mycobiota in chilean chilli capsicum annuum l. used for

64
Rodrigo Nicolas Rodriguez Quiroz Mycobiota in Chilean chilli Capsicum annuum L. used for Merkén’s production and its mycotoxin contamination Master's thesis Master in Biotechnology Work performed under the supervision of Professor Doctor Nelson Lima & Co-supervision of Professor Doctor Cledir Santos July of 2018

Upload: others

Post on 26-Jul-2022

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Mycobiota in Chilean chilli Capsicum annuum L. used for

Rodrigo Nicolas Rodriguez Quiroz

Mycobiota in Chilean chilli Capsicum annuum L.

used for Merkén’s production and its mycotoxin

contamination

Master's thesis

Master in Biotechnology

Work performed under the supervision of

Professor Doctor Nelson Lima

&

Co-supervision of

Professor Doctor Cledir Santos

July of 2018

Page 2: Mycobiota in Chilean chilli Capsicum annuum L. used for

“The best way to predict the future is to invent it.”

Alan Kay.

Page 3: Mycobiota in Chilean chilli Capsicum annuum L. used for

I

ACKNOWLEDGMENTS

First and foremost, I would like to thank my inspirational sources: my mother Gabriela Quiroz,

my father Luis Rodriguez and my grandfather Alamiro Quiroz for teach me that the key to success is

perseverance, hard work and deeply love what we do day by day.

I would also like to thank my supervisors, Prof. Dr. Nelson Lima and Prof. Dr. Cledir Santos, for

all the continuous support, dedication, patient guidance, constant availability, and for all the scientific,

social and cultural insights.

I would also like to thank my incredible Portuguese friends, Mariana Macedo, Claudio Menses,

Luis Passos, Rui Cardoso and my partner of life Filipa Carvalho for helping me in this long and hard

process away from my homeland and for showing me a different society and a new way of see the world.

I am grateful for the patience, support, help and friendship given by the researchers that daily

worked near me at the Mycology Laboratory, specially, Carla Santos, Ana Guimarães and Tiago Afonso.

I would also like to thank a very special person, Celia Soares, for all the continuous support, for

initiating me in the world of mycology, for his smiles, for her advice and for being an admirable person.

And, to all the people not mentioned above that help me in different occasions to finish this work

I leave to them my gratitude and I hope be able in the future to interchange in some way.

Thank you.

Page 4: Mycobiota in Chilean chilli Capsicum annuum L. used for

II

The present work was funded by The Advanced Human Capital Program, of the National

Commission for Scientific and Technological Research (CONICYT) Chile, Folio N° 7317076, Application

N° 73170764 and partially funded by the Universidad de La Frontera-UFRO (Temuco, Chile) through the

Project DIUFRO DI16-0135, DIUFRO DI18-0121, PIA17-0006 and EXT18-0043. Moreover, the present

work was partially being funded by the Portuguese Foundation for Science and Technology (FCT) under

the scope of the strategic funding of the UID/BIO/04469/2013 unit, COMPETE 2020 (POCI-01-0145-

FEDER-006684) and the BioTecNorte operation (NORTE-01-0145-FEDER-000004), funded by the

European Regional Development Fund through Norte2020—Programa Operacional Regional do Norte.

The Author is thankful to Miss Natalia Castillo (Chilean Social Worker) for her indefatigable support on

the development of this work and for the Mapuche Communities from the Region of La Araucanía (Chile)

for all support, contributions and for supplying red pepper and Merkén samples. Thanks are due to

Jéssica Sousa da Costa (PhD. Student of Science of Natural Resources, UFRO) and to Sebastián López

(BSc. Student of Biochemistry, UFRO) for their support in the fungal isolation evaluated in this Master

Thesis.

The experimental work developed herein was developed at the Laboratory of Mycology of the

Centre of Biological Engineering of the University of Minho, Braga, Portugal in collaboration with the Unity

of Metabolomics of the Scientific and Technological Bioresource Nucleus BIOREN-UFRO of the

Universidad de La Frontera, Temuco, Chile. The biological materials used are deposited at CCCT and

MUM culture collections.

Page 5: Mycobiota in Chilean chilli Capsicum annuum L. used for

III

RESUMO

No Chile, o fruto da planta Capsicum annuum L. cv. "Cacho de Cabra” é usado para a produção dum

piri-piri em pó chamado Merkén. As práticas usadas pelos produtores locais de Merkén são altamente empíricas

e não consideram a prevenção contra fungos micotoxigénicos pelo que, nos últimos anos, contaminações por

micotoxinas têm sido reportadas. Deste modo, o objetivo geral deste estudo é identificar os fungos potencialmente

micotoxigénicos no processo tradicional de produção do piri-piri chileno usado na produção de Merkén e detetar

contaminações por micotoxinas no Merkén produzido e comercializado no Chile. Amostras de frutos maduros de

C. annuum de 3 pontos de amostragem e amostras de Merkén foram obtidas de agricultores e mercados locais

da Zona Sul do Chile. O micobiota de frutos maduros foi isolada em meio de MEA, DRBC e DG18%. Estirpes

fúngicas isoladas foram identificadas usando técnicas clássicas e por biologia molecular. Para a taxonomia

clássica, foram avaliadas pelas características macro- e micro-morfológicas. A identificação molecular das estirpes

fúngicas foi realizada através do sequenciamento de β-tubulina (BenA) e ITS (ITS1-5.8S-ITS2). Além disso, genes

associados à produção de OTA (Otapks, Otanps, Otachl e Otatra) foram testados numa estirpe de Penicillium

crustosum e de P. verrucosum. Na extração e purificação de micotoxinas (ocratoxina A (OTA) e aflatoxinas (AFLs)

B1, B2, G1 e G2) foram usados métodos de extração líquido-líquido e colunas de imunoafinidade. Duzentos e seis

fungos foram isolados e identificados usando macro- e micro-morfologia clássica. Destes fungos, 190 foram

posteriormente identificados por técnicas de biologia molecular ao nível da espécie. Para genes associados à

produção de OTA, P. verrucosum apresenta resultados positivos para o gene Otanps e Otatra, e P. crustosum

apresenta resultados positivos apenas para o gene Otatra. Para contaminação do Merkén por micotoxinas, OTA

foi detetada em 100% amostras com valores de 0,79±0,05 a 19,81±0,70 µg/kg e quanto às aflatoxinas,

unicamente AFB1 foi detetada em 57% amostras, com valores que variaram de 0,29±0,37 a 1,67±0,32 µg/kg.

No computo geral, os resultados mostraram que o micobiota presente em frutos de C. annuum durante o processo

de produção de Merkén é cada vez mais seletiva para ocorrência de espécies de Aspergillus e Penicillium,

principalmente para a ocorrência cada vez mais seletiva de fungos potencialmente micotoxigénicos, como A. niger,

A. flavus, P. expansum e P. verrucosum. Estes resultados demonstram a importância de conhecer ainda mais o

micobiota e potenciais fungos micotoxigénicos presentes em cada etapa de produção do Merkén para estabelecer

pontos críticos de controlo para obtenção de produtos mais seguros e de melhor qualidade.

PALAVRAS-CHAVE: FUNGOS MICOTOXIGÉNICOS, PIMENTA CHILENA, MERKÉN, IDENTIFICAÇÃO DO MICOBIOTA

Page 6: Mycobiota in Chilean chilli Capsicum annuum L. used for

IV

ABSTRACT

In Chile, Capsicum annuum L. cv. "Cacho de Cabra” berry fruits are used for the manufacture of a

traditional chilli powder known as Merkén. The agricultural practices used by Merkén local producers are empirical

and do not consider the prevention of mycotoxigenic fungi and, in the last years, mycotoxin contaminations have

been reported in Merkén. Therefore, the main goal of this study was to search for the potentially mycotoxigenic

culturable mycobiota on the traditional agriculture production process of the Chilean traditional chilli used for

Merkén’s production and search for mycotoxin contamination in Merkén produced and commercialised in Chile.

Ripened fruit samples of C. annuum from 3 sampling point and Merkén samples were obtained from local farmers

and markets of the South Zone of Chile. Mycobiota from ripened fruit were isolated on MEA, DRBC and DG18

media. Isolated fungal strains were identified using classical morphology and molecular biology techniques. For

the classical taxonomy, macro- and micro-morphology traits where observed. For molecular identification of fungal

strains sequencing of β-tubulin (BenA) and ITS (ITS1-5.8S-ITS2) were performed. In addition, genes associated

with OTA production (Otapks, Otanps, Otachl and Otatra) were tested in one strain of P. crustosum and one strain

of P. verrucosum. For the extraction and clean-up of ochratoxin A (OTA) and aflatoxins (AFLs) B1, B2, G1, G2,

liquid-liquid extraction and immunoaffinity columns were performed. A total of 206 strains were obtained using

classical macro- and micro-morphology. From these fungi, 190 were identified by molecular biology techniques at

the species level. For genes associated with OTA production, P. verrucosum present positive results for Otanps

and Otatra genes, and P. crustosum present positive results for Otatra gene only. For mycotoxins contamination of

Merkén, OTA was presented in 100% samples in a range of 0.79±0.05 and 19.81±0.70 µg/kg and for AFLs, only

AFB1 were detected in 57% samples in a range of 0.29±0.37 and 1.67±0.32 µg/kg. Overall, these results show

that the mycobiota present in C. annuum berry fruits during the process of Merkén production is increasingly

selective for occurrence of Aspergillus and Penicillium species, mainly for the increasingly selective occurrence of

potentially mycotoxigenic fungi such as A. niger, A. flavus, P. expansum and P. verrucosum. These results

demonstrate the importance of knowing even more about the mycobiota and potential mycotoxigenic fungi present

in each stage used for the Merkén production, in order to establish critical control points for a safe and high-quality

product.

KEYWORDS: MICOTOXIGENIC FUNGI, CHILEAN RED PEPPER, MERKÉN, MYCOBIOTA IDENTIFICATION

Page 7: Mycobiota in Chilean chilli Capsicum annuum L. used for

INDEX

Acknowledgments .............................................................................................................................................. I

Resumo ........................................................................................................................................................... III

Abstract ........................................................................................................................................................... IV

1. Introduction .............................................................................................................................................. 1

1.1 Mycobiota and mycotoxins in chilli products ............................................................................................ 2

1.2 Identification of food-borne fungi .............................................................................................................. 6

1.2.1 Conventional microbiological techniques. ......................................................................................... 6

1.2.2 Genetic identification ........................................................................................................................ 7

1.2.3 Genetic identification of ochratoxigenic fungi ................................................................................... 10

1.3 General objective .................................................................................................................................. 13

1.3.1 Specific objectives.......................................................................................................................... 13

2. Materials and methods ............................................................................................................................ 13

2.1 Samples ............................................................................................................................................... 13

2.2 Fungal strains isolation ......................................................................................................................... 14

2.3 DNA extraction ...................................................................................................................................... 15

2.4 Genotypic identification of fungal strains ................................................................................................ 16

2.5 Genetic characterisation of potentially OTA producer fungal strains ........................................................ 17

2.6 Quantitative analyses of OTA and AFLs contamination in Merkén ........................................................... 17

2.6.1 Mycotoxins extraction and clean-up ................................................................................................ 17

2.6.2 OTA determination ......................................................................................................................... 18

2.6.3 AFLs determination ........................................................................................................................ 18

3. RESULTS ................................................................................................................................................ 19

3.1 Fungal isolation and identification .......................................................................................................... 19

3.2 Molecular Identification ......................................................................................................................... 23

3.3 Preliminary genetic characterisation of potentially OTA producer fungal strains ....................................... 30

3.4 Quantitative analyses of OTA and AFLs contamination in Merkén ........................................................... 31

4. Discussion .............................................................................................................................................. 32

5. Conclusions ............................................................................................................................................ 39

6. References .............................................................................................................................................. 40

Annex I – Isolated strains ................................................................................................................................ 50

Annex II – Standard curves of OTA and AFLs.................................................................................................... 55

Page 8: Mycobiota in Chilean chilli Capsicum annuum L. used for

1

1. INTRODUCTION

Capsicum annuum L. is the most well-known species, widespread and cultivated plants of the

genus Capsicum within the family Solanaceae. This species encompasses a wide variety of shapes and

sizes of peppers, both mild and hot, ranging from bell peppers to jalapeños to cayenne peppers. Its berry

fruits are a key element for the creation of endless condiments widely used in gastronomy. According to

the FAO the production area for chillies and peppers is over 1.9 million ha, with a production of 32 million

ton of harvested product per year, with 13.3% produced in America (FAOSTAT, 2017).

In Chile, C. annuum L cv. "Cacho de Cabra” berry fruits are used for the manufacture of a

traditional chilli powder known as Merkén. This is a powder obtained from chilli “Cacho de Cabra” with

a smoky flavour, originally produced by ancestral Amerindian Mapuche communities. These communities

live mainly in the central-south zone of Chile. According to data reported by the National Institute of

Statistics of Chile, in 2016 about 1.5 million of habitants identified themselves as members of the

Mapuche ethnicity (INESTAT, 2016). Although the Mapuche communities are currently disappearing at

an accelerated rate, their cultural and gastronomic influence are of great importance for the identity of

Chile.

Merkén is produced by Mapuche families mainly in the Regions Bio-Bio and La Araucanía and

commercialised on the national and international markets. Production of Merkén can be carried out

mainly in artisan or semi-industrial process. At the artisanal level, the production of Merkén begins with

the harvest of the fruit at the beginning of the civil year, which correspond to the summer season. After

that, the fruit is usually dried by sun exposure or in “Rucas” (typical Mapuche houses made of wood and

straw). After drying, the farmers smoke the fruit, and then crush and powder it and mix it with crushed

coriander seeds and table salt, and some of them add also crushed garlic or other spices. It gives a

darker colour and a smoky flavour to the Merkén that, once processed, generates a very characteristic

chilli. The final product has about 70% of chilli, 20% of coriander seeds and garlic, and 10% sodium

chloride (FIA, 2010).

Due to its high demand, nowadays Merkén is start to be processed industrially to promote the

standardisation of the product and obtain a commercial product of higher quality and local identity. In the

year 2015, total Chilean exports of Merkén to the world reached about 4.4 million US dollars, an

increasing of 11.3% compared to 2014 (ProChile, 2016). However, contamination with mycotoxin, mainly

Page 9: Mycobiota in Chilean chilli Capsicum annuum L. used for

2

ochratoxin A (OTA), have been reported in January 2017 by the Chilean Ministry of Health. It put the

market of this commodity under risk.

The agricultural practices used by Merkén producers are highly empirical, based on ancestral

agriculture practices and do not consider the prevention of fungal growth and further contamination with

mycotoxins. Overall, warm and humid climate in the production areas, the poor hygienic conditions during

transport, packaging and storage, may increase the risk of contamination and fungal proliferation on

peppers (Mandeel, 2005; Schweiggert et al., 2005).

To date, few publications have reported scientific data on chilli’s mycobiota. Only one study by

Ikoma et al. (2015), confirmed the presence of extremely high concentrations of OTA in dried red chilli

peppers from Chile, showing an OTA concentration range of 163.4-1059.2 µg/kg with an average OTA

amount of 355.6 µg/kg. These concentrations are very high and exceed the limits established by the

European Commission for this type of food (15 µg/kg) (EC No. 1137/2015). In addition, a recent risk

profile study published by the Chilean Agency for Quality and Food Safety (by its acronym in Spanish,

ACHIPIA) of the Ministry of Agriculture of Chile, expose the limited information about OTA contamination

and mycobiota in Merkén and Chilli berry fruits from Chile (Di Pillo and Martínez, 2018).

This makes evident the importance of devising strategies to control or reduce the presence of

mycotoxigenic fungi in early and critical stages of the food chain and to establish guidelines to support

farmers in the production of sustainable, safe and high-quality food.

1.1 Mycobiota and mycotoxins in chilli products

Foods are not considered as ecosystems because they contain a large amount of substances

and modifications induced by human, however, foods are systems rich in nutrients and organic matter

that provide an ideal habitat for the proliferation of some microorganisms, such as filamentous fungi

(Santos et al., 1998).

Abiotic factor such as temperature, pH, availability of oxygen and water activity (aw) are perhaps

the most important factors in the fungal growth in food. Temperature is one of the main physical factors

that affect the fungal growth in food and biosynthesis of its secondary metabolites. Although most fungi

are mesophiles (optimal growth between 20-25 °C), some fungi can grow under very high temperatures.

Among the thermotolerant fungi we can find Aspergillus fumigatus with a growth temperatures ranging

about 12 and above 50 °C and with optimal growth temperatures of 40-42 °C. Other thermotolerant

Page 10: Mycobiota in Chilean chilli Capsicum annuum L. used for

3

species of interest in the food industry are Aspergillus flavus and Aspergillus niger, these fungi can grow

at temperatures that vary between 8 to 45 °C (Santos et al., 1998; Bhabhra et al., 2005).

The concentration of oxygen and carbon dioxide in the system are other important factor to be

considered in the growth of food-borne fungi. Most of the fungi can easily grow and germinate up to 4%

of oxygen. Among the most common cases there is Penicillium expansum that can grow at 2.1% of

oxygen. Another outstanding example is P. roqueforti, which can withstand atmospheres of 80% of carbon

dioxide and 4% of oxygen. Some fungi that are able to survive without oxygen such as P. glabrum, P.

corylophilum, Fusarium oxysporium and F. equiseti (Santos et al., 1998; Pitt and Hocking, 2009a).

On the other hand, aw can be defined as the ratio between the partial pressure of water vapour in

foods and the saturation pressure of pure water vapour under the same assay conditions. These concept

is, certainly, one of the most important parameters to consider in food production (Santos et al., 1998).

aw plays a predominant role on growth of filamentous fungi with importance in food industry such as

Penicillium spp., Aspergillus spp. and Fusarium spp. (Frisvad and Samson, 1991; Santos et al., 1998;

Pitt and Hocking, 2009a).

Although drying process contributes to the decrease of the microbial biomass in chilli powder,

many fungi can survive in spores form and germinate when the moisture value exceeds the safety limits

for the consumer (Martin et al., 2005; Dijksterhuis, 2007). On the other hand, the warm and humid

climate in the production areas and the poor hygienic condition during packaging/storage process and

sale areas, may increase the risk of fungal proliferation and contamination (Mandeel, 2005).

Due to the components present in chilli powder manufacturing processes, fungi can fall into two

possible broad ecological categories: (1) Field and (2) Storage. Some studies have found field species

that are generally associated with plants and soil such as Alternaria spp., Cladosporium spp., Mucor spp.

and Rhizopus spp. (Almela et al., 2007; Fatimoh et al., 2017). On the other hand, storage fungi reported

in the literature are mainly related to the genus Aspergillus, Penicillium and Eurotium (Mandeel, 2005;

Almela et al., 2007; Fatimoh et al., 2017).

Due to the aw of these product, xerophilic fungi are the main spoilage microbiota found in chilli

powder. Xerophilic fungi are defined by their ability to grow under very low aw conditions. It means that

fungi have the ability to carry out their life cycles on dried or concentrated substrates, under high

percentages of soluble solids, such as salts or sugars (Pitt and Hocking, 2009b). Moderate xerophilic

fungi and mycotoxigenic fungi such as Aspergillus spp., Penicillium spp. and Fusarium spp. have been

detected among the mycobiota of Capsicum berry fruits and its processed powders. Moreover, aflatoxins

Page 11: Mycobiota in Chilean chilli Capsicum annuum L. used for

4

(AFLs) and OTA have also been found as contaminant in these fresh and processed food (Hierro et al.,

2008; Iqbal et al., 2010; Gherbawy et al., 2015).

Mycotoxins are a broad group of chemical substances produced by some species of fungi that

can cause acute or chronic intoxications leading to illness or death in animals and humans. Exposition to

mycotoxins can occur by ingestion of contaminated foodstuffs, but also by skin contact or inhalation

(Magan and Olsen, 2004).

Due to the aw influences the fungal spore germination and fungal growth, it is certainly a decisive

parameter to produce mycotoxins that could contaminate the final product. For example, Aspergillus

flavus has the minimum aw values for its growth of 0.78-0.80, but is a producer of AFLs when aw reaches

values between 0.83 and 0.87; Aspergillus ochraceus has the minimum aw values for its growth of 0.76-

0.83, but it is a producer of OTA when reaches values of aw between 0.83 and 0.87, and Penicillium

expansum has minimum aw values for its growth of 0.82-0.85, but it is a producer of patulin when reaches

aw of 0.99 (Frisvad and Samson, 1991; Pitt and Hocking, 2009a; Mannaa and Kim, 2017).

The mycotoxins of great significance in chilli powder and chilli fruits include AFLs, OTA,

zearalenone (ZEA), trichothecenes and sterigmatocystin (Santos et al., 2011; Gambacorta et al., 2018;

Hossain et al., 2018). Due to high toxicity, the most important mycotoxin are AFLs followed by OTA and

ZEA. Normally AFLs, OTA and ZEA are produced by fungi in growing crops, while AFLs can also be

produced during storage. Poor hygienic conditions and deficient aw control during processing are main

causes for mycotoxin production in chilli (Santos et al., 2008; Mannaa and Kim, 2017).

AFLs are producers by fungi of the genus Aspergillus section Flavi (especially A. flavus, A. nomius,

A. parasiticus and A. tamari) and can be toxic, carcinogenic, mutagenic and teratogenic to most animal

species, being classified as the most dangerous and extremely carcinogenic mycotoxins (Smith et al.,

1994; Paterson and Lima, 2010b).

OTA is mainly produced by Aspergillus sections Circumdati and Nigri and Penicillium verrucosum

(Almela et al., 2007; Gil-Serna et al., 2009b). Species of Aspergillus section Nigri are among the most

important sources of mycotoxins contamination of foods and feeds, mainly OTA. OTA is toxic to animals

and the International Agency for Research on Cancer classified it as a group 2B carcinogen, i.e., a possible

human carcinogen (IARC, 1993).

ZEA is produced by species of the genus Fusarium and can be classified as a weak estrogen,

non-steroidal, that has mechanisms of competitiveness with the estrogens of the human body activating

and deactivating its metabolic pathways (Smith et al., 1994).

Page 12: Mycobiota in Chilean chilli Capsicum annuum L. used for

5

Mycotoxin determination and quantification in Chilli powder generally consist of five steps: 1)

sampling, 2) sample preparation, 3) extraction, 4) clean-up and 5) quantification (Hu et al., 2006; Santos

et al., 2008; Zhang et al., 2018). The extraction and purification (clean-up) depend on the

physicochemical properties of the sample (Hu et al., 2006; Zhang et al., 2018). The extraction in paprika

and chilli powder is usually done by solvent organic such as methanol, dichloromethane, chloroform,

acetone/hexane, and organic solvent mixed with water or other solvent, such as methanol-water,

acetonitrile-water and benzene-acetic acid (Santos et al., 2010a; Santos et al., 2010b). The extract

purification (clean-up) is necessary because solvents are not selective and can co-extract substances

(pigments and lipids) from the sample (Pal et al., 2005; Hu et al., 2006). Some authors used hexane,

ether or centrifugation to remove the lipids (Pal et al., 2005). However, a more selective treatment can

be selected for clean-up, such as liquid-liquid extraction, solid-phase extraction, strong anion exchange or

immunoaffinity columns (Patel et al., 1996; Pal et al., 2005; Hu et al., 2006; Almela et al., 2007; Zhang

et al., 2018).

The quantification of mycotoxins from paprika and chilli powder include high performance thin-

layer chromatography (HPTLC), high performance liquid chromatography (HPLC), gas chromatography

(GC) and mass spectrometry (MS). Normally AFLs and OTA are analysed by HPLC with fluorescence

detection (Patel et al., 1996; Soares et al., 2010; Abrunhosa et al., 2014). Other methods like Enzyme

Linked Immuno Sorbent Assay (ELISA) are used to detect and quantify mycotoxins and usually do not

need the clean-up step. However, the possibility of false positives because of cross-reaction and

interference in complex matrixes are reported (Hu et al., 2006). Nowadays, due to the low availability of

equipment and high costs associated, the use of HPLC remains the most common.

Mycotoxins contamination of chilli powder has been reported worldwide. The range of

contamination for AFLs has varied between below the limit of detection (<LOD) and 124.6 μg/kg (Santos

et al., 2008), for OTA has varied between 0.4 and 355 μg/kg (El-Kady et al., 1995; Fazekas et al., 2005;

Hierro et al., 2008; Santos et al., 2010c; Ikoma et al., 2015) and for ZEA has varied between <LOD and

131 μg/kg (Patel et al., 1996; Santos et al., 2010b; Motloung et al., 2018).

The European Commission has established regulations for AFLs in Capsicum fruits with

maximum tolerable limits set at 10 µg/kg for total AFLs (AFLB1+AFLB2+AFLG1+AFLG2) and at 5.0 µg/kg

for AFLB1 (EC No. 1881/2006). A new regulation established for the maximum levels of OTA in spices

of 20 μg/kg for Capsicum powder and 15 μg/kg for mixtures of Capsicum with other species (EC No.

1137/2015). According to the European Commission Regulations, the maximum tolerable ZEA

concentration is not yet established for chilli powder.

Page 13: Mycobiota in Chilean chilli Capsicum annuum L. used for

6

1.2 Identification of food-borne fungi

The identification of fungal species has been carried out mainly by conventional microbiological

techniques, such as macro- and microscopic features on different culture media, and DNA-based

methods, such as the amplification and sequencing of different DNA barcodes.

1.2.1 Conventional microbiological techniques.

Over the past two centuries, the identification of different filamentous fungi were performed by

classical microbiological techniques. These methods allow the differentiation according morphological

criteria and microscopic features in accordance with appropriate dichotomous keys (Pitt et al., 1993).

At the time of performing a conventional identification, a series of parameters must be taken into

consideration, such as the macroscopic characteristics (e.g., colony colour and cultural characteristics)

and microscopic (e.g., spore size and shape) phenotypic characteristics observed from cultures grown

on different media (Pitt et al., 1993; Ryu et al., 2015).

When the identification of filamentous fungi is based on morphological features the use of

standard culture media is important because the morphology and colonies colour vary from medium to

medium. In the case of mycological analyses in foods, the aw is a fundamental parameter for successful

fungal growth. Many of the culture media found in the literature for fungal isolation from chilli powder are

formulated considering its the aw (Pitt et al., 1993; Mandeel, 2005; Martin et al., 2005; Santos et al.,

2008; Santos et al., 2011).

Selective fungal enumeration in chilli powder has been achieved using oxytetracycline-glucose

yeast extract agar (OGYEA) incubated at 25 ºC for 5 days (Almela et al., 2007), yeast extract glucose

chloramphenicol agar (YGC agar) at 25 ºC for 5 days and potato dextrose agar (PDA) at 28 ºC for 2-5

days (Schweiggert et al., 2005). In the case of food-borne fungi, Dicloran Rose Bengal Chloramphenicol

Agar (DRBC Agar) (King et al., 1979; Mandeel, 2005), Dicloran 18% Glycerol Agar (DG18) (Hocking and

Pitt, 1980), Czapex-Dox Agar (CD) (Thom and Church, 1926; Mandeel, 2005), Aspergillus

Flavus/Parasiticus Agar (AFPA) (Pitt et al., 1983), Corn Yeast Agar (CYA) and Malt Extract Agar (MEA)

(Thom and Church, 1926; Mandeel, 2005; Martin et al., 2005) are the most commonly used media.

Usually, for the study of mycobiota, these media are incubated at temperature between 21 and

28 ºC for 5-15 days. In the most of cases, due to it reduces interferences from both bacteria and rapidly

growing fungi, DG18 is the medium of choice for the isolation of moderate xerophilic fungi from food with

low aw (0.955) (Hocking and Pitt, 1980). However, in some cases it is necessary to use culture media

Page 14: Mycobiota in Chilean chilli Capsicum annuum L. used for

7

with very low aw such as Malt Extract 50% Glucose Agar (MY50G) with aw of 0.89 or Malt Extract Yeast

Extract 70% Glucose Fructose Agar (MY70GF) with aw of 0.75 (Ryu and Wolf-Hall, 2015).

After incubation, the microorganisms are visualised by light field microscopy using different

stains. Most of the dyes used in mycology are slow acting because the fungal cell wall is highly resistant

to staining due to their composition. The most used dyes in mycology include lactofucsins, lactophenol

blue, cotton blue and erythrosine in ammonia (Pitt and Hocking, 1992).

Although the identification by conventional techniques is fundamental for the identification of

filamentous fungi, they are time consuming, subjective and usually requires taxonomic expertise. For this

reason, more efficient and non-subjective methods have been used in recent years, such as DNA-based

methods (Frisvad and Samson, 2004; Geiser et al., 2007; Rodrigues et al., 2009; Simões et al., 2013;

Silva et al., 2015; Decontardi et al., 2018).

1.2.2 Genetic identification

With the advance of molecular biology in the last quarter of XX century, cutting-edge technologies

have been developed for the identification of microorganisms based on their genomic DNA. The DNA-

based techniques permit rapid, sensitive and specific detection, of filamentous fungi in food samples and

are faster than conventional methods (Frisvad and Samson, 2004; Geiser et al., 2007; Rico-Muñoz et al.,

2018).

The identification of filamentous fungi from chilli powder has been achieved using PCR-based

methods (Santos et al., 2010a; Santos et al., 2010b). These methods allow the specific detection of small

amounts of target organisms by amplifying their DNA in a considerably short time frame. The most

common used methods are conventional and real-time PCR assays, but next generation DNA sequencing

is becoming more common (Rodrigues et al., 2011; Silva et al., 2015; Decontardi et al., 2017; Decontardi

et al., 2018).

In this context, the extraction of high quality and quantity DNA from fungi is one of the most

important parameters to consider. The DNA extraction protocol has a pronounced effect on the result of

a molecular study and must be optimised for each food product or in some cases for each kind of fungi.

On the other hand, some studies have shown that different methods of DNA extraction, as well as different

purities or qualities can affect the interpretation of results (Costea et al., 2017).

The DNA extraction of fungi in culture media can present in some cases great difficulties due to

the cellular structure of each fungus. Fungi can produce unicellular structures such as conidia in the case

of Aspergillus and Penicillium or multicellular structures such as Alternaria and Fusarium. These

Page 15: Mycobiota in Chilean chilli Capsicum annuum L. used for

8

structures can be thin-walled, but can also be thick-walled with large amounts of melanin (melanised

walls). Certain species such as Cladosporium and Alternaria (Dothideomycetes) produce resistant and

melanised cell walls and the rupture of these cells is difficult. It is expected that DNA recovery will be

different for each cell type (Summerbell, 2011).

In the case of DNA purity, this can be very important at the time of using PCR-based methods, if

the DNA is used in a PCR based assay, then compounds in the DNA extract can affect or inhibit the PCR

reaction by binding or denaturing the Taq DNA polymerase (Hayat et al., 2012)

Methods based on PCR techniques have a theoretic detection limit down to 1 to 10 molecules of

DNA. However, the concentration, quality, purity and PCR conditions (melting temperature, hybridization

time and elongation time) substantially affect its performance and its veracity. However, internal

amplification controls can be used to determine the presence of inhibitory substances or failure of a PCR

(Hoorfar et al., 2004).

In the last years, different target housekeeping genes have been used for the identification of

fungi. These are usually divided into two groups: 1) single-copy and 2) multicopy. Single-copy gene

analyses (for example, genes encoding actin, calmodulin, chitin synthetase, heat shock proteins 90 or

superoxide dismutase) are highly species-specific; however, since they are single-copy, they may not be

detected easily. In contrast, the amplification of multicopy genes considerably increases the sensitivity

due to a greater number of copies of the target gene in the sample (Mitchell et al., 1995; Chen et al.,

2002; Cabañes, 2002).

Although the single-copy genes are more species-specific, the most used target genes in the last

years are the multicopy genes found in the complex of ribosomal DNA complex (rDNA). Due to the fact

that ribosomes are responsible for protein synthesis, these genes have been conserved in all life forms,

presenting small alterations generated by evolutionary processes. These genes are excellent genetic

markers for the identification of fungi (Mitchell et al., 1995) and recommended as universal DNA barcode

marker for fungal identification (Schoch et al., 2012).

In fungi, the rDNA is a tandem formation of 40-240 copies per genome in the haploid genome

of these microorganisms (O´Donnel, 1992). It comprises the coding regions of the small subunit of the

ribosome (18S), the 5.8S gene and the large subunit of the ribosome (28S). Separating the subunits 18S-

5.8S and 5.8S-28S are non-coding regions Internal Transcribed Spacer (ITS) (Chen et al., 2002; Iwen et

al., 2002).

The ITS rDNA regions have been accepted as the official DNA barcode for fungi and is used as

universal region to identification of fungal species (Hebert et al., 2003; Geiser et al., 2007; Schoch et al.,

Page 16: Mycobiota in Chilean chilli Capsicum annuum L. used for

9

2012). In the case of chilli powder the identification of A. flavus, A. niger, A. ochraceus, A. tubingensis

and A. westerdijkiae have been archived amplifying the region ITS1-5.8S-ITS2 (Gil-Serna, 2009b; Santos

et al., 2010a; Santos et al., 2010b). Also, the primers 5.8S1/5.8S2 have been using to amplify the 5.8S

region to identify A. carbonarius (Patiño et al., 2005), A. steynii (Gil-Serna, 2009a), A. flavus and A.

parasiticus (González-Salgado et al., 2008).

Although the ITS region is considered as a universal barcoding, the resolution of this region can

be insufficient for species level identification especially for fungi of food origin as such as Aspergillus,

Cladosporium, Fusarium and Penicillium. For this reason, sequencing of a secondary barcode is

recommended (Peterson, 2008; Roe et al., 2010; Varga et al., 2011; Peterson, 2012; Pérez‐Izquierdo

et al., 2017).

Other barcoding genes such as β-tubulin (BenA), Calmodulin (CaM) and translation elongation

factor-1 alpha (TEF-1α) have been used in many fungal identification and classification studies in recent

years (Perrone et al., 2011; Rodrigues et al., 2011; Varga et al., 2011; Soares et al., 2012; Frisvad et

al., 2013; Visagie et al., 2014; Decontardi et al., 2017; Decontardi et al., 2018).

For Aspergillus and Penicillium species, the most commonly used genes are BenA and CaM. In

the case of the tubulin superfamily, only α-, β- and γ-tubulins have homologues in fungal genomes

(Dutcher, 2001). The β-tubulin paralogs are represented in Aspergillus by two genes called BenA and

tubC. The BenA gene encodes polypeptides designated as β1- and β2-tubulin. This gene is an excellent

molecular marker for these fungi and is currently the third most used gene in fungal multilocus

phylogenies (Feau et al., 2011).

Although the genes to be amplified and sequenced must be sufficiently distinctive to identify

genetically related fungi, the use of the databases is another important point to consider. The results of

sequencing and identification depend strongly on the quality and quantity of information deposited in the

databases. In relation to the GenBank database, it is estimated that about 20% of fungal identifications

deposited have errors or are unreliable (Nilsson et al., 2006; Bidartondo et al., 2008).

Nowadays, different analytical techniques allow the identification of microorganisms with the use

of biomolecules as fingerprint. The Matrix-Assisted Laser Desorption Ionization-Time-Of-Flight Mass

Spectrometry (MALDI-TOF MS) technique can be used for fungal identification. MALDI-TOF MS has a short

turnaround time, is robust and can be reliable. The technique measures the difference in size of

biomolecules such as proteins, peptides, glycoproteins, biomarkers, sugars and also organic molecules.

The use of MALDI-TOF MS for the identification of food spoilage filamentous fungi has successfully been

Page 17: Mycobiota in Chilean chilli Capsicum annuum L. used for

10

explored (Santos et al., 2010d; Rodrigues et al., 2011; Chaves da Silva et al., 2015; Santos et al., 2015;

Chang et al., 2016; da Cruz Pedrozo Miguel et al., 2017; Lima and Santos, 2017).

Fourier transform infrared spectroscopy (FT-IR) is another powerful analytical technique that can

be applied for characterising the chemical composition of very complex probes such as filamentous fungi.

This technique has been successfully applied for the identification of filamentous fungi. Fungal FT-IR

typing is fast, effective and reagent-free. Moreover, it requires a small quantity of biomass. When FT-IR is

applied on the fungal cell, it generates a unique ‘‘fingerprint’’ representing the total cell chemical

composition such as lipids, proteins, nucleic acids and polysaccharides. FT-IR spectra are highly specific

for each fungal strain and species (Santos et al., 2010e).

For this reason, the most accurate approach to get a precise and reliable identification is the

polyphasic identification, merging the morphological, chemical and genetic trait results in the final

identification (Frisvad and Samson, 2004; Geiser et al., 2007; Rodrigues et al., 2009; Simões et al.,

2013; Silva et al., 2015; Decontardi et al., 2018).

1.2.3 Genetic identification of ochratoxigenic fungi

The early and rapid detection of possible mycotoxin-producing fungi is important to reduce the

negative impacts on food commodity. In recent years, PCR systems have been developed for the detection

and differentiation of the main species and groups of mycotoxigenic fungi, by means of the amplification

of sequences that code for associated enzymes in the metabolic pathways of production of these

mycotoxins. These strategies have been successfully applied to aflatoxin-producing fungi (Geisen, 1996;

Shapira et al., 1996; Rodrigues et al., 2009), OTA (Geisen et al., 2006), fumonisins (Paterson, 2006a;

Gonzalez-Jaen et al., 2004) and also to patulin producers (Paterson, 2006a; Paterson, 2006b).

The ocratoxina A (OTA) is a mycotoxin produced by the secondary metabolism of many

filamentous species. Biosynthetically, it is a pentaketide derived from the dihydrocoumarins family

coupled to β-phenylalanine (El-Khoury and Atoui, 2010). Although there is much information about the

various toxigenic properties of OTA, unlike other important mycotoxins, not much is known about the

pathway of OTA biosynthesis. It is widely believed that the isocoumarin group is a pentaketide formed

from acetate and malonate through a polyketide synthesis route (Niessen et al., 2005).

In 1979 Huff and Hamilton proposed a biosynthetic route based on a mechanistic model

according to the structure of the OTA (Figure 1). In this model, a polyketide synthase (PKS) is considered

as a key enzyme in the biosynthesis of OTA in a similar way to other polyketide mycotoxins such as

fumonisins (Proctor et al., 1999) and aflatoxins (Bhatnagar et al., 2003).

Page 18: Mycobiota in Chilean chilli Capsicum annuum L. used for

11

According to Huff and Hamilton (1979), OTA biosynthesis can be summarized in three steps: the

first step involves the synthesis of polyketides of ochratoxin α via Mellein using a polyketide synthase.

The second step includes addition of the acyl group to Mellein, oxidizing the molecule to 7-carboxy-mellein

and forming OTβ. Followed by this chlorination by a chloroperoxidase leads to OTα. OTα then transforms

into a mixed anhydride, an activation reaction that uses ATP. The second precursor phenylalanine is

synthesized through the metabolic pathway of shikimic acid, followed by activation of the ethyl ester so

that it can participate in the subsequent acyl shift reaction. In the third step, the binding of these activated

precursors through a synthetase, generating OTC, an ethyl ester of OTA: the de-esterification by an

esterase or transesterification is the last step in this hypothetical biosynthetic route (Figure 1).

Figure 1. Schematic representation of the hypothetical OTA biosynthetic

pathway. Adapted from Huff and Hamilton (1979).

OTβ

OTα

OTC

OTA

Page 19: Mycobiota in Chilean chilli Capsicum annuum L. used for

12

In 2006, Geisen et al. identified in Penicillium nordicum fragments of genes highly related to the

production of OTA, which are related to the metabolic pathway proposed by Huff and Hamilton (1979).

Two fragments were identified (Figure 2), a fragment of 10 kb DNA, containing three genes: A gene with

high homology with an alkaline serine protease gene (aspPN), representing the upper edge of the cluster.

A fragment carrying a large part (approximately 2 kb) of the 5' end of a polyketide synthase (otapksPN,

≈500 bp for P. nordicum) and a complete non-ribosomal peptide synthetase (otanpsPN, ≈750 bp for P.

nordicum). The second 4.5 kb fragment harboring three open reading frames encoding putative OTA

biosynthetic proteins: an incomplete open reading frames at the 5' end of the fragment demonstrated

homology with an organic anionic transporter of rat kidneys (otatraPN, ≈400 bp for P. nordicum). A

complete open reading frames of 951 nucleotides that has a limited homology with a chloroperoxidase

(otachlPN) of Gluconobacter oxidans. At the 3' end of this DNA fragment there is an incomplete open

reading frame of a possible nitrate transporter (ntraPN) (Geisen et al., 2006).

Figure 2. Map of the characterised OTA biosynthetic genes. aspPN = alkaline serine protease; otapksPN = ochratoxin polyketide synthase; otanpsPN = ochratoxin non-ribosomal peptide synthetase; otachlPN = ochratoxin

chloroperoxidases; otatraPN = ochratoxin transport protein and ntraPN= nitrate transporter. Adapted from Geisen et al. (2006).

Page 20: Mycobiota in Chilean chilli Capsicum annuum L. used for

13

1.3 General objective

The main aim of the present work is search for the potentially mycotoxigenic culturable mycobiota

on the traditional agriculture production process of the Chilean traditional chilli Capsicum annuum L. cv.

"Cacho de Cabra” used for Merkén’s production and search for mycotoxin contamination in Merkén

produced and commercialised in Chile.

1.3.1 Specific objectives

1. Isolate the culturable mycobiota on three time points of the traditional agriculture production process

of the Chilean traditional chilli Capsicum annuum L. cv. "Cacho de Cabra” used for Merkén’s production;

2. Identify the isolated fungal strains based on classical morphology and molecular biology techniques;

3. Characterise the genes associated with OTA production for potentially ochratoxigenic strains;

4. Search for OTA and AFLs contamination in Merkén produced and commercialised in Chile.

2. MATERIALS AND METHODS

2.1 Samples

Samples of chilli C. annuum were provided by 8 farmers belonging to the follow localities of the

Chilean Region of La Araucanía (Figure 3): Nueva Imperial (2 farmers), Hualacura (1 farmer), Cholchol

(1 farmer) and Purén (4 farmers). Samples were collected at 3 different sampling points of berry fruits

production: (1) just at the day of ripe fruits harvest; (2) after 1 month of harvest (drying process); and (3)

during the smoking process. A total of 240 samples of chilli berry fruits (8 farmers, 10 berry fruits per

farmer distributed in 3 different sampling points of berry fruits production) were collected. Berry fruits

samples were used for mycobiota isolation. In addition, 10 g of Merkén samples from each farmer and

from 13 different regional markets were collected in order to determine OTA and AFLs contamination. All

Merkén samples were stored at 4 °C until extraction and quantification.

Page 21: Mycobiota in Chilean chilli Capsicum annuum L. used for

14

2.2 Fungal strains isolation

Isolation of fungal strains was carried out at the Laboratory of the Chilean Culture Collection of

Type Strains (CCCT/UFRO, http://ccct.ufro.cl/), which is hosted at the BIOREN-UFRO Scientific and

Technological Bioresource Nucleus (http://bioren.ufro.cl/) of the Universidad de La Frontera, Temuco,

Chile.

In order to isolate the mycobiota, each fruit was cut and placed on MEA (malt extract 20g/L,

mycological peptone 1g/L, agar 20g/L, glucose 20 g/L), DRBC (KH2PO4 1g/L, MgSO4·7H2O 0.5 g/L,

Figure 3. Location of the sampling regions in the Region of La Araucanía, Chile. (Adapted from Biblioteca Nacional del Congreso de Chile, URL:www.bcn.cl/siit/mapas_vectoriales/mapoteca, access on

15/06/2018).

Page 22: Mycobiota in Chilean chilli Capsicum annuum L. used for

15

mycological peptone 5g/L, dicloran 0.002g/L, chloramphenicol 0.1 g/L, agar 15g/L, glucose 10 g/L,

rose bengal 0.025 g/L) and DG18 (peptone 5g/L, glucose 10 g/L; KH2PO4 1g/L, MgSO4·7H2O 0.5 g/L,

dicloran 0.002 g/L, agar 15g/L) in Petri dishes. Each plate was incubated for 7 days at 27 °C in the

dark. After isolation, the fungal strains were identified at genus level based on macro- and micro-

morphological traits. Classical fungal identification followed the taxonomic keys and guides available for

each specific genus.

All fungal strains were then sent to the Laboratory of Mycology of the Center of Biological

Engineering of the University of Minho (Braga, Portugal, www.ceb.uminho.pt/amg) for molecular biology

identification. All fungal strains were deposited at CCCT/UFRO and at the Micoteca da Universidade do

Minho (MUM, www.micoteca.deb.uminho.pt/). In this latter case, the international regulations were

compiled and a Material Transfer Agreement was established.

2.3 DNA extraction

Genomic DNA of each isolates was extracted using a modified protocol describe by Rodrigues et

al. (2009). Briefly, spores of each strain were transferred from a 7 days old culture into 50 mL falcon

tubes containing 25 mL of Malt Extract-Glucose-Yeast Extract-Peptone Medium (MGYP, malt extract 3

g/L, glucose 10 g/L, yeast extract 3 g/L, peptone 5g/L). Samples were incubated at room temperature

for 5 days in the dark, at 150 rpm in a shaker.

Fungal biomasses were filtrated and stored at -20 °C. For DNA extraction, 100 mg of biomasses

were transferred into a 1.5 mL Eppendorf tube containing 100 µL of lysis buffer (200 mM Tris-HCl pH

8.5, 250 mM NaCl, 25 mM EDTA, 0.5% (w/v) SDS). Cells lyses were performed using a pellet pestle for

3-4 min. After mechanical lysis, 900 µL of lysis buffer was added and the samples were incubated for 1

hour at 65 °C. Samples were centrifuged at 14000 x g for 10 min at room temperature and 800 µL of

upper phase was transferred into a new 2 mL Eppendorf tube.

Polysaccharides and proteins were precipitated by adding 1 mL of cold sodium acetate (3 M, pH

5.5). Samples were gently mixed by inversion, placed at -20 °C for 10 min and centrifuged at 14000 x g

for 10 min at room temperature. Clean supernatant was then transferred to a new tube and precipitated

with one volume of cold isopropanol (-20 °C). Samples were gently mixed by inversion for 2 minutes,

incubated at −20 °C for 2 h and centrifuged at 14000 x g for 10 min. DNA pellets were washed twice

with 1.0 mL of cold 70% ethanol, centrifuged at 14000 x g for 10 min and dried using a SpeedVac

Concentrator. DNA samples were suspended on 50 µL of ultra-pure water and stored at -20 °C. DNA

Page 23: Mycobiota in Chilean chilli Capsicum annuum L. used for

16

samples were subjected to quality assessment by quantification of total DNA using Thermo Scientific™

NanoDrop™ instrument and by electrophoresis agarose gel 1% (w/v) for 45 min at 80 V. Electrophoresis

agarose gel SYBR® Safe DNA Gel Stain (Invitrogen) was used as a staining element and NZYDNA ladder

III was used as a DNA molecular weight marker.

2.4 Genotypic identification of fungal strains

In order to identify the fungal strains, amplification of BenA and ITS genes were performed.

The BenA gene was amplified using the primers Bt2a (5´-GGT AAC CAA ATC GGT GCT GCT TTC-

3´) and Bt2b (5´-ACC CTC AGT GTA GTG ACC CTT GGC-3´) design by Glass and Donaldson (1995). For

BenA reactions used 25 µL Taq DNA polymerase Master Mix 2x (VWR Life Science), 2 µL primer Bt2a

10 mM, 2 µL primer Bt2b 10 mM, 2 µL of total DNA and 19 µL of distilled water free of proteases to a

final volume of 50 µL were used. PCR parameters used in the thermal cycler were: 95 °C for 3 min, 35

cycles of 95 °C for 1 min, 56 °C for 45 s, 72 °C for 90 s and a final extension at 72 °C for 10 min.

The ITS1-5.8S-ITS2 rDNA region was amplified using the primers ITS1 (5´-TCC GTA GGT GAA

CCT GCG G-3´) and ITS4 (5´-TCC TCC GCT TAT TGA TAT GCC-3´) design by White et al. (1990). For

ITS reaction 25 µL Taq DNA polymerase Master Mix 2x (VWR Life Science), 2 µL primer ITS1 10 mM, 2

µL primer ITS4 10 mM, 2 µL of Total DNA and 19 µL of distilled water free of proteases to a final volume

of 50 µL were used. PCR parameters used in the thermal cycler were: 94 °C for 3 min, 35 cycles of 94

°C for 1 min, 55 °C for 1 min, 72 °C for 1 min and a final extension at 72 °C for 5 min.

Bands were visualized on 1% (w/v) agarose gel supplemented with SYBR® Safe DNA Gel Stain

(Invitrogen) as a staining element and NZYDNA ladder III as a DNA molecular weight marker. PCR

products were cleaned using NZYGelpure kit according to the manufacturer’s instruction and send to Stab

Vida Oporto Laboratories (Centro de Testagem Molecular, Vairão, Portugal;

http://www.stabvida.com/pt), sending 3 µL of the forward primer and 10 µL of genomic DNA. Gene

sequences were provided in AB1 format file by Stabvida Laboratories.

Every sequence was opened into the analysis software package Mega7 (Molecular Evolutionary

Genetics Analysis version 7.0 for bigger datasets; Kumar et al., 2016) and the quality of the sequences

was verified using as reference the quality graphics sent by Stab Vida. Each sequence was compared

with the GenBank database sequences using the Basic Local Alignment Search Tool (BLAST

https://blast.ncbi.nlm.nih.gov/). Sequences were aligned into the analysis software package Mega7

using MUSCLE alignment (Robert, 2004). Dendrograms were then deduced, opening the sequences into

Page 24: Mycobiota in Chilean chilli Capsicum annuum L. used for

17

the pattern analysis software package Mega7 and using the Neighbour-Joining method (Saitou and Nei,

1987). All positions containing gaps and missing data were eliminated. The evolutionary distances were

computed using the Kimura 2-parameter method (Kimura, 1980). All sequences of the reference strains,

also included for clustering, were obtained by searching on the website of the fungal collection CBS-KNAW

(http://www.westerdijkinstitute.nl/collections/); also GenBank database

(http://www.ncbi.nlm.nih.gov/genbank/).

2.5 Genetic characterisation of potentially OTA producer fungal strains

In order to identify, preliminarily, a putative OTA biosynthetic genes of Penicillium species.

Genomic DNA of P. verrucosum (CCCT 18.33) and P. crustosum (CCCT 18.46) were subjected to PCR

using the primers, Otapks1 (5´-TAC GGC CAT CTT GAG CAA CGG CAC TGC C-3´) Otapks2 (5´-ATG CCT

TTC TGG GTC CGA TA-3´), Otanps_for (5´-AGT CTT CGC TGG GTG CTT CC-3`), Otanps_rev (5´-CAG

CAC TTT TCC CTC CAT CTA TCC- 3`), Otachl_for (5´-CGT GAT GGC CAG ATG GGC GTG-3`), Otachl_rev

(5´-CCG TCT CTC CAT TCT CTT CCT-3`), Otatra_for (5´-GGT CGG GCC GAT GTT TGA TCG-3`) and

Otatra_rev (5´-CCT CGC ATC TTG TAA GGA ACG C-3`) design by Geisen et al. (2006). Such internal

amplification control (IAC), ITS1-5.8S-ITS2 region using ITS1/ITS4 primers were used. Penicillium

nordicum (CBS 112573) was used like positive control and A. carbonarius (MUM 01.08) was used like

negative control.

For PCR reaction 25 µL Taq DNA polymerase Master Mix 2x (VWR Life Science), 2 µL of each

primer 10 mM, 2 µL of Total DNA and 16 µL of distillate water free of proteases to a final volume of 50

µL. The temperature ramps used in the thermal cycler was 95 °C for 3 min, 33 cycles of 95 °C for 30

min, 60 °C for 40 s, 72 °C for 60 s and a final extension at 72 °C for 10 min. The Samples will be

visualized on 1% agarose gel supplemented with Green Safe Premium (Nzytech MB13201) as a DNA

identifier. Five μL Ladder III Molecular Weight Marker and 5 μL of each amplified sample will be loaded

and the gel run at 75 V for 30 min.

2.6 Quantitative analyses of OTA and AFLs contamination in Merkén

2.6.1 Mycotoxins extraction and clean-up

Page 25: Mycobiota in Chilean chilli Capsicum annuum L. used for

18

Extraction and clean-up of AFLs and OTA from Merkén samples using liquid-liquid extraction and

immunoaffinity columns were performed with two independent replicates. For the extraction, 2 g of

Merkén were mixed with extraction solution (0.2 g of NaCl; 10 ml of methanol:water 8:2; and 5 mL of

hexane) in Erlenmeyer flasks of 100 mL and shake on a mechanical shaker at 150 rpm for 1 hour at

room temperature. The solution was filtered through WhatmanTM Nº 4 filter paper and separate using a

separatory funnel. After phase separation, 5 mL of the aqueous layer were diluted with 30.7 mL of

Phosphate Buffered Saline (PBS) buffer (137 mM NaCl; 2.7 mM KCl; 10 mM Na2HPO4; 2 mM KH2PO4;

pH 7.4) and filtered using glass microfibre filters WhatmanTM. For the clean-up, 10 mL of solution extracted

previously was added to AflaOchra HPLC immunoaffinity columns VICAM column at flow rate of 2

mL/min. the column was washed with 10 mL of PBS buffer at flow rate of 2mL/min and finally the

mycotoxins were eluted with 1.5 mL of methanol. The eluted samples were quantified by HPLC.

2.6.2 OTA determination

Ochratoxin A quantification was performed according to Abrunhosa et al. (2014) using High

Performance Liquid Chromatography (Waters, Milford, MA, USA) with a reverse-phase C18 silica gel

column (250x4.6 mm, 5 μm), equipped with a Varian 9002 pump (Agilent, Palo Alto, CA, USA), a Varian

Prostar 410 autosampler and Jasco FP-920 fluorescence detector (Jasco Europe, Cremella, Italy).

Excitation and emission wavelengths was set at 333 and 460 nm, respectively. An isocratic mobile phase

of acetonitrile/water/acetic acid (99:99:2, v/v/v) was used with a flow rate of 1.0 mL/min. OTA was

identified by comparing the retention time of the peak samples with the standards. OTA determination in

samples was based on the external standard calibration method, using an OTA concentration range of

0.05–100 ppb. Recovery rates was calculated using Merkén samples spiked with 50 and 100 ppb of

OTA.

2.6.3 AFLs determination

Aflatoxins quantification was performed according to Soares et al. (2010) using High

Performance Liquid Chromatography (Waters, Milford, MA, USA) with a reverse-phase C18 silica gel

column (4.6x250 mm, 5 µm), equipped with a Varian 9002 pump (Agilent, Palo Alto, CA, USA), a Varian

Prostar 410 autosampler and Jasco FP-920 fluorescence detector (Jasco Europe, Cremella, Italy).

Excitation and emission wavelengths was set at 365 and 435 nm, respectively. An isocratic mobile phase

of water/acetonitrile/methanol (3:1:1, v/v/v) was used with a flow rate of 1.0 mL/min. AFLs were

Page 26: Mycobiota in Chilean chilli Capsicum annuum L. used for

19

identified by comparing the retention time of the peak samples with the standards. AFLs determination

in samples was based on the external standard calibration method, using an AFLs concentration range

of 0.1–20 ppb. Recovery rates was calculated using Merkén samples spiked with 50 and 100 ppb of

each AFB1, AFB2, AFG1 and AFG2.

3. RESULTS

3.1 Fungal isolation and identification

A total of 206 fungal strains belonging to 9 fungal genera were identified using classical macro-

and micro-morphology. The strains were isolated from berry fruits of C. annuum used for the manufacture

of Merkén in all sampling points. From the total isolated fungal strains, 86 strains (41.7%) were isolated

on DG18, 80 strains (38.8 %) were isolated on DRBC and 40 strains (19.4 %) were isolated on MEA.

From total fungi, 190 strains (92.2 %) were identified by molecular biology techniques at the species level.

All the fungal identified strains were deposited at both CCCT/UFRO and MUM culture collections with

their respective unique strain identification numbers (Annex I, Table A).

From the sampling point I (just at the day of ripe fruits harvest), 61 fungal strains were identified

(Figure 4). From these, strains were distributed among Penicillium spp. (25), Fusarium spp. (14),

Alternaria spp. (9), Aspergillus spp. (7) and others (6). In the case of Penicillia, the species with greatest

abundance were P. veridicatum (6; 24%), P. brevicompactum (5; 20%) and P. crysogenum (4; 16%).

Whereas for fusaria the species with greatest abundance were F. oxysporum (5; 38%) followed by F.

esquiseti (4; 29%).

For Alternaria, A. alternata (8; 89%) represents the most abundant species. Aspergilli represents

a low amount of fungi in these sampling point being A. versicolor (2; 29%), A. dimorphicus (2; 29%) and

A. fructus (2; 29%) the most representative. The strains that did not belong to these major taxa were

identified and grouped as “Other species”, which were, Cladosporium westerdijkieae (3; 50%),

Trichoderma koningiopsis (2; 33%) and Boeremia exigua (1; 17%).

From the sampling point II (drying process), 46 fungal strains were identified (Figure 5). These

strains were distributed among: Penicillium spp. (24), Aspergillus spp. (12), Alternaria spp. (7) and

“Others species” (3). In this sampling point, Penicillia were the most predominant species, with greatest

abundance for P. crustosum (8; 33%) followed by P. glabrum (6; 25%). For Aspergilli, A. niger (6; 50%)

and A. flavus (3; 25%) were the most representative ones. For Alternaria, A. alternata (5; 71%) represents

Page 27: Mycobiota in Chilean chilli Capsicum annuum L. used for

20

the most abundant species. Cladosporium westerdijkieae (1; 33%), Trichurus spiralis (1; 33%) and

Colletotrichum coccodes (1; 33%) were also identified and grouped as “Others species”.

Figure 4. Total of fungal strains isolated in the sampling point I (just at the day of ripe fruits harvest). Number and percentage of strains: (A) at genus level; (B) Penicillium; (C)

Aspergillus; (D) Fusarium; (E) Alternaria and (F) Other species.

Page 28: Mycobiota in Chilean chilli Capsicum annuum L. used for

21

From the sampling point III (during the smoking process), 83 fungal strains were identified (Figure

6). These strains were distributed among: Penicillium spp. (63), Aspergillus spp. (18), and “Others

species” (2). In these case, Penicillium spp. and Aspergillus spp. were the most dominants ones. For

Penicillia, P. glabrum (13; 21%), P. brevicompactum (13; 21%), P. cyclopium (8; 13 %) and P. crustosum

Figure 5. Total of fungal strains isolated in the sampling II (drying process). Number and percentage of strains: (A) at genus level; (B) Penicillium; (C) Aspergillus; (D) Alternaria and (E) Other species.

Page 29: Mycobiota in Chilean chilli Capsicum annuum L. used for

22

(7; 11%) were the most abundant ones. For Aspergilli, A. niger (5; 28%), A. awamori (3; 17%), A. flavus

(2; 11%), A. fumigatus (2; 11%) and A. pseudoglaucus (2; 11%) were the most abundant ones.

Trichoderma viride (1; 50%) and Microascus cinereus (1; 50%) were identified and grouped as “Others

species”. In this sampling point, Alternaria and Fusarium are not detected.

Figure 6. Total of fungal strains isolated in the sampling III (during the smoking process). Number and

percentage of strains: (A) at genus level; (B) Penicillium; (C) Aspergillus; and (D) Other species.

Page 30: Mycobiota in Chilean chilli Capsicum annuum L. used for

23

From all sampling points, Alternaria spp., Aspergillus spp., Boeremia spp., Cladosporium spp.,

Colletotrichum spp., Fusarium spp., Microascus spp., Penicillium spp., Trichoderma spp. and Trichulus

spp. were identify. All genera found in the all sampling point were group in a Venn diagram and analysed

using a Heat map (Figure 7).

The most representative genera were Aspergillus and Penicillium, being representatives of both

species in the three sampling points. From a point of view of relative abundance Penicillium, Aspergillus

and Alternaria were the most abundant respectively.

From all sampling points, 49 species were identify using DNA sequence. All species found in

were group in a Venn diagram and analysed using a Heat map (Figure 8).

In the Venn diagram it is observed that A. flavus, P. brevicompactum, P. cyclopium, P. glabrum,

P. polonicum and P. viridicatum were identified in the three sampling points. From a point of view of

relative abundance P. glabrum, P. brevicompactum and P. crustosum were the most abundant

respectively.

Figure 7. Venn diagram of all genus found in all sampling points (A) and Heat map showing the relative abundance of all fungi at genus level found in all sampling points (B)

Page 31: Mycobiota in Chilean chilli Capsicum annuum L. used for

24

3.2 Molecular Identification

A total of 190 fungi from all sampling point were analysed by molecular biology. From these, 136

strains were sequenced and identified using BenA amplicons and 54 strains using ITS1-5.8S-ITS2. From

BenA sequences, 32 sequences corresponded to Aspergillus species and 93 correspond to Penicillium

species. For the construction of the phylogenetic trees, only BenA sequences from Aspergillus and

Penicillium species were used.

Robustness of Aspergillus species identification was added to the cluster by the addition of BenA

nucleotide sequences of A. awamori (NRRL 35710; NRRL 4760; CBS 55765; GL 125), A. niger (CBS

55465; KACC 46497; KACC 46495; CBS 101699; CBS 101698), A. tubingensis (CBS 13448), A. flavus

(CBS 10092), A. fructus (NRRL 239), A. pseudoglaucus (CBS 379.75), A. fumigatus (CBS 112389; NRRL

A B

Figure 8. Venn diagram of all species identified in all sampling points (A) and Heat map showing the relative abundance of all fungi at species level found in all sampling points (B).

Page 32: Mycobiota in Chilean chilli Capsicum annuum L. used for

25

132) from GenBank database. All strains coming from this study grouped very well with their respective

reference strains.

In the case of Aspergillus species, the gene sequencing was well supported by its respective

dendrograms (Figure 9). Aspergillus niger (CCCT 18.11, 18.12, 18. 14, 18.15, 18.19, 18.22, 18.26,

18.27, 18.28, 18.30 and 18.89), A. awamori (CCCT 18.16, 18.18, 18.20, 18.50 and 18.53) and A.

brasiliensis (CCCT 18.23) are related in a well-defined cluster forming A. niger clade (bootstrap >80%).

The cluster of A. luchuensis (CCCT 18.70) and A. tubingensis (CCCT 18.21) are associated with poor

values of bootstrap of 64%, however, these fungi belong to the same clade (A. tubingensis clade). The A.

tubingensis and A. niger clade correspond to A. niger aggregate of the section Nigri.

All A. flavus (CCCT 18.35, 18.36, 18.120, 18.126, 18.128 and 18.186) are related forming the

section Flavi with very well bootstrap values of 99%. In the case of A. fructus (CCCT 18.145 and 18.182)

and A. sydowii (CCCT 18.92), a well-defined phylogenetic closeness with bootstrap values of 99% is

observed. Aspergillus sydowii belongs to A. sydowii subclade and A. fructus belongs to A. versicolor

subclade, both fungi are grouped in the clade A. versicolor and are belonging to section Versicolores.

Two isolated strains of A. pseudoglaucus (CCCT 18.85 and 18.97) are related in a cluster with

bootstrap of 99%. Finally, the strains A. fumigatus CCCT 18.94 and 18.103 are related with their

reference strains in a cluster with bootstrap of 99%.

Page 33: Mycobiota in Chilean chilli Capsicum annuum L. used for

26

Figure 9. Evolutionary relationships of Aspergillus isolates (BenA). The optimal tree with the sum of branch length = 0.59762008 is shown. The analysis involved 48 nucleotide sequences. All positions containing gaps and missing

data were eliminated. There were a total of 179 positions in the final dataset. “T”: type strain. Trichocoma paradoxa CBS 788.83 is used as outgroup.

Page 34: Mycobiota in Chilean chilli Capsicum annuum L. used for

27

Robustness of Penicillium species identification was added to the cluster by the addition of BenA

nucleotide sequences of P. expansum (CBS 28197), P. crustosum (CBS 101025 and 47184), P.

viridicatum (CBS 101034), P. polonicum (CBS 101479), P. cyclopium (CBS 47784 and 101136), P.

chrysogenum (CBS 306.48), P. brevicompactum (CBS 110067, 110068, 110069 and 257.29), P.

angulare (NRRL 35683), P. glabrum (NRRL 66390 and CBS 125543) from GenBank database. All strains

coming from this study grouped very well with their respective reference strain.

In the case of Penicillium species, the genes sequencing were well supported by its respective

dendrograms (Figure 10). Penicillium expansum (CCCT 18.131 and 18.116) are related in a well-defined

cluster with bootstrap values >90%. For P. polonicum (CCCT 18.29, 18.42, 18.55, 18.72, 18.80 and

18.88), P. viridicatum (CCCT 18.43, 18.44, 18.95, 18.111, 18.192, 18.196 and 18.204), P.

melanoconidium (CCCT 18.37 and 18.54), P. cyclopium (CCCT 18.60, 18.64, 18.66, 18.71, 18.81,

18.83, 18.84, 18.106, 18.124, 18.203 and 18.215) and P. crustosum (CCCT 18.46, 18.52, 18.62,

18.77, 18.82, 18.96, 18.99, 18.102, 18.117, 18.119, 18.123, 18.132, 18.134 and 18.138) the cluster

present bootstrap values >70%.

All these strains belong to the serie Viridicata. Penicillium verrucosum (CCCT 18.33) and P.

discolor (CCCT 18.91) belong to the serie Verrucosa and serie Camemberti, respectively. This large

cluster of series Viridicata, Verrucosa and Camemberti correspond to the section Fasciculata. The three

P. chrysogenum (CCCT 18.170, 18.200 and 18.207) were grouped in a cluster with bootstrap values of

99%.

In the case of P. bialowiezense (CCCT 18.38), P. brevicompactum (CCCT 18.31, 18.39, 18.49,

18.57, 18.58, 18.59, 18.67, 18.68, 18.74, 18.90, 18.104, 18.105, 18.113, 18.129, 18.130, 18.139,

18.168 and 18.194) and P. buchwaldii (CCCT 18.107) were grouped in a very well cluster with bootstrap

value of 99%. These strains belong to the section Brevicompacta. Penicillium paraherquei (CCCT 18.40

and 18.65) are present in one cluster with bootstrap value of 100%. Penicillium waskmanii (CCCT 18.65)

and P. citrinum do not belong to the same cluster, however, they are very close related at genetic level;

both fungal species belong to the section Citrina.

Penicillium angulare (CCCT 18.109 and 18.110) and P. adametzioides (CCCT 18.108) are

associated in a cluster with bootstrap values of 99%. These fungi belong to section Sclerotiora. Finally, P.

glabrum (CCCT 18.32, 18.47, 18.51, 18.61, 18.63, 18.75, 18.76, 18.86, 18.87, 18.93, 18.96, 18.100,

Page 35: Mycobiota in Chilean chilli Capsicum annuum L. used for

28

18.101, 18.115, 18.118, 18.122, 18.135, 18.137 and 18.144) are associated in a well-defined cluster

(bootstrap value of 100%). these filamentous fungi belong to section Aspergilloides.

Figure 10. Evolutionary relationships of Penicillium isolates (BenA). The optimal tree with the sum of branch length

= 1.25708392 is shown. The analysis involved 109 nucleotide sequences. All positions containing gaps and

missing data were eliminated. There were a total of 276 positions in the final dataset. “T”: type strain. Talaromyces

flavus CBS 310.38 is used as outgroup.

Page 36: Mycobiota in Chilean chilli Capsicum annuum L. used for

29

Figure 10. Continued.

Talaromyces flavus CBS 310.38

Page 37: Mycobiota in Chilean chilli Capsicum annuum L. used for

30

3.3 Preliminary genetic characterisation of potentially OTA producer fungal strains

In order to identify, preliminarily, a putative OTA biosynthetic genes, amplification of otapksPN,

otanpsPN, otachlPN and otatraPN were performed (Figure 11). Penicillium verrucosum (CCCT 18.33),

Penicillium crustosum (CCCT 18.46), Penicillium nordicum (CBS 112573) and Aspergillus carbonarius

(MUM 01.08) were used for this analysis.

Considering the four genes under study, our results show that P. nordicum (CBS 112573) was

positive for the otanpsPN (≈700 bp), otapksPN (≈500 bp) and otatraPN (≈420 bp) amplicons,

nevertheless, negative for otachlPN. For P. verrucosum (CCCT 18.33), this was positive for the otanpsPN

(≈700 bp) and otatraPN (≈420 bp) and negative for otachlPN and otapksPN. In the case of P. crustosum

(CCCT 18.46), this was positive only for otatraPN (≈420 bp). A. carbonarius (MUM 01.08) was negative

for all genes tested (present only the IAC).

Figure 11. Agarose gel electrophoretic pattern of amplification of otapksPN, otanpsPN, otachlPN and otatraPN for P. nordicum (P.n.), P. verrucosum (P.v.), P. crustosum (P.c.) and A. carbonarius (A.c.). B= blank.

Page 38: Mycobiota in Chilean chilli Capsicum annuum L. used for

31

3.4 Quantitative analyses of OTA and AFLs contamination in Merkén

For verifying the co-occurrence of AFLs and OTA, twenty one Merkén samples from 8 farmers

and 13 markets from the Chilean Region of La Araucanía were analysed by HPLC with fluorescence

detection (Table 1-2). The retention time for the mycotoxins analysed was: AFB1, 17.18 min, AFB2, 14.39

min, AFG1, 12.90 min, AFG2, 11.00 min and OTA 11.28 min (For standard curves, see Annex II).

In the case of AFLs, only AFB1 were detected in 12 samples of Merkén (57%). AFB2, AFG1 and

AFG2 were not detected in all Merkén samples. Concentrations of AFB1 were in the range of 0.19±0.26-

1.44±0.10 µg/kg for 6 farmer samples. In markets samples, only 6 samples present AFB1

concentrations in the range of 0.29±0.37-1.67±0.32 µg/kg. The samples present very low concentration

of AFLs (near to the LOD). In relation to recovery rates (%), AFB1, AFB2, AFG1 and AFG2 present

48.91±3.61%, 66.58±0.97%, 46.16±0.65% and 70.47±0.57% respectively.

For OTA, the samples analysed showed a wide range of contamination. OTA was detected in all

Merkén samples in the range of 0.79±0.05-5.99±0.68 µg/kg for farmer samples and 0.83±0.83-

19.81±0.70 µg/kg for markets samples. The recovery rates (%) for OTA was 97.57±19.91%.

Samples OTA

(µg/kg)

AFB1 (µg/kg) AFB2 (µg/kg) AFG1 (µg/kg) AFG2 (µg/kg)

Farmer I 5.71±0.11 0.64±0.30* ND ND ND

Farmer II 3.38±0.24 0.19±0.26* ND ND ND

Farmer III 1.99±1.99 0.72±1.01* ND ND ND

Farmer IV 2.46±0.29 0.45±0.63* ND ND ND

Farmer V 2.39±0.31 0.50±0.11* ND ND ND

Farmer VI 0.91±0.26 1.44±0.10* ND ND ND

Farmer VII 5.99±0.68 ND ND ND ND

Farmer VIII 0.79±0.05 ND ND ND ND

Table 1. AFLs and OTA results from 8 farmers of the La Araucania Region, Chile. Values are the

average of two independent replicates ± standard deviation. Data shown in this table were not

corrected for recoveries.

*= Very close to LOD; ND= Not Detected.

Page 39: Mycobiota in Chilean chilli Capsicum annuum L. used for

32

4. DISCUSSION

Three different culture media (DG18, DRBC and MEA) were used to isolate the mycobiota on

each step of Capsicum annuum L. cv. "Cacho de Cabra” berry fruits chain production used for Merkén

production. The culture medium DG18 followed by DRBC showed the greatest potential to isolate fungi

from the collected samples (42% and 39%, respectively). This is due to the lower aw of both culture media

Samples OTA (µg/kg) AFB1 (µg/kg) AFB2 (µg/kg) AFG1 (µg/kg) AFG2 (µg/kg)

Market 1 2.89±0.39 ND ND ND ND

Market 2 2.8±0.32 ND ND ND ND

Market 3 13.56±1.56 ND ND ND ND

Market 4 6.49±2.05 0.95±0.23* ND ND ND

Market 5 3.23±0.60 1.62±0.23* ND ND ND

Market 6 3.50±0.12 0.29±0.41* ND ND ND

Market 7 10.36±0.69 ND ND ND ND

Market 8 1.37±0.13 0.29±0.37* ND ND ND

Market 9 1.43±0.08 ND ND ND ND

Market 10 1.70±0.17 0.33±0.47* ND ND ND

Market 11 5.99±2.73 ND ND ND ND

Market 12 19.81±0.70 1.67±0.32* ND ND ND

Market 13 0.83±0.83 ND ND ND ND

Table 2. AFLs and OTA results from 13 markets of the La Araucania Region, Chile. Values are the

average of two independent replicates ± standard deviation. Data shown in this table were not

corrected for recoveries.

*= Very close to LOD; ND= Not Detected.

Page 40: Mycobiota in Chilean chilli Capsicum annuum L. used for

33

(0.955 and 0.950, respectively) that usually gives good results for the isolation of xerophilic filamentous

fungi.

In this work, the amplification and sequencing of the housekeeping BenA gene was used for the

identification of fungi at the species level. Gene sequencing is reported in the last years as the gold

standard for the fungal identification at species level (Skouboe et al., 1999; Samson et al., 2004a; Geiser

et al., 2007; Houbraken et al., 2011; Rodrigues et al., 2011; Schoch et al., 2012; Soares et al., 2012;

Decontardi et al., 2017; Decontardi et al., 2018).

Our results showed that species from Alternaria, Aspergillus, Fusarium and Penicillium were the

most commonly species isolated from each step of C. annuum berry fruits chain production used for

Merkén production (Figure 7). These filamentous fungi found in chilli berry fruits can fall into two broad

ecological categories, field and storage fungi. Field fungi belonging to Alternaria and Fusarium species

could be of phytopathological and mycotoxicological risk during the field growth and ripening of the fruits.

Meanwhile, storage-related fungi that belong mainly to the genera Aspergillus and Penicillium could cause

a serious problem of spoilage and mycotoxin contamination of these agri-food products.

Previous studies reported mycotoxigenic fungi and mycotoxins in different varieties of C. annuum

(Hierro et al., 2008; Santos et al., 2011; Salari et al., 2012; Ham et al., 2016). In the case of Aspergillus

and Penicillium, some studies have identified a high incidence of these fungi and mycotoxins as AFLs

and OTA (Santos et al., 2008; Santos et al., 2011; Van de Perre et al., 2014; Ikoma et al., 2015;

Gambacorta et al., 2018; Motloung et al., 2018).

Due to the climatic conditions of the central-south region of Chile which is characterised as

temperate oceanic, the fungi that are mostly favoured, at growth and colonise ability, are those belonging

to the Penicillium genus. At ecologically point of view, Penicillium species are mostly saprophytes and the

majority of species show optimal growth at moderate to low temperatures (near of 25 °C) and are capable

to grow at aw below 0.9. (Frisvad and Samson, 1991; Pitt and Hocking, 2009a). According to the results

observed in the present study, Penicillium species was found in all sampling point as the most abundant

filamentous fungus.

In the sampling point I, just at the day of ripe fruits harvest, the main fungal species found belong

to Alternaria, Fusarium and Penicillium genera. According to the results obtained in the present study, a

high incidence of Penicillium followed by Fusarium and Alternaria was found in each step of chain

production.

Among Fusarium species, F. oxysporum was the most abundant (36%). Fusarium oxysporum is

a pathogenic fungus common in soils and the causative agent of Fusarium wilt, a deadly vascular wilting

Page 41: Mycobiota in Chilean chilli Capsicum annuum L. used for

34

syndrome in plants (Booth, 1971). In the case of Alternaria species, A. Alternata was the most abundant

one (89%). Fusarium oxysporum is a pathogen for plants, mainly affecting leaves and less severely to the

stolons, overall leading to the complete death of the plant. These results are consistent with the conditions

of the fruit and its proximity to the field (Nelson et al., 1981; Silva and Singh, 2012; Bajwa et al., 2010).

In the sampling point II, after 1 month of harvest (drying process), Fusarium was not found. In

contrast, Penicillium and Aspergillus began to be the abundant and unique fungal genera found in the

dry form of the berry fruits, taking into consideration the culture media used for the mycobiota isolation.

Penicillium crustosum and P. glabrum were the most representatives Penicillium species. Aspergillus

niger and A. flavus were found as other important fungal species in the sampling point II. These results

can be explained due to the decreasing in the aw (drying process).

The low aw could contribute to selectivity for moderate xerophilous fungi. Moreover, the ability of

some Aspergillus and Penicillium species to generate ascospores can be an advantage in relation to other

fungi, once it allows their survival to drying process (Dijksterhuis, 2007), e.g., conidia of A. niger, P.

chrysogenum and P. glabrum, exhibit a high resistance and low reduction of viable spores in a range of

56 and 62 °C (Baggerman and Samson, 1988).

In the sampling point III, during the smoking process, results obtained show a total absence of

Alternaria and Fusarium species. In this case, species belonging to Penicillium, Aspergillus, Trichoderma

and Microaascos were found. For Penicillium genus, the section Fasciculata, (P. crustosum, P.

viridicatium, P. verrucosum, P. discolor, P. melanoconidium, P. polonicum and P. cyclopium) was the

most representative. Most species of this section can grow well at 15 to 25 °C (except those species of

the series Verrucosa, i.e. P. verrucosum), and at low aw (Houbraken et al., 2016). Section Fasciculata

contains species that commonly occur on stored or manufactured foods (Houbraken et al., 2016).

Also, in the sampling point III, the follow species belonging to series Viridicata were found: P.

crustosum, P. viridicatium, P. melanoconidium, P. polonicum and P. cyclopium. Penicillium species

belonging to the series Viridicata are typically associated with stored cereal grains and those belonging to

series Verrucosa are associated with stored cereal grains and dried or salted meat products (Frisvad and

Samson, 2004; Houbraken et al., 2016). Species belonging to series Camemberti (i.e., P. discolor)

typically occur on foods with a high protein or lipid content like Capsicum annum berry fruits (Houbraken

et al., 2016).

On the other hand, fungal species belonging to the section Brevicompacta (i.e., P.

brevicompactum) and the section Aspergilloides (i.e., P. glabrum) represent the seconds most

representative groups in all the Peniciillia found at this sampling point. Penicillium brevicompactum grows

Page 42: Mycobiota in Chilean chilli Capsicum annuum L. used for

35

between -2°C and 30°C, with an optimum growth temperature around 23 °C, and can grow in an

environment with aw below to 0.78. It is one of the most xerophilic species of the Penicillium genus.

Penicillium glabrum can grows under temperature ranging from 0 to 35 °C and has been referred as a

xerophilic fungus (Pitt and Hocking, 1997).

Still about the fungal sample isolated from the sampling point III, the species of Aspergillus

section Nigri (A. niger, A. awamori and A. tubingensis) were the most representative (51%). Aspergillus

species have the capacity to develop in high temperature conditions. Relatively low aw allows them to

adapt and colonise a wide variety of cereals and dried fruits. Temperature ranging from15 to 30 °C and

aw ranging from 0,97 to 0.85 means the best growth conditions for Aspergillus section Nigri (Astoreca et

al., 2007).

On the other hand, the section Flavi (A. flavus, 11%), represented the second most representative

group in all Aspergillus species found at the sampling point III as well as in the sampling point II (25%).

Aspergillus flavus has been reported by some authors as an important fungus in the production of

mycotoxins and food contamination such as almonds, maize, peanuts (Rodrigues et al., 2011; Soares et

al., 2012; Fountain et al., 2015; Fakruddin et al., 2015).

The ecophysiology of the fungi, their natural occurrence on the environmental conditions at the

place of berry fruits of C. annuum production, and the low aw of the studied substrate can explain the

reason these fungal species are the most common found in all sampling points. Interestingly, according

to the fungal species succession observed in the present study, based on the mycobiota isolated from

each step of C. annuum berry fruits chain production, used for Merkén production, a selectively increasing

on the occurrence of Aspergillus and Penicillium species was observed (Figure 7 and 8).

Potential mycotoxigenic fungi found in the present study include 1) for Penicillium species: P.

brevicompactum, potentially producer of Brevianamide A (BrA) and Mycophenolic acid;, P. verrucosum,

potentially producer of BrA, Citrinin (CIT) and OTA; P. viridicatum, potentially producer of BrA and

Xanthomegnin; P. citrinum, potentially producer of CIT; P. chrysogenum, potentially producer of

Cyclopiazonic acid (CPA); P. crustosum, potentially producer CPA and Penitrem A; and P. expansum,

potentially producer of Patulin and CIT); and 2) for Aspergillus species: A. niger, potentially producer of

OTA and Fumonisin B2 and B4; and A. flavus, potentially producer of AFls and CPA.

Overall, in the present study, it was observed that Aspergillus section Nigri (i.e., A. niger) was the

group of potentially mycotoxigenic fungi with the highest incidence. Previous studies reported that

mycotoxigenic strains of Aspergillus section Nigri has a high incidence in samples with similar

characteristics (Martín et al., 2005; Almela et al., 2007; Varga et al., 2011). Taking into consideration

Page 43: Mycobiota in Chilean chilli Capsicum annuum L. used for

36

the current scenario of climate change and at the point of view of food safety, the results obtained in the

present study can represent a worrying.

Some authors have referred to the impact of climate change on the increasing of mycotoxin

production by mycotoxigenic fungi, which increases the incidence of mycotoxins with carcinogenic effects

in food products (Paterson and Lima, 2010a; Paterson and Lima 2010b; Battilani et al., 2016). For this

reason, contamination problems in food commodity, and particularly Merkén, can be increased leading,

consequently, to a seriously problem of public health in people consuming Merkén in Chile and abroad.

At genetic point of view, BenA region seems suitable. It shown good clustering with bootstrap

values higher than 70% (Hall, 2013), both for the strains obtained from chilli berry fruits in all sampling

points and for the reference strains.

In the case of genetics similarities of Aspergillus species, no atypical grouping values were found.

The overall clustering was good with bootstrap values commonly above 85%. However, closely related

species (i.e., A. niger and A. awamori, that belong to the section Nigri, Clade A. niger) were not

discriminated (Figure 9). These species are very closely related, are morphologically indistinguishable

and in some cases the genetic regions used to differentiate them are not effective, as is the case of ITS

regions.

The status of A. awamori has been revised several times and has been named with various

synonyms (Sakaguci et al., 1951; Samson et al., 2004b; Varga et al., 2011). However, in recent years,

due to their genetic differences, A. awamori has been classified as a cryptic phylogenetic specie of A.

niger (Perrone et al., 2011). Results obtained in the present study show that gene BenA may not be the

most appropriate choice for molecular identification of these species and, in general, sequencing of more

than one gene is required to obtain a robust identification. Some authors show similar results in their

phylogenetic trees using BenA, and in most cases, the use of more than one gene, or multilocus,

phylogenetic trees help to differentiate A. awamori from A. niger (Perrone et al., 2011; Varga et al., 2011).

Phylogenetic analyses of calmodulin sequences (CaM) and the translation elongation factor-1

alpha (TEF-1α) have been reported as good genetic regions for separating A. niger from A. awamori

(Perrone et al., 2011; Varga et al., 2011). In addition, some sequences of genes coding for enzymes,

associated in the production of OTA reported by Geisen (2006), such as chloroperoxidase, show a great

performance to differentiate A. niger from A. awamori at the phylogenetic level (Varga et al., 2011). The

amplification and sequencing of these regions in A. niger and A. awamori could add required information

to separate these fungal species found in chilli berry fruits.

Page 44: Mycobiota in Chilean chilli Capsicum annuum L. used for

37

The other species grouped in this phylogenetic tree do not present atypical groupings, that is,

they are associated to their respective strains and reference strains with values of bootstrap over 90%.

In the case of genetics similarities of Penicillium species, no atypical grouping values were found

(Figure 10). The overall clustering was good with bootstrap values commonly above 70%. Moreover, some

strains show variability within their clusters. The isolated strains of P. expansum (CCCT 18.131 and

18.116) were not associated to the same cluster with their reference strain (CBS 28197). However, they

present a bootstrap of 91%, which shows its high similarity of the gene BenA. In addition, more reference

strains from GenBank database could provide more robustness to the cluster and associate these species

in the same cluster.

The same trend is found in the clusters generated for P. brevicompactum (Figure 10). Some

intraspecific variability can be found in this section. Although the clusters are well defined, this variability

could indicate that P. brevicompactum could be more diversified than is currently known.

In the case of the cluster generated for P. crustosum, P. viridicatum, P. melanoconidium, P.

polonicus and P. cyclopium, these are genetically very close to each other. Several studies show similar

groupings among Penicillium species, which is not surprising, due to the fact that these fungi are classified

within the Section Fasciculata (Frisvad et al., 2004; Frisvad et al., 2013; Visagie et al., 2014a; Visagie et

al., 2014b).

Concerning all P. chrysogenum strains, these are associated in a well-identified cluster with

Bootstrap values of 99% (Figure 10). It is important to note that in Penicillium, BenA has limitations to

discriminate P. chrysogenum from P. camembertii species complexes. Studies conducted by Houbraken

et al. (2012) on the Penicillium section Chrysogena, showed that a large number of species are

phylogenetically closely related, and showed that different genes suggest different phylogenies. For

example, in some cases, P. chrysogenum cannot be differentiated from P. allii-sativii, even though CaM

distinguishes P. chrysogenum from P. allii-sativi (Houbraken et al., 2012). The results obtained in the

present study show a good grouping of P. chrysogenum, which can due to the low amount of isolates of

P. chrysogenum and species of the series Camemberti (Figure 10).

The fungal strains from chilli berry fruits classified as P. glabrum showed a high intraspecific

variability, which is observed in a high number of clusters within a larger cluster (Figure 10, Section

Aspergilloides). These data are supported by those information found in the literature.

Penicillium glabrum presents a great variability at the genetic level. Several studies show that

different sequence types of BenA within P. glabrum can be found (Basílio et al., 2006, Barreto et al.,

Page 45: Mycobiota in Chilean chilli Capsicum annuum L. used for

38

2011). Some authors classify the isolates of P. glabrum as "Penicillium glabrum complex" due to the

great variability between these species (Barreto et al., 2011; Houbraken et al., 2014).

In relation with the genetic characterisation of potentially OTA producer fungal strains, our

preliminary results (Figure 11) show, low congruence with the results reported in literature (Geisen et al.,

2006). For gene amplification, P. nordicum (CBS 112573) was positive for otapksPN (≈500 bp),

otanpsPN(≈720 bp), and otatraPN (≈420 bp). P. verrucosum (CCCT 18.33) was positive for otanpsPN

(≈720 bp) and otatraPN (≈420 bp). In P. crustosum (CCCT 18.46) only otatraPN (≈420 bp) was present.

Although the PCR reaction was performed correctly due to the presence of the IAC (ITS1-5.8S-

ITS2 region) it was not possible amplify all expected genes. First, we associate the absence of the

otachlPN gene due to a low temperature of annealing, this is because the primers used to amplify this

gene have a melting temperature of 68.1°C, which is high in comparison to the other primers and at the

alignment temperature used in the reaction (60 °C). However, to clarify this hypothesis, this gene was

separately amplified using the annealing temperature of 60 °C and no band was obtained (data not

shown). The negative results in this experiment can be due to an inefficient amplification protocol.

Although these data are preliminary, it is advisable to optimize the PCR conditions and optimise the

primers and DNA concentration, to obtain a better amplification and associate these genes with the

potential ochratoxigenic fungi in this samples.

In this study, the mycotoxins contamination of Merkén from the farmers and markets of the La

Araucania region are ranged between 0.79±0.05 to 19.81±0.70 µg/kg for OTA and 0.29±0.37 to

1.67±0.32 µg/kg for AFB1.

The Merkén samples showed a 100% OTA contamination however, only the sample from market

12 presents OTA concentration higher (19.8 ±0.70 μg / kg) than those established by the European

Union. Although the concentrations of OTA in the Merkén samples are within the European limit, the high

incidence of OTA is a serious warning. On the other hand, our results show that OTA concentration is in

general higher in the samples collected from markets than those provided by the farmers. This could be

due to poor hygiene conditions and increased humidity in the stores. Many studies show that the

occurrence of OTA contamination is greater in condiments in the process of storage and sale (Mandeel,

2005; Schweiggert et al., 2005; Santos et al., 2008; Santos et al., 2010c).

The high incidence of OTA could be due to the high incidence of black Aspergillus (i.e. A. niger

complex) in the early stages of Merkén production. Our results show that the Merkén production process

promote this type of mycotoxigenic fungi. However, it is necessary to analyse the mycobiota in subsequent

processes (grinding, additives, storage and sale) to verify other possible ochratoxigenic fungi.

Page 46: Mycobiota in Chilean chilli Capsicum annuum L. used for

39

OTA is classified as a group 2B carcinogen (IARC, 1993) and several studies show its high toxicity

and cumulative effect on the mammal’s body (Smith et al., 1994; Peterson and Lima, 2010b; Paradellsa

et al., 2015). The high incidence of OTA in the Merkén samples, added to the high and continuous

consumption of food condiment by the Chilean population, should be an important point to consider by

the health authorities. Studies conducted by Ikoma et al. (2015) show that the high incidence of OTA in

red chili peppers in specific geographical areas could be associated with a high incidence of Gallbladder

Cancer in the population.

In the case of AFLs, only AFB1 are present in Merkén samples (57%) in a very low concentration.

Our results only show 6 isolates of A. flavus that could not contaminate the samples significantly.

5. CONCLUSIONS

This work provides for the first time information about mycobiota in berry fruits of Capsicum

annuum L. cv. "Cacho de Cabra” used for Merkén’s production. This study confirms Aspergillus and

Penicillium are the two dominant fungal genera that emerge in the early stages of the Merkén production

process. Overall, these results show that the mycobiota present in C. annuum berry fruits during its

production process, before the Merkén production, is increasingly selective for occurrence potentially

mycotoxigenic fungi such as A. niger, A. flavus, P. expansum and P. verrucosum. Based on the identified

fungal species, the raw material used for Merkén production can be at risk of mycotoxin contamination.

The high co-occurrence of potentially mycotoxigenic fungi found in berry fruits of C. annuum used

for Merkén’s production and the high incidence of OTA in the Merkén samples, warning for the

assessment of the mycotoxigenic potential of isolated strains in the 3 points of fungal isolation.

The genetic characterisation of the mycotoxigenic potential of isolated fungal strains, and the

identification of mycobiota and mycotoxins in the milling and Merkén storage, should be continued in

order to stablish possible contamination points, providing input for effective strategies to mitigate

mycotoxin contamination and identification of the critical control points to generate a safe and high-quality

food product.

Page 47: Mycobiota in Chilean chilli Capsicum annuum L. used for

40

6. REFERENCES

Abarca, M.L., Accensi, F., Bragulat, M.R. & Cabañes F.J. (2001). Current importance of ochratoxin A-

producing Aspergillus spp. Journal of Food Protection, 64, 903-906.

Abrunhosa, L., Ines, A., Rodrigues, A.I., Guimaraes, A., Pereira, V.L., Parpot, P., Mendes Faia, A. &

Venâncio, A. (2014). Biodegradation of ochratoxin A by Pediococcus parvulus isolated from Douro

wines. International Journal of Food Microbiology 188, 45-52.

Agustini, B.C., Silva, L.P., Bloch, C., Jr., Bonfim, T.M. & da Silva, G.A. (2014). Evaluation of MALDI-TOF

mass spectrometry for identification of environmental yeasts and development of supplementary

database. Applied Microbiology and Biotechnology, 98, 5645-5654.

Almela, L., Rabe, V., Sánchez, B., Torrella, F., López-Pérez, J.P., Gabaldón, J.A. & Guardiola L. (2007).

Ochratoxin A in red paprika: relationship with the origin of the raw material. Food Microbiology,

24, 319-327.

Astoreca, A., Magnoli, C., Ramirez, M.L., Combina, M. & Dalcero, A. (2007). Water activity and

temperature effects on growth of Aspergillus niger, A. awamori and A. carbonarius isolated from

different substrates in Argentina. International Journal of Food Microbiology, 119, 314-318.

Baggerman, W.I. & Samson, R.A. (1988). Heat resistance of fungal spores. Food-borne fungi 3rd edition

(Samson, R. A., Reenen-Hoekstra, E. S. van, eds). Centraalbureau voor Schimmelcultures, Baarn,

The Netherlands.

Bajwa, R., Mukhtar, I. & Mushtaq, S. (2010). New report of Alternaria alternata causing leaf spot of Aloe

vera in Pakistan. Canadian Journal of Plant Pathology, 32, 490-492.

Barreto, M.C., Houbraken, J., Samson, R.A., Frisvad, J.C. & San-Romão, M.V. (2011). Taxonomic studies

of the Penicillium glabrum complex and the description of a new species P. subericola. Fungal

Diversity, 49, 23-33.

Basílio, M.C., Gaspar, R., Silva-Pereira, C. & San Romão, M.V. (2006). Penicillium glabrum cork colonising

isolates-preliminary analysis of their genomic similarity. Revista Iberoamericana de Micología, 23,

151-154.

Battilani, P., Toscano, P., Van der Fels-Klerx, H.J., Moretti, H.J., Camardo-Leggieri, M., Brera, C., Rortais,

A., Goumperis, T. & Robinson, T. (2016). Aflatoxin B1 contamination in maize in Europe

increases due to climate change. Nature Scientific Report, 6, 24328.

Bhabhra, R. & Askew, D.S. (2005). Thermotolerance and virulence of Aspergillus fumigatus: role of the

fungal nucleolus. Medical Mycology Supplement, 43, 87-93.

Bhatnagar, D., Ehrlich, K.C. & Cleveland, T.E. (2003). Molecular genetic analysis and regulation of

aflatoxin biosynthesis. Applied Microbiology and Biotechnology, 61, 83-93.

Bidartondo, M.I., et al., (2008). Preserving Accuracy in GenBank. Science, 319, 1616.

Booth, C. (1971). The genus Fusarium. Kew, Surrey, UK. Commonwealth Mycological Institute 237 pp.

Cabañes, F.J. (2002). Identificación de hongos: de la morfologia al análisis molecular. In: Ecologia dos

fungos, Santos M., Venâncio A., Lima N., (Eds.). Micoteca da Universidade do Minho, Braga, 1,

19-29.

Page 48: Mycobiota in Chilean chilli Capsicum annuum L. used for

41

Chang, S., Carneiro-Leao, M.P., de Oliveira, B.F., Souza-Motta, C., Lima N., Santos C. & de Oliveira N.T.

(2016). Polyphasic approach including MALDI-TOF MS/MS analysis for identification and

characterisation of Fusarium verticillioides in Brazilian corn kernels. Toxins, 8, 1-13.

Chaves da Silva, F., Chalfoun, S.M., Batista, L.R., Santos C. & Lima, N. (2015). Use of a polyphasic

approach including MALDI-TOF MS for identification of Aspergillus section Flavi strains isolated

from food commodities in Brazil. Annals of Microbiology, 65, 2119-2129.

Chen, S.C.A., Halliday, C.L. & Meyer, W. (2002). A review of nucleic acid-based diagnostic tests for

systematic mycoses with an emphasis on polymerase chain reaction-based assays. Medical

Mycology, 40, 333-357.

Costea, P.I., Zeller, G., Sunagawa, S., Pelletier, E., Alberti, A., Levenez, F., Tramontano, M., Driessen,

M., Hercog, R., Jung, F.E., Kultima, J.R., Hayward, M.R., Coelho, L.P., Allen-Vercoe,

E., Bertrand, L., Blaut, M., Brown, J.R.M., Carton, T., Cools-Portier, S., Daigneault, M., Derrien,

M., Druesne, A., de Vos, W.M., Finlay, B.B., Flint, H.J., Guarner, F., Hattori, M., Heilig, H., Luna,

R.A., Van Hylckama Vlieg, J., Junick, J., Klymiuk , I., Langella, P., Le Chatelier, E., Mai,

V., Manichanh, C., Martin, J.C., Mery, C., Morita, H., O'Toole, P.W., Orvain, C., Patil,

KR., Penders, J., Persson, S., Pons, N., Popova, M., Salonen, A., Saulnier, D., Scott, K.P., Singh,

B., Slezak, K., Veiga, P., Versalovic, J., Zhao, L., Zoetendal, E.G., Ehrlich, S.D., Dore, J. & Bork,

P. (2017). Towards standards for human faecal sample processing in metagenomic studies.

Nature Biotechnology, 35, 1069-1076.

Da Cruz Pedrozo Miguel, M.G., De Castro Reis, L.V., Efraim, P., Santos, C., Lima, N. & Schwan R.F.

(2017). Cocoa fermentation: Microbial identification by MALDI-TOF MS, and sensory evaluation

of produced chocolate. Food Science and Technology, 77, 362-369.

Decontardi, S., Mauro, A., Lima, N. & Battilani, P. (2017). Survey of Penicillia associated with Italian

grana cheese. International Journal of Food Microbiology 246, 25-31.

Decontardi, S., Soares, C., Lima, N. & Battilani P. (2018). Polyphasic identification of Penicillia and

Aspergilli isolated from Italian grana cheese. Food Microbiology 73, 137-149.

Di Pillo,F., & Martínez, N. (2018). Ocratoxina A en ají y merkén, Chile. Perfil de Riesgo/ACHIPIA

N.°02/2018 version Nº1. available in https://www.achipia.gob.cl/inicio/prensa-y-

publicaciones/evaluaciones-de-riesgo access on 26/06/2018

Dijksterhuis, J. (2007). Heat resistant ascospores. in: Dijksterhuis, J., Samson, R.A., (Eds.), Food

Mycology - A multifaceted approach to Fungi and Food. Boca Raton, FL: CRC Press, 101-120.

Dutcher, S.K. (2001). The tubulin fraternity: alpha to beta. Current Opinion in Cell Biology 13, 49-54.

El-Kady, I.A., El-Maraghy, S.S.M. & Mostafa, M.E. (1995). Natural occurrence of mycotoxins in different

spices in Egypt. Folia Microbiologica, 40, 297-300.

El-Khoury, A. & Atoui, A. (2010). Ochratoxin A: General Overview and Actual Molecular Status. Toxins, 2,

461-493.

European Commission (2006). Commission Regulation (EU) No. 1881/2006 of 19 December 2006

setting maximum levels for certain contaminants in foodstuffs. Official Journal of the European

Union, 364, 5-24.

Page 49: Mycobiota in Chilean chilli Capsicum annuum L. used for

42

European Commission (2015). Commission Regulation (EU) No. 1137/2015 of 13 July 2015 amending

Regulation (EC) No. 1881/2006 as regards the maximum level of Ochratoxin A in Capsicum spp.

spices. Official Journal of the European Union, 185, 11-12.

Fakruddin, M., Chowdhury, A., Hossain, M.N. & Ahmed, M.M. (2015). Characterization of aflatoxin

producing Aspergillus flavus from food and feed samples. SpringerPlus, 4, 159.

FAOSTAT, (2017). About all core data. Available at fao.org/faostat/en/?#data/QC/visualize. Visited 15

June 2017.

Fatimoh, A.O., Moses, A.A., Adekunle, O.B. & Dare, O.E. (2017). Isolation and Identification of Rot Fungi

on Post-Harvest of Pepper (Capsicum annuum L.) Fruits. AASCIT Journal of Biology, 3, 24-29.

Fazekas, B., Tar, A. & Kovács, M. (2005). Aflatoxin and ochratoxin A content of spices in Hungary. Food

Additives and Contaminants, 22, 856-863.

Feau, N., Decourcelle, T., Husson, C., Desprez-Loustau, M.L. & Dutech, C. (2011). Finding single copy

genes out of sequenced genomes for multilocus phylogenetics in non-model fungi. PLoS One 6,

18803.

FIA, Ministerio de Agricultura. (2010). Resultados y Lecciones en Ají Merkén con Alto Valor Agregado.

Proyecto de Innovación en Región de La Araucanía, Chile, 79, 6-10.

Fountain, J.C., Khera, P., Yang, L., Nayak, S.N., Scully, B.T., Lee, R.D., Chen, Z.Y., Kemerait, R.C.,

Varshney, R.K. & Guo, B. (2015). Resistance to Aspergillus flavus in maize and peanut: Molecular

biology, breeding, environmental stress, and future perspectives. The Crop Journal, 3, 229-237.

Frisvad, J.C. & Samson, R.A. (1991). Filamentous fungi in foods and feeds: ecology, spoilage and

mycotoxin production. In: Hand-book of Applied Mycology- Foods and Feeds, Arora D.K., Mukerji

K.G. & Marth E.H. (Eds.), 1, 31-68.

Frisvad, J.C. & Samson, R.A. (2004). Polyphasic taxonomy of Penicillium subgenus Penicillium. A guide

to identification of the food and air-borne terverticillate Penicillia and their mycotoxins. Studies in

Mycology, 49, 1-173.

Frisvad, J.C., Houbraken, J., Popma, S. & Samson, R.A. (2013). Two new Penicillium species Penicillium

buchwaldii and Penicillium spathulatum, producing the anticancer compound asperphenamate.

FEMS Microbiology Letters, 339, 77-92.

Galloa, A., Knox, B.P., Bruno, K.S., Solfrizzo, M., Baker, S.E. & Perrone, G. (2014). Identification and

characterization of the polyketide synthase involved in ochratoxin A biosynthesis in Aspergillus

carbonarius. International Journal of Food Microbiology, 179, 10-17.

Gambacorta, L., Magistà, D., Perrone, G., Murgolo, S., Logrieco, A.F. & Solfrizzo, M. (2018). Co-

occurrence of toxigenic moulds, aflatoxins, ochratoxin A, Fusarium and Alternaria mycotoxins in

fresh sweet peppers (Capsicum annuum) and their processed products. World Mycotoxin Journal,

11, 159-173.

Geisen, R. (1996). Multiplex polymerase chain reaction for the detection of potential aflatoxin and

sterigmatocystin producing fungi. Systematic and Applied Microbiology, 19, 388-392.

Geisen, R., Schmidt-Heydt, M. & Karolewiez, A. (2006). A gene cluster of the ochratoxin A biosynthetic

genes in Penicillium. Mycotoxin Research, 22, 134-141.

Geiser, D.M., Klich, M.A., Frisvad, J.C., Peterson, S.W., Varga, J. & Samson, R.A. (2007). The current

status of species recognition and identification in Aspergillus. Studies in Mycology, 59, 1-10.

Page 50: Mycobiota in Chilean chilli Capsicum annuum L. used for

43

Gherbawy, Y.A., Shebany, Y.M., Hussein, M.A. & Maghraby, T.A. (2015). Molecular detection of

mycobiota and aflatoxin contamination of chili. Archives of Biological Sciences, 67, 223-34.

Gil-Serna, J., González-Salgado, A., González-Jaén, M.T., Vázquez, C. & Patiño, B. (2009a). ITS-based

detection and quantification of Aspergillus ochraceus and Aspergillus westerdijkiae in grapes and

green coffee beans by real-time quantitative PCR. International Journal of Food Microbiology,

131, 162-167.

Gil-Serna, J., Vázquez, C., Sardiñas, N., González-Jaén, M.T. & Patiño, B. (2009b). Discrimination of the

main Ochratoxin A-producing species in Aspergillus section Circumdati by specific PCR assays.

International Journal of Food Microbiology, 136, 83-87.

Glass, L. & Donaldson, G.C. (1995). Development of primer sets designed for use with the PCR to amplify

conserved genes from filamentous ascomycetes. Applied and Environmental Microbiology, 61,

1323-1330.

Gonzalez-Jaen, M.T., Mirete, S., Patino, B., Lopez-Errasquin, E. & Covadonga, V. (2004). Genetic markers

for the analysis of variability and for production of specific diagnostic sequences in fumonisin-

producing strains of Fusarium verticillioides. European Journal of Plant Pathology, 110, 525-532.

González-Salgado, A., González-Jaén, M.T., Vazquez, C. & Patiño, B. (2008). Highly sensitive PCR-based

detection methods specific for Aspergillus flavus in wheat flour. Food Additives & Contaminants,

25, 758-764.

Hall, B.G. (2013). Building phylogenetic trees from molecular data with MEGA. Molecular Biology and

Evolution, 30, 1229-1235.

Ham, H., Kim, S., Kim, M.H., Lee, S., Hong, S.K., Ryu, J.G. & Lee, T. (2016). Mycobiota of ground red

pepper and their aflatoxigenic potential. Journal of Microbiology, 54, 832-837.

Hayat, A., Paniel, N., Rhouati, A., Marty, J.L. & Barthelmebs, L. (2012). Recent advances in ochratoxin

A-producing fungi detection based on PCR methods and ochratoxin A analysis in food matrices.

Food Control, 26, 401-415.

Hebert, P.D.N., Cywinska, A., Ball S.L. & de Waard J.R. (2003). Biological identifications through DNA

barcodes. Proceedings of the Royal Society of London B, 270, 313-321.

Hierro, J.M.H., Garcia-Villanova, J., Torreno, P.R. & Fonseca, I.M.T. (2008). Aflatoxins and ochratoxin A

in red paprika for retail sale in Spain: occurrence and evaluation of a simultaneous analytical

method. Journal of Agricultural and Food Chemistry, 56, 751-756.

Hocking, A.D. & Pitt, J.I. (1980). Dichloran-glycerol medium for enumeration of xerophilic fungi from low

moisture foods. Applied and Environmental Microbiology, 39, 488-492.

Hoorfar, J., Cook, N., Malorny, B., Wagner, M., De Medici, D., Abdulmawjood, A. & Fach, P. (2004).

Diagnostic PCR: making internal amplification control mandatory. Journal of Applied

Microbiology, 96, 221-222.

Hossain, Md.N., Talukder, A., Afroze, F., Rahim, Md.M., Begum, S., Haque, Md.Z. & Ahmed, M.M.

(2018). Identification of aflatoxigenic fungi and detection of their aflatoxin in red chilli (Capsicum

annuum) samples using direct cultural method and HPLC. Advances in Microbiology, 8, 42-53.

Houbraken, J., Frisvad, J.C. & Samson, R.A. (2011). Taxonomy of Penicillium section citrina. Studies in

Mycology, 70, 53-138.

Page 51: Mycobiota in Chilean chilli Capsicum annuum L. used for

44

Houbraken, J., Frisvad, J.C., Seifert, K.A., Overy, D.P., Tuthill, D.M., Valdez, J.G. & Samson, R.A. (2012).

New penicillin-producing Penicillium species and an overview of section Chrysogena. Persoonia,

29, 78-100.

Houbraken, J., Visagie, C.M., Seifert, K.A., et al. (2014). A taxonomic and phylogenetic revision of

Penicillium section Aspergilloides. Studies in Mycology, 78, 373-451.

Houbraken, J., Wang, L., Lee, H.B. & Frisvad, J.C. (2016). New sections in Penicillium containing novel

species producing patulin, pyripyropens or other bioactive compounds. Persoonia, 36, 299-314.

Hu, Y., Zhang, Z. & He, Y. (2006). Determination of aflatoxin in high-pigment content samples by matrix

solid-phase dispersion and high-performance liquid chromatography. Journal of Agricultural and

Food Chemistry, 54, 4126-4130.

IARC. (1993). Some naturally occurring substances: food items and constituents, heterocyclic aromatic

amines and mycotoxins. In: Monographs on the evaluation of the carcinogenic risk of chemicals

to humans. International Agency for Research on Cancer, Lyon, 56, 489-521.

Ikoma, T., Tsuchiya, Y., Asai, T., Okano, K., Endoh, K., Yamamoto, M. & Nakamura, K. (2015). Ochratoxin

contamination of red chili peppers from Chile, Bolivia and Peru showing high incidences of

gallbladder cancer. Asian Pacific Journal of Cancer Prevention, 16, 5897-5991.

INESTAT. (2016). About all core data, CENSO 2016. Avaliable at

http://www.ine.cl/estadisticas/censos/censos-de-poblacion-y-vivienda. Visited 25 June 2017.

Iqbal, S.Z., Paterson, R.R.M., Bhatti, I.A. & Asi, M.R. (2010). Survey of aflatoxins in chilies from Pakistan

produced in rural, semi-rural and urban environments. Food Additives and Contaminants, 3, 268-

274.

Iwen, P.C., Hinrichs, S.H. & Rup, M.E. (2002). Utilization of the internal transcribed spacers regions as

molecular targets to detect and identify human fungal pathogens. Medical Mycology, 40, 87-109.

Kimura, M. (1980). A simple method for estimating evolutionary rate of base substitutions through

comparative studies of nucleotide sequences. Journal of Molecular Evolution, 16, 111-120.

King, A.D., Hocking, A.D. & Pitt. J.I. (1979). Dichloran-rose bengal medium for the enumeration and

isolation of molds from foods. Applied and Environmental Microbiology, 37, 959-964.

Kolecka, A., Khayhan, K., Arabatzis, M., Velegraki, A., Kostrzewa, M., et al. (2014). Efficient identification

of Malassezia yeasts by matrix-assisted laser desorption ionization time of flight mass

spectrometry (MALDI-TOF MS). British Journal of Dermatology, 170, 332-341.

Kolecka, A., Khayhan, K., Groenewald, M., Theelen, B., Arabatzis, M., Velegraki, A., Kostrzewa, M., Mares,

M., Taj-Aldeen, S.J. & Boekhout, T. (2013). Identification of medically relevant species of

arthroconidial yeasts by use of matrix-assisted laser desorption ionization-time of flight mass

spectrometry. Journal of Clinical Microbiology, 51, 2491-2500.

Kumar, S., Stecher, G. & Tamura, K. (2016). MEGA7: molecular evolutionary Genetics analysis version

7.0 for bigger datasets. Molecular Biology and Evolution, 33, 1870-1874.

Lima, N. & Santos, C. (2017). MALDI-TOF MS for identification of food spoilage filamentous fungi. Current

Opinion in Food Science, 13, 26-30.

Magan, N. & Olsen, M. (2004). Risk assessment and risk management of mycotoxins in food. In:

Mycotoxins in Food-Detection and Control. Cambridge: Woodhead Publishing, 1, 3-27.

Mandeel, Q.A. (2005). Fungal contamination of some imported spices. Mycopathologia, 159, 291-298.

Page 52: Mycobiota in Chilean chilli Capsicum annuum L. used for

45

Mannaa, M. & Kim, K.D. (2017). Influence of Temperature and Water Activity on Deleterious Fungi and

Mycotoxin Production during Grain Storage. Mycobiology, 45, 240-254.

Marklein, G., Josten, M., Klanke, U., Muller, E., Horre, R., Maier, T., Wenzel, T., Kostrzewa, M., Bierbaum,

G., Hoerauf, A. & Sahl, H.G. (2009). Matrix-assisted laser desorption ionization-time of flight mass

spectrometry for fast and reliable identification of clinical yeast isolates. Journal of Clinical

Microbiology, 47, 2912-2917.

Martin, A., Aranda, E., Benito, M.J., Pérez-Nevado, F. & Córdoba, M.G. (2005). Identification of fungal

contamination and determination of mycotoxigenic moulds by micellar electrokinetic capillary

chromatography in smoked paprika. Journal of Food Protection, 68, 815-822.

Mitchell, J.I., Roberts, P.J. & Moss, S.T. (1995). Sequence or structure? A short review on the application

of nucleic acid sequence information to fungal taxonomy. Mycologist, 9, 67-75.

Motloung, L., De Saeger, S., De Boevre, M., Detavernier, C., Audenaert, K., Adebo, O.A. & Njobeh, P.B.

(2018). Study on mycotoxin contamination in South African food spices. World Mycotoxin Journal,

In press.

Nelson, P.E., Toussoun, T.A. & Cook, R.J. (1981). Fusarium: Diseases, Biology and Taxonomy. University

Park, PA, USA: Pennsylvania State University Press.

Niessen, L., Schmidt, H., Muhlencoert, E., Farber, P., Karolewiez, A. & Geisen, P. (2005). Advances in

the molecular diagnosis of Ochratoxin A-producing fungi. Food Additives and Contaminants, 22,

324-334.

Nilsson, R.H., Ryberg, M., Kristiansson, E., Abarenkov, K., Larsson, K.H. & Koljalg, U. (2006). Taxonomic

reliability of DNA sequences in public sequence databases: a fungal perspective. PLoS One, 1,

1-59.

Nilsson, R.H., Sánchez-García, M., Ryberg, M., Abarenkov, K., Wurzbacher, C. & Kristiansson, E. (2017).

Read quality-based trimming of the distal ends of public fungal DNA sequences is nowhere near

satisfactory. MycoKeys, 26, 13-24.

O'Callaghan, J., Coghlan, A., Abbas, A., García-Estrada, C., Martín, J.F. & Dobson, A.D.W. (2013).

Functional characterization of the polyketide synthase gene required for ochratoxin A biosynthesis

in Penicillium verrucosum. International Journal of Food Microbiology, 161, 172-181.

O'Donnell, K. (1992). Ribosomal DNA internal transcribed spacers are highly divergent in the

phytopathogenic ascomycete Fusarium sambucinum (Gibberella pulicaris). Current Genetics, 22,

213-220.

Pal, A., Acharya, D., Saha, D., Roy, D. & Dhar T.K. (2005). In situ simple cleanup during immunoassay:

a simple method for rapid detection of aflatoxin B1 in food samples. Journal of Food Protection,

68, 169-2177.

Paradellsa, S., Rocamondea, B., Llinaresa, C., Herranz-Pérez, V., Jimenez, M., Garcia-Verdugo, J.M.,

Zipancic, I., Soria, J.M. & Garcia-Esparza, M.A. (2015). Neurotoxic effects of ochratoxin A on the

subventricular zone of adult mouse brain. Journal of Applied Toxicology, 35, 737-751.

Patel, S., Hazel, C.M., Winterton, A.G.M. & Mortby, E. (1996). Survey of ethnic foods for mycotoxins. Food

Additives and contaminants, 13, 883-841.

Paterson, M. & Lima, N. (2010a). How will climate change affect mycotoxins in food? Food Research

International, 43, 1902-1914.

Page 53: Mycobiota in Chilean chilli Capsicum annuum L. used for

46

Paterson, M. & Lima, N. (2010b). Toxicology of mycotoxins. In: Molecular, Clinical and environmental

toxicology. Luch A (Ed). Clinical Toxicology, 2, 31-63.

Paterson, R.R.M. (2006a). Identification and quantification of mycotoxigenic fungi by PCR. Process

Biochemistry, 41, 1467-1474.

Paterson, R.R.M. (2006b). Primers from the isoepoxydon dehydrogenase gene of the patulin biosynthetic

pathway to indicate critical control points for patulin contamination of apples. Food Control, 17,

741-744.

Patiño, B., González-Salgado, A., González-Jaén, M.T. & Vázquez, C. (2005). PCR detection assays for

the ochratoxin-producing Aspergillus carbonarius and Aspergillus ochraceus species.

International Journal of Food Microbiology, 104, 207-214.

Pavlovic, M., Mewes, A., Maggipinto, M., Schmidt, W., Messelhausser, U., Balsliemke, J., Hormansdorfer,

S., Busch, U. & Huber, I. (2014). MALDI-TOF MS based identification of food-borne yeast isolates.

Journal of Microbiological Methods, 106, 123-128.

Pérez‐Izquierdo, L., Morin, E., Maurice, J.P., Martin, F., Rincón, A. & Buée, M. (2017). A new promising

phylogenetic marker to study the diversity of fungal communities: The Glycoside Hydrolase 63

gene. Molecular Ecology Resource, 17, 1-11.

Perrone, G., Stea, G., Epifani, F., Varga, J., Frisvad J. C. & Samson, R. A. (2011). Aspergillus niger

contains the cryptic phylogenetic species A. awamori. Fungal Biology 115, 1138-1150.

Peterson, S.W. (2008). Phylogenetic analysis of Aspergillus species using DNA sequences from four loci.

Mycologia, 100, 205-226.

Peterson, S.W. (2012). Aspergillus and Penicillium identification using DNA sequences: barcode or MLST?

Applied Microbiology and Biotechnology, 95, 339-344.

Pitt, J.I. & Hocking, A.D. (1997). Fungi and Food Spoilage, 2nd ed., Sydney, Blackie Academic &

Professional.

Pitt, J.I. & Hocking, A.D. (2009a). The Ecology of Fungal Food Spoilage. In: Fungi and Food Spoilage.

Springer, Boston, MA, 3-9.

Pitt, J.I. & Hocking, A.D. (2009b). Fungi and Food Spoilage, third ed. New York: Springer.

Pitt, J.I., Hocking, A.D. & Glenn, D.R. (1983). An improved medium for the detection of Aspergillus flavus

and A. parasiticus. Journal of Applied Bacteriology, 54, 109-114.

Pitt, J.I., Hocking, A.D., Samson, R.A. & King, A.D. (1992). Recommended methods for mycological

examination of food. In: Modern methods for mycology, Samson R.A., Hocking A.D., Pitt J.I. &

King A.D. (Eds.), 1, 343-348.

ProChile. (2016). Ficha De Mercado Producto. El Mercado De Merkén (Ají Seco) En Argentina

2016/Oficina Comercial En Buenos Aires Argentina, 1, 1-2.

Proctor, R.H., Desjardins, A.E., Plattner, R.D. & Hohn, T.M. (1999). A polyketide synthase gene required

for biosynthesis of fumonisin mycotoxins in Gibberella fujikuroi mating population A. Fungal

Genetics and Biology, 27, 100-112.

Quintilla, R., Kolecka, A., Casaregola, S., Daniel, H.M., Houbraken, J., Kostrzewa, M., Boekhout, T. &

Groenewald, M. (2018). MALDI-TOF MS as a tool to identify foodborne yeasts and yeast-like fungi.

International Journal of Food Microbiology, 266, 109-118.

Rico-Muñoz, E., Samson, R.A. & Houbraken, J. (2018). Mould spoilage of foods and beverages: Using

the right methodology, Food Microbiology. In Press.

Page 54: Mycobiota in Chilean chilli Capsicum annuum L. used for

47

Robert, C. (2004). MUSCLE: multiple sequence alignment with high accuracy and high throughput.

Nucleic Acids Research, 32, 1792-1797.

Rodrigues, P., Santos, C., Venâncio, A. & Lima, N. (2011). Species identification of Aspergillus section

Flavi isolates from Portuguese almonds using phenotypic, including MALDI-TOF ICMS, and

molecular approaches. Journal of Applied Microbiology 111, 877-892.

Rodrigues, P., Venâncio, A., Kozakiewicz, Z. & Lima, N. (2009). A polyphasic approach to the identification

of aflatoxigenic and non-aflatoxigenic strains of Aspergillus Section Flavi isolated from Portuguese

almonds. International Journal of Food Microbiology, 129, 187-193.

Roe, A.D., Rice, A.V., Bromilow, S.E., Cooke, J.E. & Sperling, F.A. (2010). Multilocus species identification

and fungal DNA barcoding: insights from blue stain fungal symbionts of the mountain pine beetle.

Molecular Ecology Resources, 10, 946-959.

Ryu, D. & Wolf-Hall, C. (2015). Yeast and molds. in: Salfinger, Y., Tortorello, M.L. (Eds.), Compendium of

Methods for the Microbiological Examination of Foods, fifth edition, American Public Health

Association, Washington D. C., 277-286.

Saitou, N. & Nei, M. (1987). The neighbor-joining method: a new method for reconstructing phylogenetic

trees. Molecular Biology and Evolution, 4, 406-425.

Sakaguci, K., Iizuka, H. & Yamazaki, S. (1951). A study on black Aspergilli. Journal of the Agricultural

Chemical Society of Japan 24, 138-142.

Salari, R., Najafi, H., Boroushaki, M., Mortazavi, S. & Fathi-Najafi, M. (2012). Assessment of the

microbiological quality and mycotoxin contamination of Iranian red pepper spice. Journal of

Agriculture Science and Technology, 14, 1511-1521.

Samson, R.A., Houbraken, J.A.M.P., Kuijpers, A.F.A., Frank, J.M. & Frisvad J.C. (2004b). New ochratoxin

or sclerotium producing species in Aspergillus section Nigri. Studies in Mycology, 50, 45-61.

Samson, R.A., Seifert, K.A., Kuijpers, A.A., Houbraken, J.A.M.P. & Frisvad, J.C. (2004a). Phylogenetic

analysis of Penicillium subgenus Penicillium using partial β-tubulin sequences. Studies in

Mycology, 49, 175-200.

Santos, C., Fraga, M.E., Kozakiewicz, Z. & Lima, N. (2010e). Fourier transform infrared as a powerful

technique for the identification and characterization of filamentous fungi and yeasts. Research in

Microbiology, 161, 168-175.

Santos, C., Paterson, R.R.M., Venâncio, A. & Lima, N. (2010d). Filamentous fungal characterizations by

matrix-assisted laser desorption/ionization time‐of‐flight mass spectrometry. Journal of Applied

Microbiology, 108, 375-385.

Santos, C., Ventura, J.A., Costa, H., Fernandes, P.M.B. & Lima N. (2015). MALDI-TOF MS to identify the

pineapple pathogen Fusarium guttiforme and its antagonist Trichoderma asperellum on decayed

pineapple. Tropical Plant Pathology, 40, 227-232.

Santos, I., Venâncio, A. & Lima, N. (1998). Fungos Contaminantes na Indústria Alimentar. Micoteca da

Universidade do Minho, Braga, 1, 1-128.

Santos, L., Kasper, R., Gil-Serna, J., Marín, S., Sanchis, V. & Ramos, A.J. (2010a). Effect of Capsicum

carotenoids on growth and ochatoxin A production by chilli and paprika Aspergillus spp. Isolates.

International Journal of Food and Microbiology, 142, 354-359.

Page 55: Mycobiota in Chilean chilli Capsicum annuum L. used for

48

Santos, L., Kasper, R., Sardiñas, N., Marín, S., Sanchis, V. & Ramos, A. J. (2010b). Effect of Capsicum

carotenoids on growth and aflatoxins production by Aspergillus flavus isolated from paprika and

chilli. Food Mycrobiology, 27, 1064-1070.

Santos, L., Marín S., Sanchis, V. & Ramos, A.J. (2008). Capsicum and mycotoxin contamination: state of

the art in a global context. Food Science and Technology International, 14, 5-20.

Santos, L., Marín, S., Mateo, E.M., Gil-Serna, J., Valle-Algarra, F.M., Patiño, B. & Ramos, A.J. (2011).

Mycobiota and co-occurrence of mycotoxins in Capsicum powder. International Journal of Food

Microbiology, 151, 270-276.

Santos, L., Marín, S., Sanchis, V. & Ramos, A.J. (2010c). Co-occurrence of aflatoxins, ochratoxin A and

zearalenone in Capsicum powder samples available on the Spanish market. Food Chemistry,

122, 826-830.

Schoch, C. L., Seifert, K. A., Huhndorf, S., Robert, V., Spouge, J. L., Levesque, C. A., Chen, W. & Fungal

Barcoding Consortium. (2012). Nuclear ribosomal internal transcribed spacer (ITS) region as a

universal DNA barcode marker for Fungi. Proceedings of the National Academy of Sciences, 109,

6241-6246.

Schweiggert, U., Mix, K., Schieber, A. & Carle, R. (2005). An innovative process for the production of

spices through immediate thermal treatment of the plant material. Innovative Food Science &

Emerging Technologies, 6, 143-153.

Shapira, R., Paster, N., Eyal, O., Menasherov, M., Mett, A. & Salomon, R. (1996). Detection of

aflatoxingenic molds in grains by PCR. Applied and Environmental Microbiology, 62, 3270-3273.

Silva, F.C., Chalfoun, S.M., Batista, L.R., Santos, C. & Lima, N. (2015). Use of polyphasic approach

including MALDI-TOF MS for identification of Aspergillus section Flavi strains isolated from food

commodities in Brazil. Annals of Microbiology, 65, 2119-2129.

Silva, W.L. & Singh, R. (2012). First Report of Alternaria alternata Causing Leaf Spot on Aloe vera in

Louisiana. Plant Disease Journal, 96, 1379.

Simões, M.F., Pereira, L., Santos, C. & Lima, N. (2013). Polyphasic identification and preservation of

fungal diversity: concepts and applications. In: Management of Microbial Resources in the

Environment. Malik, A., Grohmann, E., Alves, M. (Eds.), Springer, Dordrecht, 91-117.

Skouboe, P., Frisvad, J.C., Lauritsen, D., Boysen, M., Taylor, J.W. & Rossen, L. (1999). Nucleotide

sequences from the ITS region of Penicillium species. Mycological Research, 103, 873-881.

Smith, J.E., Lewis, C.W., Anderson, J.G. & Solomons, G.L. (1994). Mycotoxins in human nutrition and

health. European Commission: Agro-Industrial Research Division, 3, 202-222.

Soares, C., Rodrigues, P., Freitas-Silva, O., Abrunhosa, L. & Venâncio, A. (2010). HPLC method for

simultaneous detection of aflatoxins and cyclopiazonic acid. World Mycotoxin Journal, 3, 225-

231.

Soares, C., Rodrigues, P., Peterson, S.W., Lima, N. & Venâncio, A. (2012). Three new species of

Aspergillus section Flavi isolated from almonds and maize in Portugal. Mycologia, 104, 682-697.

Summerbell, R.C.G., Green, B.J., Corr, D. & Scott, J.A. (2011). Molecular methods for bioaerosol

characterization. In: Microorganisms in home and indoor work environments. Flannigan, D.S.,

R.A.; Miller, J.D. (Ed.) CRC Press, Boca Raton. 247-263.

Thom C. & Church M.B. (1926). The Aspergilli. Baltimore, MD: Williams & Wilkins. 39.

Page 56: Mycobiota in Chilean chilli Capsicum annuum L. used for

49

Van de Perre, E., Deschuyffeleer, N., Jacxsens, L., Vekeman, F., Van der Hauwaert, W., Asam, S., Rychlik,

M., Devlieghere, F. & De Meulenaer, B. (2014). Screening of moulds and mycotoxins in tomatoes,

bell peppers, onions, soft red fruits and derived tomato products. Food Control, 37, 165-170.

Varga, J., Frisvad, J.C., Kocsubé, S., Brankovics, B., Tóth, B., Szigeti, G. & Samson, R.A. (2011). New

and revisited species in Aspergillus section Nigri. Studies in Mycology, 69, 1-17.

Visagie, C.M., Hirooka, Y., Tanney, J.B., Whitfield, E., Mwange, K., Meijer, M., Amed, A.S., Seifert, K.A.

& Samson, R.A. (2014a). Aspergillus, Penicillium and Talaromyces isolated from house dust

samples collected around the world. Studies in Mycology 78, 63-139.

Visagie, C.M., Houbraken, J., Frisvad, J.C., Hong, S.-B., Klaassen, C.H.W., Perrone, G., Seifert, K.A.,

Varga, J., Yaguchi, T. & Samson, R.A. (2014b). Identification and nomenclature of the genus

Penicillium. Studies in Mycology, 78, 343-371.

White, T.J., Burns, T., Lee, S. & Taylor, J.W. (1990). Amplification and direct sequencing of fungal

ribosomal RNA genes for phylogenetics. In: PCR Protocols: A Guide to Methods and Applications.

Innis, M.A., Gelgard, D.H., Sninsky, J.J., White, T.J. (Eds.), Academic Press, New York, 315-322.

Zhang, J., Zhu, L., Chen, H., Li, M., Zhu, X., Gao, Q., Wang, D. & Zhang, Y. (2016). A Polyketide synthase

encoded by the gene An15g07920 is involved in the biosynthesis of ochratoxin A in Aspergillus

niger. Journal of Agricultural and Food Chemistry, 64, 9680-9688.

Zhang, L., Dou, X.W., Zhang, C., Logrieco, A.F. & Yang, M.H. (2018). A Review of Current Methods for

Analysis of Mycotoxins in Herbal Medicines. Toxins, 10, 65.

Page 57: Mycobiota in Chilean chilli Capsicum annuum L. used for

50

ANNEX I – ISOLATED STRAINS

Table A-I. Sampling point, isolated media, gene sequencing, fungal Identification and CCCT/UFRO code

assigned to all strain identified at specie level using gene sequencing.

Sampling Isolated Media

Gene Fungal ID CCCT/UFRO Code

Point sequencing 3 MEA BenA Aspergillus niger CCCT 18.11

2 DG18 BenA Aspergillus niger CCCT 18.12

3 MEA BenA Aspergillus niger CCCT 18.14

3 DG18 BenA Aspergillus niger CCCT 18.15

3 DRBC BenA Aspergillus awamori CCCT 18.16

2 DG18 BenA Alternaria atra CCCT 18.17

2 MEA BenA Aspergillus awamori CCCT 18.18

2 DG18 BenA Aspergillus niger CCCT 18.19

2 DRBC BenA Aspergillus awamori CCCT 18.20

3 DRBC BenA Aspergillus tubingensis CCCT 18.21

3 MEA BenA Aspergillus niger CCCT 18.22

2 MEA BenA Aspergillus brasiliensis CCCT 18.23

3 DG18 BenA Aspergillus niger CCCT 18.26

2 DG18 BenA Aspergillus niger CCCT 18.27

2 MEA BenA Aspergillus niger CCCT 18.28

3 MEA BenA Penicillium polonicum CCCT 18.29

2 DG18 BenA Aspergillus niger CCCT 18.30

3 DG18 BenA Penicillium brevicompactum CCCT 18.31

3 DRBC BenA Penicillium glabrum CCCT 18.32

3 DRBC BenA Penicillium verrucosum CCCT 18.33

3 DG18 BenA Aspergillus oryzae CCCT 18.35

3 MEA BenA Aspergillus flavus CCCT 18.36

3 DRBC BenA Penicillium melanoconidium CCCT 18.37

3 DRBC BenA Penicillium bialowiezense CCCT 18.38

3 DRBC BenA Penicillium brevicompactum CCCT 18.39

3 DRBC BenA Penicillium paraherquei CCCT 18.40

3 DG18 BenA Penicillium polonicum CCCT 18.42

3 MEA BenA Penicillium viridicatum CCCT 18.43

3 DRBC BenA Penicillium viridicatum CCCT 18.44

3 MEA BenA Penicillium crustosum CCCT 18.46

3 DRBC BenA Penicillium glabrum CCCT 18.47

3 MEA BenA Penicillium brevicompactum CCCT 18.49

3 MEA BenA Aspergillus awamori CCCT 18.50

3 DRBC BenA Penicillium glabrum CCCT 18.51

3 DG18 BenA Penicillium crustosum CCCT 18.52

3 DRBC BenA Aspergillus awamori CCCT 18.53

3 DG18 BenA Penicillium melanoconidium CCCT 18.54

3 MEA BenA Penicillium polonicum CCCT 18.55

3 DRBC BenA Penicillium paraherquei CCCT 18.56

3 MEA BenA Penicillium brevicompactum CCCT 18.57

Page 58: Mycobiota in Chilean chilli Capsicum annuum L. used for

51

Table A-I. Continued

Sampling Point

Isolated Media Gene

sequencing Fungal ID CCCT/UFRO Code

3 DRBC BenA Penicillium brevicompactum CCCT 18.58

3 DG18 BenA Penicillium brevicompactum CCCT 18.59

3 DG18 BenA Penicillium cyclopium CCCT 18.60

3 MEA BenA Penicillium glabrum CCCT 18.61

3 MEA BenA Penicillium crustosum CCCT 18.62

3 DG18 BenA Penicillium glabrum CCCT 18.63

3 MEA BenA Penicillium cyclopium CCCT 18.64

3 DRBC BenA Penicillium waksmanii CCCT 18.65

3 DG18 BenA Penicillium cyclopium CCCT 18.66

3 DG18 BenA Penicillium brevicompactum CCCT 18.67

3 DRBC BenA Penicillium brevicompactum CCCT 18.68

3 MEA BenA Aspergillus luchuensis CCCT 18.70

3 DRBC BenA Penicillium cyclopium CCCT 18.71

3 DRBC BenA Penicillium polonicum CCCT 18.72

3 DRBC BenA Microascus cinereus CCCT 18.73

3 DG18 BenA Penicillium brevicompactum CCCT 18.74

3 DRBC BenA Penicillium glabrum CCCT 18.75

3 DG18 BenA Penicillium glabrum CCCT 18.76

3 DRBC BenA Penicillium crustosum CCCT 18.77

3 DRBC BenA Penicillium glabrum CCCT 18.78

3 DRBC BenA Penicillium polonicum CCCT 18.80

3 DRBC BenA Penicillium cyclopium CCCT 18.81

3 DRBC BenA Penicillium crustosum CCCT 18.82

3 DG18 BenA Penicillium cyclopium CCCT 18.83

3 DRBC BenA Penicillium cyclopium CCCT 18.84

3 DG18 BenA Aspergillus pseudoglaucus CCCT 18.85

3 DRBC BenA Penicillium glabrum CCCT 18.86

3 MEA BenA Penicillium glabrum CCCT 18.87

3 DG18 BenA Penicillium polonicum CCCT 18.88

2 MEA BenA Aspergillus niger CCCT 18.89

3 MEA BenA Penicillium CCCT 18.90

3 DG18 BenA Penicillium discolor CCCT 18.91

3 DRBC BenA Aspergillus sydowii CCCT 18.92

3 MEA BenA Penicillium glabrum CCCT 18.93

3 MEA BenA Aspergillus fumigatus CCCT 18.94

3 DRBC BenA Penicillium viridicatum CCCT 18.95

2 DRBC BenA Penicillium crustosum CCCT 18.96

3 MEA BenA Aspergillus pseudoglaucus CCCT 18.97

3 DRBC BenA Trichoderma viride CCCT 18.98

3 DRBC BenA Penicillium crustosum CCCT 18.99

3 MEA BenA Penicillium glabrum CCCT 18.100

3 DRBC BenA Penicillium glabrum CCCT 18.101

3 MEA BenA Penicillium crustosum CCCT 18.102

3 DG18 BenA Aspergillus fumigatus CCCT 18.103

3 DRBC BenA Penicillium brevicompactum CCCT 18.104

Page 59: Mycobiota in Chilean chilli Capsicum annuum L. used for

52

Table A-I. Continued

Sampling Point

Isolated Media Gene

sequencing Fungal ID CCCT/UFRO Code

3 DG18 BenA Penicillium brevicompactum CCCT 18.105

3 DRBC BenA Penicillium cyclopium CCCT 18.106

3 DRBC BenA Penicillium buchwaldii CCCT 18.107

3 MEA BenA Penicillium adametzioides CCCT 18.108

3 MEA BenA Penicillium angulare CCCT 18.109

3 DRBC BenA Penicillium angulare CCCT 18.110

3 DG18 BenA Penicillium viridicatum CCCT 18.111

3 DRBC BenA Penicillium brevicompactum CCCT 18.113

2 DG18 BenA Penicillium glabrum CCCT 18.115

2 DG18 BenA Penicillium expansum CCCT 18.116

2 DG18 BenA Penicillium crustosum CCCT 18.117

2 DRBC BenA Penicillium glabrum CCCT 18.118

2 DRBC BenA Penicillium crustosum CCCT 18.119

3 DRBC BenA Aspergillus flavus CCCT 18.120

2 DRBC BenA Penicillium glabrum CCCT 18.122

2 DG18 BenA Penicillium crustosum CCCT 18.123

2 MEA BenA Penicillium cyclopium CCCT 18.124

2 DRBC BenA Penicillium citrinum CCCT 18.125

2 DG18 BenA Aspergillus flavus CCCT 18.126

2 MEA BenA Aspergillus flavus CCCT 18.128

2 DG18 BenA Penicillium brevicompactum CCCT 18.129

2 DG18 BenA Penicillium brevicompactum CCCT 18.130

2 DG18 BenA Penicillium expansum CCCT 18.131

2 DG18 BenA Penicillium crustosum CCCT 18.132

2 MEA BenA Penicillium crustosum CCCT 18.134

2 MEA BenA Penicillium glabrum CCCT 18.135

1 DG18 BenA Alternaria botrytis CCCT 18.136

2 DG18 BenA Penicillium glabrum CCCT 18.137

2 DG18 BenA Penicillium crustosum CCCT 18.138

1 DG18 BenA Penicillium brevicompactum CCCT 18.139

2 DG18 BenA Alternaria botrytis CCCT 18.140

2 DRBC ITS1-5.8S-ITS2 Penicillium commune CCCT 18.141

2 DRBC ITS1-5.8S-ITS2 Penicillium crustosum CCCT 18.142

2 DG18 ITS1-5.8S-ITS2 Penicillium polonicum CCCT 18.143

1 DRBC BenA Penicillium glabrum CCCT 18.144

1 DRBC BenA Aspergillus fructus CCCT 18.145

1 DRBC ITS1-5.8S-ITS2 Cladosporium westerdijkieae CCCT 18.146

1 DG18 ITS1-5.8S-ITS2 Alternaria alternata CCCT 18.147

1 DG18 ITS1-5.8S-ITS2 Penicillium steckii CCCT 18.148

1 DRBC ITS1-5.8S-ITS2 Penicillium viridicatum CCCT 18.149

2 MEA ITS1-5.8S-ITS2 Penicillium polonicum CCCT 18.150

1 DG18 ITS1-5.8S-ITS2 Cladosporium westerdijkieae CCCT 18.151

1 DG18 ITS1-5.8S-ITS2 Cladosporium westerdijkieae CCCT 18.152

1 DG18 ITS1-5.8S-ITS2 Penicillium brevicompactum CCCT 18.154

1 DRBC ITS1-5.8S-ITS2 Penicillium viridicatum CCCT 18.155

Page 60: Mycobiota in Chilean chilli Capsicum annuum L. used for

53

Table A-I. Continued

Sampling Point

Isolated Media Gene

sequencing Fungal ID CCCT/UFRO Code

1 DRBC ITS1-5.8S-ITS2 Penicillium polonicum CCCT 18.156

2 DG18 ITS1-5.8S-ITS2 Aspergillus flavus CCCT 18.157

2 DG18 ITS1-5.8S-ITS2 Cladosporium westerdijkieae CCCT 18.158

2 DG18 ITS1-5.8S-ITS2 Penicillium viridicatum CCCT 18.159

2 DG18 ITS1-5.8S-ITS2 Penicillium glabrum CCCT 18.160

1 DRBC ITS1-5.8S-ITS2 Penicillium polonicum CCCT 18.161

1 DRBC ITS1-5.8S-ITS2 Fusarium equiseti CCCT 18.162

1 DRBC ITS1-5.8S-ITS2 Trichoderma koningiopsis CCCT 18.163

1 DRBC ITS1-5.8S-ITS2 Penicillium freii CCCT 18.164

1 DRBC ITS1-5.8S-ITS2 Fusarium oxysporum CCCT 18.165

1 DG18 BenA Fusarium incarnatum CCCT 18.166

1 DRBC ITS1-5.8S-ITS2 Fusarium redolens CCCT 18.167

1 DG18 BenA Penicillium brevicompactum CCCT 18.168

1 DRBC ITS1-5.8S-ITS2 Alternaria alternata CCCT 18.169

1 DG18 BenA Penicillium chrysogenum CCCT 18.170

1 DG18 ITS1-5.8S-ITS2 Alternaria alternata CCCT 18.171

1 DG18 ITS1-5.8S-ITS2 Alternaria alternata CCCT 18.172

1 DG18 BenA Fusarium equiseti CCCT 18.173

1 DRBC ITS1-5.8S-ITS2 Fusarium equiseti CCCT 18.174

1 DRBC ITS1-5.8S-ITS2 Penicillium freii CCCT 18.175

1 DG18 ITS1-5.8S-ITS2 Penicillium dipodomyicola CCCT 18.176

1 DG18 ITS1-5.8S-ITS2 Aspergillus versicolor CCCT 18.177

1 DRBC ITS1-5.8S-ITS2 Trichoderma koningiopsis CCCT 18.178

1 DRBC ITS1-5.8S-ITS2 Penicillium viridicatum CCCT 18.179

1 DG18 BenA Boeremia exigua CCCT 18.180

1 DG18 ITS1-5.8S-ITS2 Alternaria alternata CCCT 18.181

1 DRBC BenA Aspergillus fructus CCCT 18.182

1 DG18 ITS1-5.8S-ITS2 Fusarium redolens CCCT 18.183

1 DRBC ITS1-5.8S-ITS2 Aspergillus versicolor CCCT 18.185

1 DG18 BenA Aspergillus flavus CCCT 18.186

1 DRBC ITS1-5.8S-ITS2 Penicillium glabrum CCCT 18.187

2 DRBC BenA Trichurus spiralis CCCT 18.188

2 MEA ITS1-5.8S-ITS2 Alternaria alternata CCCT 18.189

2 DG18 ITS1-5.8S-ITS2 Alternaria alternata CCCT 18.190

1 DRBC ITS1-5.8S-ITS2 Alternaria alternata CCCT 18.191

1 DRBC BenA Penicillium viridicatum CCCT 18.192

1 DG18 ITS1-5.8S-ITS2 Fusarium oxysporum CCCT 18.193

1 DRBC BenA Penicillium brevicompactum CCCT 18.194

1 DG18 ITS1-5.8S-ITS2 Fusarium oxysporum CCCT 18.195

1 DRBC BenA Penicillium viridicatum CCCT 18.196

1 DG18 ITS1-5.8S-ITS2 Penicillium chrysogenum CCCT 18.197

1 DG18 ITS1-5.8S-ITS2 Fusarium oxysporum CCCT 18.198

1 DRBC ITS1-5.8S-ITS2 Penicillium brevicompactum CCCT 18.199

1 DRBC BenA Penicillium chrysogenum CCCT 18.200

1 DRBC ITS1-5.8S-ITS2 Alternaria alternata CCCT 18.201

Page 61: Mycobiota in Chilean chilli Capsicum annuum L. used for

54

Table A-I. Continued

Sampling Point

Isolated Media Gene

sequencing Fungal ID CCCT/UFRO Code

1 DRBC ITS1-5.8S-ITS2 Aspergillus dimorphicus CCCT 18.202

1 DG18 BenA Penicillium cyclopium CCCT 18.203

1 DRBC BenA Penicillium viridicatum CCCT 18.204

1 DRBC ITS1-5.8S-ITS2 Aspergillus dimorphicus CCCT 18.205

1 DG18 ITS1-5.8S-ITS2 Fusarium oxysporum CCCT 18.206

1 DRBC BenA Penicillium chrysogenum CCCT 18.207

1 DG18 BenA Fusarium incarnatum CCCT 18.208

1 DG18 ITS1-5.8S-ITS2 Fusarium equiseti CCCT 18.209

1 DG18 ITS1-5.8S-ITS2 Alternaria alternata CCCT 18.210

2 MEA ITS1-5.8S-ITS2 Alternaria alternata CCCT 18.211

2 MEA ITS1-5.8S-ITS2 Alternaria alternata CCCT 18.212

2 MEA ITS1-5.8S-ITS2 Alternaria alternata CCCT 18.213

2 DG18 ITS1-5.8S-ITS2 Colletotrichum coccodes CCCT 18.214

1 DG18 BenA Penicillium cyclopium CCCT 18.215

1 DG18 BenA Fusarium incarnatum CCCT 18.216

Page 62: Mycobiota in Chilean chilli Capsicum annuum L. used for

55

ANNEX II – STANDARD CURVES OF OTA AND AFLS

Figure A-II. Calibration curve of Ochratoxin A (OTA).

Figure B-II. Calibration curve of Aflatoxin G2.

y = 1075,8x + 736,3R² = 0,9972

0

20000

40000

60000

80000

100000

120000

0 20 40 60 80 100 120

Are

a (m

V/m

in)

ng/mL

OTA

y = 2010,6x + 324,81R² = 0,9993

0,00

5000,00

10000,00

15000,00

20000,00

25000,00

30000,00

35000,00

40000,00

45000,00

0 5 10 15 20 25

Are

a (m

V/m

in)

ng/mL

AFG2

Page 63: Mycobiota in Chilean chilli Capsicum annuum L. used for

56

Figure C-II. Calibration curve of Aflatoxin B2.

Figure D-II. Calibration curve of Aflatoxin B1.

y = 1887,3x + 320,1R² = 0,9994

0,00

5000,00

10000,00

15000,00

20000,00

25000,00

30000,00

35000,00

40000,00

0 5 10 15 20 25

Are

a (m

V/m

in)

ng/mL

AFB1

y = 2777,1x + 442,73R² = 0,9994

0,00

10000,00

20000,00

30000,00

40000,00

50000,00

60000,00

0 5 10 15 20 25

Are

a (m

V/m

in)

ng/mL

AFB2

Page 64: Mycobiota in Chilean chilli Capsicum annuum L. used for

57

Figure E-II. Calibration curve of Aflatoxin G1.

y = 1038,4x + 211,16R² = 0,999

0,00

5000,00

10000,00

15000,00

20000,00

25000,00

0 5 10 15 20 25

Are

a (m

V/m

in)

ng/mL

AFG1