the determination of metabolic activity of the biofilms. Calcofluor white and Nile red fluorescent stains successfully dyed the hyphae of microscopic fungi.
Article
Full-text available
Patients with chromoblastomycosis (CBM) suffer chronic tissue lesions that are hard to treat. Considering that biofilm is the main growth lifestyle of several pathogens and it is involved with both virulence and resistance to antimicrobial drugs, we have investigated the ability of CBM fungi to produce this complex, organized and multicellular structure. Fonsecaea pedrosoi and Phialophora verrucosa conidial cells were able to adhere on a polystyrene abiotic substrate, differentiate into hyphae and produce a robust viable biomass containing extracellular matrix. Confocal laser scanning microscopy (CLSM) and scanning electron microscopy (SEM) showed the tridimensional architecture of the mature biofilms, revealing a dense network of interconnected hyphae, inner channels and amorphous extracellular polymeric material. Interestingly, the co-culture of each fungus with THP-1 macrophage cells, used as a biotic substrate, induced the formation of a mycelial trap covering and damaging the macrophages. In addition, the biofilm-forming cells of F. pedrosoi and P. verrucosa were more resistant to the conventional antifungal drugs than the planktonic-growing conidial cells. The efflux pump activities of P. verrucosa and F. pedrosoi biofilms were significantly higher than those measured in conidia. Taken together, the data pointed out the biofilm formation by CBM fungi and brought up a discussion of the relevance of studies about their antifungal resistance mechanisms.
Article
Full-text available
Dermatophytes, fungi that cause dermatophytosis, can invade keratinized tissues in humans and animals. The biofilm-forming ability of these fungi was described recently, and it may be correlated with the long treatment period and common recurrences of this mycosis. In this study, we evaluated the anti-dermatophytic and anti-biofilm activity of 2-hydroxychalcone (2-chalcone) in the dark and photodynamic therapy (PDT)-mediated and to determine its mechanism of action. Trichophyton rubrum and Trichophyton mentagrophytes strains were used in the study. The antifungal susceptibility test of planktonic cells, early-stage biofilms, and mature biofilms were performed using colorimetric methods. Topographies were visualized by scanning electron microscopy (SEM). Human skin keratinocyte (HaCat) monolayers were also used in the cytotoxicity assays. The mechanisms of action of 2-chalcone in the dark and under photoexcitation were investigated using confocal microscopy and the quantification of ergosterol, reactive oxygen species (ROS), and death induction by apoptosis/necrosis. All strains, in the planktonic form, were inhibited after treatment with 2-chalcone (minimum inhibitory concentration (MIC) = 7.8-15.6 mg/L), terbinafine (TRB) (MIC = 0.008–0.03 mg/L), and fluconazole (FLZ) (1–512 mg/L). Early-stage biofilm and mature biofilms were inhibited by 2-chalcone at concentrations of 15.6 mg/L and 31.2 mg/L in all tested strains. However, mature biofilms were resistant to all the antifungal drugs tested. When planktonic cells and biofilms (early-stage and mature) were treated with 2-chalcone-mediated PDT, the inhibitory concentrations were reduced by four times (2–7.8 mg/L). SEM images of biofilms treated with 2-chalcone showed cell wall collapse, resulting from a probable extravasation of cytoplasmic content. The toxicity of 2-chalcone in HaCat cells showed higher IC50 values in the dark than under photoexcitation. Further, 2-chalcone targets ergosterol in the cell and promotes the generation of ROS, resulting in cell death by apoptosis and necrosis. Overall, 2-chalcone-mediated PDT is a promising and safe drug candidate against dermatophytes, particularly in anti-biofilm treatment.
Article
Full-text available
An increasing number of people is affected by fungal biofilm-based infections, which are resistant to the majority of currently-used antifungal drugs. Such infections are often caused by species from the genera Candida, Aspergillus or Cryptococcus. Only a few antifungal drugs, including echinocandins and liposomal formulations of amphotericin B, are available to treat such biofilm-based fungal infections. This review discusses combination therapy as a novel antibiofilm strategy. More specifically, in vitro methods to discover new antibiofilm combinations will be discussed. Furthermore, an overview of the main modes of action of promising antibiofilm combination treatments will be provided as this knowledge may facilitate the optimization of existing antibiofilm combinations or the development of new ones with a similar mode of action.
Article
Full-text available
The cell wall is essential for fungal viability and is absent from human hosts; thus, drugs disrupting cell wall biosynthesis have gained more attention. Caspofungin is a member of a new class of clinically approved echinocandin drugs to treat invasive aspergillosis by blocking β-1,3-glucan synthase, thus damaging the fungal cell wall. Here, we demonstrate that caspofungin and other cell wall stressors can induce galactosaminogalactan (GAG)-dependent biofilm formation in the human pathogen Aspergillus fumigatus . We further identified SomA as a master transcription factor playing a dual role in both biofilm formation and cell wall homeostasis. SomA plays this dual role by direct binding to a conserved motif upstream of GAG biosynthetic genes and genes involved in cell wall stress sensors, chitin synthases, and β-1,3-glucan synthase. Collectively, these findings reveal a transcriptional control pathway that integrates biofilm formation and cell wall homeostasis and suggest SomA as an attractive target for antifungal drug development.
Article
Full-text available
The high clinical mortality and economic burden posed by invasive fungal infections (IFIs), along with significant agricultural crop loss caused by various fungal species, has resulted in the widespread use of antifungal agents. Selective drug pressure, fungal attributes, and host- and drug-related factors have counteracted the efficacy of the limited systemic antifungal drugs and changed the epidemiological landscape of IFIs. Species belonging to Candida, Aspergillus, Cryptococcus, and Pneumocystis are among the fungal species showing notable rates of antifungal resistance. Drug-resistant fungi from the environment are increasingly identified in clinical settings. Furthermore, we have a limited understanding about drug class-specific resistance mechanisms in emerging Candida species. Establishment of antifungal stewardship programs in both clinical and agricultural fields and inclusion of species identification, antifungal susceptibility testing, and therapeutic drug monitoring practices in the clinic can minimize the emergence of drug-resistant fungi. New antifungal drugs featuring promising therapeutic profiles have great promise to treat drug-resistant fungi in the clinical setting. Mitigating antifungal tolerance, a prelude to the emergence of resistance, also requires development of effective and fungal-specific adjuvants to be used in combination with systemic antifungals.
Article
Full-text available
Pathogenic microorganisms and their chronic pathogenicity are significant concerns in biomedical research. Biofilm-linked persistent infections are not easy to treat due to resident multidrug-resistant microbes. Low efficiency of various treatments and in vivo toxicity of available antibiotics drive the researchers toward the discovery of many effective natural anti-biofilm agents. Natural extracts and natural product-based anti-biofilm agents are more efficient than the chemically synthesized counterparts with lesser side effects. The present review primarily focuses on various natural anti-biofilm agents, i.e., phytochemicals, biosurfactants, antimicrobial peptides, and microbial enzymes along with their sources, mechanism of action via interfering in the quorum-sensing pathways, disruption of extracellular polymeric substance, adhesion mechanism, and their inhibitory concentrations existing in literature so far. This study provides a better understanding that a particular natural anti-biofilm molecule exhibits a different mode of actions and biofilm inhibitory activity against more than one pathogenic species. This information can be exploited further to improve the therapeutic strategy by a combination of more than one natural anti-biofilm compounds from diverse sources.
Article
Full-text available
Pathogenic microorganisms and their chronic pathogenicity are significant concerns in biomedical research. Biofilm-linked persistent infections are not easy to treat due to resident multidrug-resistant microbes. Low efficiency of various treatments and in vivo toxicity of available antibiotics drive the researchers toward the discovery of many effective natural anti-biofilm agents. Natural extracts and natural product-based anti-biofilm agents are more efficient than the chemically synthesized counterparts with lesser side effects. The present review primarily focuses on various natural anti-biofilm agents, i.e., phytochemicals, biosurfactants, antimicrobial peptides, and microbial enzymes along with their sources, mechanism of action via interfering in the quorum-sensing pathways, disruption of extracellular polymeric substance, adhesion mechanism, and their inhibitory concentrations existing in literature so far. This study provides a better understanding that a particular natural anti-biofilm molecule exhibits a different mode of actions and biofilm inhibitory activity against more than one pathogenic species. This information can be exploited further to improve the therapeutic strategy by a combination of more than one natural anti-biofilm compounds from diverse sources.
Article
Full-text available
Filamentous fungi differentiate along complex developmental programs directed by abiotic and biotic signals. Currently, intrinsic signals that govern fungal development remain largely unknown. Here we show that an endogenously produced and secreted fungal oxylipin, 5,8-diHODE, induces fungal cellular differentiation, including lateral branching in pathogenic Aspergillus fumigatus and Aspergillus flavus, and appressorium formation in the rice blast pathogen Magnaporthe grisea. The Aspergillus branching response is specific to a subset of oxylipins and is signaled through G-protein coupled receptors. RNA-Seq profiling shows differential expression of many transcription factors in response to 5,8-diHODE. Screening of null mutants of 33 of those transcription factors identifies three transcriptional regulators that appear to mediate the Aspergillus branching response; one of the mutants is locked in a hypo-branching phenotype, while the other two mutants display a hyper-branching phenotype. Our work reveals an endogenous signal that triggers crucial developmental processes in filamentous fungi, and opens new avenues for research on the morphogenesis of filamentous fungi.
Article
Full-text available
The oral cavity is a highly diverse microbial environment in which microorganisms interact with each other, growing as biofilms on biotic and abiotic surfaces. Understanding the interaction among oral microbiota counterparts is pivotal for clarifying the pathogenesis of oral diseases. Candida spp. is one of the most abundant fungi in the oral mycobiome with the ability to cause severe soft tissue lesions under certain conditions. Paracoccidioides spp., the causative agent of paracoccidioidomycosis, may also colonize the oral cavity leading to soft tissue damage. It was hypothesized that both fungi can interact with each other, increasing the growth of the biofilm and its virulence, which in turn can lead to a more aggressive infectivity. Therefore, this study aimed to evaluate the dynamics of mono-and dual-species biofilm growth of Paracoccidioides brasiliensis and Candida albicans and their infectivity using the Galleria mellonella model. Biomass and fungi metabolic activity were determined by the crystal violet and the tetrazolium salt reduction tests (XTT), respectively, and the colony-forming unit (CFU) was obtained by plating. Biofilm structure was characterized by both scanning electronic-and confocal laser scanning-microscopy techniques. Survival analysis of G. mellonella was evaluated to assess infectivity. Our results showed that dual-species biofilm with P. brasiliensis plus C. albicans presented a higher biomass, higher metabolic activity and CFU than their mono-species biofilms. Furthermore, G. mellonella larvae infected with P. brasiliensis plus C. albicans presented a decrease in the survival rate compared to those infected with P. brasiliensis or C. albicans, mainly in the form of biofilms. Our data indicate that P. brasiliensis and C. albicans coexistence is likely to occur on oral mucosal biofilms, as per in vitro and in vivo analysis. These data further widen the knowledge associated with the dynamics of fungal biofilm growth that can potentially lead to the discovery of new therapeutic strategies for these infections.
Article
Aims: Biofilms formed in drinking water distribution systems serve as a continuous source of fungal infections. Biofilms are thick aggregates of adherent microorganisms including pathogenic species of fungi. Respiratory diseases and skin allergy reactions are caused by drinking water containing biofilm forming fungus and bacteria. One of the main causes of nosocomial infections and respiratory diseases in hospitals is due to the fungal biofilm formation in machines, catheters and other surgical instruments. There are some antifungal drugs which are used to control biofilm formation to minimize the infection rate. Methodology and results: The present study was conducted to isolate and identify Aspergillus species which are the main fungal spp. responsible for the biofilm formation in drinking water and to check their antifungal susceptibility against antifungal drugs. The isolated fungal samples from drinking water were cultivated on Potato dextrose agar for the isolation of Aspergillus species. Isolated Aspergillus species were identified on the basis of cultural, morphological and microscopic examination. Then in-vitro ability of biofilm produced by isolated Aspergillus species was estimated using microtitre plate method and quantification by crystal violet assay. Antifungal susceptibility testing against isolated fungal spp. was done by antifungal drug Amphotericin B. Results: From results, it is concluded that drinking water of labs, hospitals and common water chillers were more prevelant by Aspergillus species whereas water from reverse osmosis plants showed negative results. From microtitre plate method and crystal violet assay, it was concluded that Aspergillus spp. are Susceptible against Amphotericin B drug as compared to miconazole. Keywords: Aspergillus spp; Biofilm; Drinking water; Disk diffusion method; Amphotericin B
Article
Protection of ureteral stents against infections is a significant current challenge raised by increasing the number of stentings and infections, microbial resistance to conventional antibiotics and multi-drugs treatments, side effects for the patients like damage of surrounding tissue and high morbidity. Surface engineered biomaterials and ureteral stents, reducing biofilm formation and encrustation would contribute to mitigation of the problem. The aim of this review is to present the progress during the last 5 years in the development of surface engineered antimicrobial biomaterials and stents with expectation to raise some new fruitful ideas in this direction. Strategies aimed at preventing, disrupting, or removing adherent microbes and biofilms from biomaterials and ureteral stents are its main subject. Various antibacterial agents, modifications, and coatings as well as preparation, deposition and characterization techniques are among the topics covered. A brief market analysis is also included covering the significance of ureteral stents associated infections and a mode of development of a biofilm. The review of the progress during the last 5 years shows a continuing interest in surface modification and coating employing three principal anti-biofilm strategies: 1) mechanical detachment; 2) killing microbial cells and 3) creation of low-adhesive surfaces. The known surface engineering solutions report a reduced, to some extent, biofilm formation and encrustation of biomaterials and ureteral stents, but none of them is able to totally stop their development. Some new trends are observed, such as complementary antimicrobial protection by coating and flow dynamic; biodegradable coatings releasing antimicrobial agent; quasi irremovable surface coatings, delivering drug or antimicrobial agent as well as new carbon and biodegradable materials; bacteriophages and phage cocktails, etc. Almost all of them are under intense in vitro studies; only few of them were studied in vivo animal models and none in humans. Some of the liquid infused coatings, already tested in animal models, seem to be the closest to clinical application but so far no one has been applied to ureteral stents. This review outlines some future research directions and major challenges in the surface engineered biomaterials and ureteral stents. It gives ideas how, by surface engineering, to approach more closely the “clean” ureteral stent to not allow microbial adhesion and encrustation and thus sharply reduce the ureteral stents associated infections and stent dysfunction. The most important prerequisite of the non-toxic “clean” surface is to be low adhesive. Its anti-biofilm performance could be improved by including bio-surfactant and/or inhibitor of quorum sensing (QS) and/or inhibitor of the crosslinking of the exopolymeric substances (EPSs). The key to identifying a “universal” surface that would repel/release all microbial cells is maybe hidden in the in-depth understanding of the mechanism of the initial reversible adsorption of the EPSs secreted by microbial cells, since it initiates the whole biological cascade of biofilm development.