CANDIDA IN UPPER DIGESTIVE TRACT CARCINOGENESIS Mikko Nieminen ACADEMIC DISSERTATION

CANDIDA YEASTS AND CARCINOGENIC ACETALDEHYDE
IN UPPER DIGESTIVE TRACT CARCINOGENESIS
Mikko Nieminen
ACADEMIC DISSERTATION
Research Unit on Acetaldehyde and Cancer, Biomedicum
Department of Bacteriology and Immunology, Haartman Institute
Department of Oral and Maxillofacial Diseases, Helsinki University Central Hospital
Department of Periodontology, Institute of Dentistry
University of Helsinki, Finland
Respiratory Research Group, Institute of Inflammation and Repair
Education and Research Centre, University Hospital of Southern Manchester
University of Manchester, United Kingdom
Finnish Doctoral Programme in Oral Sciences (FINDOS)
To be presented, with the permission of the Faculty of Medicine of the
University of Helsinki, for public examination in lecture hall 2, Biomedicum 1,
th
Haartmaninkatu 8, Helsinki, on 24 October 2014, at 12 o’clock noon.
© 2014 by Mikko Nieminen
Cover image by Dr Ranjith Rajendran, University of Glasgow
(SEM of Candida albicans SC5314 biofilm)
ISBN 978-951-51-0131-0 (pbk.)
ISBN 978-951-51-0132-7 (PDF)
ISSN 2342-3161 (print)
ISSN 2342-317X (online)
http://ethesis.helsinki.fi
Hansaprint Oy
Helsinki 2014
2
SUPERVISED BY:
Riina Rautemaa-Richardson
D.D.S., PhD, FRCPath, Docent
Department of Bacteriology and Immunology, Haartman Institute
Department of Oral Medicine, Institute of Dentistry
University of Helsinki, Helsinki, Finland
Faculty of Human and Medical Sciences, Institute of Inflammation and Repair
University Hospital of South, Wythenshawe Hospital
The University of Manchester, Manchester, United Kingdom
Timo Sorsa
Professor, D.D.S., PhD, Dipl. Perio
Department of Periodontology, Institute of Dentistry
University of Helsinki, Helsinki, Finland
Division of Periodontology, Department of Dental Medicine
Karolinska Institutet, Huddinge, Sweden
Mikko Salaspuro
Professor (emeritus), M.D., PhD
Research Unit on Acetaldehyde and Cancer, Biomedicum
Faculty of Medicine, University of Helsinki, Helsinki, Finland
REVIEWED BY:
Markus Haapasalo
Professor, D.D.S., PhD
Department of Oral Biological & Medical Sciences
University of British Columbia, Vancouver, Canada
Bernhard Hube
Professor, PhD
Department of Microbial Pathogenicity Mechanisms
Leibniz Institute for Natural Product Research and Infection Biology
Hans Knöll Institute
Friedrich Schiller University, Jena, Germany
OPPOSED BY:
Neil A.R. Gow
Professor, PhD, FRSE, FSB, FAAM
School of Medical Sciences, Institute of Medical Sciences
University of Aberdeen, Aberdeen, United Kingdom
3
“Don’t play what’s there, play what’s not there.”
Miles Davis
4
LIST OF CONTENTS
1 LIST OF ORIGINAL PUBLICATIONS ......................................................................... 8
2 ABBREVIATIONS ................................................................................................... 9
3 ABSTRACT ........................................................................................................... 11
4 INTRODUCTION .................................................................................................. 13
5 REVIEW OF THE LITERATURE ............................................................................... 16
5.1 CANDIDA YEASTS AND THE HUMAN MYCOBIOME ........................................ 16
5.2 PATHOGENESIS OF CANDIDA YEASTS ............................................................. 18
5.2.1 Virulence factors..................................................................................... 18
5.2.1.1 Polymorphism and phenotypic switching ................................... 18
5.2.1.2 Biofilm lifestyle ............................................................................ 19
5.2.1.3 Adhesion, invasion and hydrolysis .............................................. 21
5.2.2 Infection and disease .............................................................................. 22
5.2.2.1 Oral candidosis ............................................................................ 23
5.2.2.2 Oesophageal candidosis .............................................................. 24
5.2.2.3 Invasive candidosis ...................................................................... 24
5.2.3 Interaction with bacteria ........................................................................ 25
5.2.4 Mucosal immunity .................................................................................. 25
5.2.4.1 Innate immunity .......................................................................... 26
5.2.4.2 Adaptive immunity ...................................................................... 27
5.3 TREATMENT STRATEGIES ............................................................................... 28
5.3.1 Treatment of infection ........................................................................... 29
5.3.1.1 Topical treatment ........................................................................ 29
5.3.1.1.1 Polyenes ........................................................................... 29
5.3.1.1.2 Azoles ............................................................................... 30
5.3.1.1.3 Chlorhexidine ................................................................... 30
5.3.1.2 Systemic treatment ..................................................................... 30
5.3.1.2.1 Triazoles ........................................................................... 31
5.3.1.2.2 Echinocandins .................................................................. 31
5.3.1.2.3 Polyenes ........................................................................... 32
5.3.1.2.4 Antifungal vaccines .......................................................... 32
5.3.2 Suppressive therapy ............................................................................... 32
5.3.3 Probiotics as an alternative approach .................................................... 33
5.3.4 Alternative anti-biofilm agents ............................................................... 34
5.4 ETHANOL AND ACETALDEHYDE METABOLISM IN THE HUMAN BODY ........... 34
5.4.1 Ethanol oxidation in humans .................................................................. 36
5.4.1.1 Alcohol dehydrogenase (ADH) .................................................... 36
5.4.1.2 Microsomal ethanol oxidising system (MEOS) ............................ 36
5.4.1.3 Aldehyde dehydrogenase (ALDH) ............................................... 37
5.4.2 Hepatic metabolism ............................................................................... 37
5.4.3 Extra-hepatic metabolism ...................................................................... 38
5.4.3.1 Metabolism in the upper digestive tract ..................................... 38
5.4.3.2 Microbial metabolism ................................................................. 39
5
5.4.3.2.1 Bacterial metabolism ....................................................... 39
5.4.3.2.2 Candida metabolism ........................................................ 40
5.5 ACETALDEHYDE IN CARCINOGENESIS OF THE UPPER DIGESTIVE TRACT ....... 44
5.5.1 Ethanol-related toxicity .......................................................................... 44
5.5.2 Acetaldehyde-related toxicity and carcinogenicity ................................ 44
5.5.3 Aetiology and epidemiology of upper digestive tract cancers ............... 46
5.5.3.1 Main risk factors .......................................................................... 47
5.5.3.2 Candida and cancer ..................................................................... 48
6 AIMS OF THE STUDY ........................................................................................... 49
7 MATERIALS AND METHODS ................................................................................ 50
7.1 STUDY DESIGN ................................................................................................ 50
7.2 MATERIALS ..................................................................................................... 51
7.2.1 Candida strains and isolates ................................................................... 51
7.2.2 Reagents ................................................................................................. 52
7.2.2.1 Starter cultures............................................................................ 53
7.2.2.2 Primers ........................................................................................ 53
7.2.2.3 Antibodies ................................................................................... 54
7.3 METHODS ....................................................................................................... 55
7.3.1 Candida growth conditions .................................................................... 55
7.3.1.1 Planktonic cells ............................................................................ 55
7.3.1.2 Biofilm formation ........................................................................ 55
7.3.1.2.1 In vitro biofilm .................................................................. 55
7.3.1.2.2 Biofilm susceptibility to D,L-2-hydroxyisocaproic acid
(HICA)
......................................................................................... 56
7.3.1.2.3 In vivo biofilm ................................................................... 56
7.3.2 Fermentation using starter cultures ....................................................... 56
7.3.3 Isolation and identification of the microbiota in starter cultures .......... 57
7.3.3.1 Assessment of microbial growth and isolation of microorganisms ..................................................................................................... 57
7.3.3.2 DNA extraction ............................................................................ 57
7.3.3.3 16S and 18S sequencing and data analysis ................................. 58
7.3.4 Gas chromatography .............................................................................. 58
7.3.4.1 Planktonic cells and biofilms ....................................................... 58
7.3.4.2 Starter cultures and fermentation .............................................. 59
7.3.5 Analysis of Adh activity ........................................................................... 60
7.3.6 Evaluation of Candida albicans biofilm formation ................................. 60
7.3.6.1 XTT reduction assay .................................................................... 60
7.3.6.2 PicoGreen assay .......................................................................... 61
7.3.6.3 Scanning electron microscopy (SEM) .......................................... 61
7.3.7 Gene expression analysis........................................................................ 61
7.3.7.1 RNA extraction ............................................................................ 61
7.3.7.2 cDNA synthesis ............................................................................ 62
7.3.7.3 RT-qPCR analysis ......................................................................... 62
7.3.8 In vivo biofilm murine model .................................................................. 62
7.3.8.1 Ethics statement ......................................................................... 62
6
7.3.8.2 Diffusion chamber model ............................................................ 62
7.3.9 Immunohistochemistry .......................................................................... 63
7.3.10 Statistical analysis ................................................................................... 64
8 RESULTS .............................................................................................................. 65
8.1 PRODUCTION OF ACETALDEHYDE AND ETHANOL BY CANDIDA YEASTS IN
VITRO ..................................................................................................................... 65
8.1.1 Planktonic Candida albicans and non-albicans Candida cultures .......... 65
8.1.1.1 Acetaldehyde production in ethanol incubation ........................ 65
8.1.1.2 Acetaldehyde production in glucose, fructose or xylitol
incubation or in co-incubation with ethanol................................................ 66
8.1.1.3 Adh activity.................................................................................. 67
8.1.2 Mixed bacteria/yeast starter culture during fermentation .................... 67
8.1.2.1 Pilot experiments ........................................................................ 67
8.1.2.2 Distribution of yeast and bacterial species in starter cultures .... 68
8.1.2.3 Changes in microbial counts and pH during fermentation ......... 69
8.1.2.4 Acetaldehyde and ethanol production during fermentation
and correlation with microbiota .................................................................. 69
8.1.3 Candida albicans biofilms ....................................................................... 69
8.1.3.1 Pilot experiments ........................................................................ 69
8.1.3.2 Acetaldehyde production from ethanol and glucose at an
acidic and a neutral pH ................................................................................ 70
8.2 INHIBITORY EFFECT OF D,L-2-HYDROXYISOCAPROIC ACID (HICA) ON
CANDIDA ALBICANS BIOFILM FORMATION AND ACETALDEHYDE METABOLISM
IN VITRO .................................................................................................................. 71
8.2.1 Pilot experiment - HICA susceptibility .................................................... 71
8.2.2 Effect of HICA on metabolic activity, biomass and ultrastructure of
biofilms .............................................................................................................. 72
8.2.3 Acetaldehyde production during challenge with HICA and
caspofungin ....................................................................................................... 73
8.2.4 Transcriptional changes (CaADH1, CaADH2, CaALD4, CaALD5,
CaALD6, CaACS1 and CaACS2) in biofilms ......................................................... 74
8.2.4.1 Basal and relative gene expression ............................................. 74
8.2.4.2 Correlation with acetaldehyde levels .......................................... 75
8.3 ANTI-BIOFILM AND ANTI-INFLAMMATORY PROPERTIES OF HICA IN VIVO .... 75
9 DISCUSSION ........................................................................................................ 77
9.1 PRINCIPLE FINDINGS....................................................................................... 77
9.2 LIMITATIONS OF THE STUDY .......................................................................... 81
9.3 STRENGTHS OF THE STUDY ............................................................................ 83
9.4 RELEVANCE OF THE STUDY ............................................................................. 84
10 CONCLUSIONS .................................................................................................... 85
11 ACKNOWLEDGEMENTS ....................................................................................... 87
12 REFERENCES........................................................................................................ 90
7
1
LIST OF ORIGINAL PUBLICATIONS
This thesis is based on the following original articles, which are referred to in the text
by Roman numerals as indicated below:
I
Nieminen MT, Uittamo J, Salaspuro M, Rautemaa R. Acetaldehyde production
from ethanol and glucose by non-Candida albicans yeasts in vitro. Oral
Oncology 2009; 45(12): e245-8.
II
Uittamo J, Nieminen MT, Kaihovaara P, Bowyer P, Salaspuro M, Rautemaa R.
Xylitol inhibits carcinogenic acetaldehyde production by Candida species.
International Journal of Cancer 2011; 129(8): 2038-41.
III
Nieminen MT, Novak-Frazer L, Collins R, Dawsey SP, Dawsey SM, Abnet CC,
White RE, Freedman ND, Mwachiro M, Bowyer P, Salaspuro M, Rautemaa R.
Alcohol and acetaldehyde in African fermented milk mursik – A possible
etiological factor for high incidence of esophageal cancer in western Kenya.
Cancer Epidemiology, Biomarkers & Prevention 2013; 22(1): 69-75.
IV
Nieminen MT, Novak-Frazer L, Richardson V, Rajendran R, Sorsa T, Ramage G,
Bowyer P, Rautemaa R. A novel antifungal is active against Candida albicans
biofilms and inhibits mutagenic acetaldehyde production in vitro. PLoS one
2014; 9(5): e97864.
V
Nieminen MT, Hernandez M, Novak-Frazer L, Kuula H, Ramage G, Bowyer P,
Warn P, Sorsa T, Rautemaa R. D,L-2-hydroxyisocaproic acid (HICA) attenuates
inflammatory responses in a murine Candida albicans biofilm model. Clinical
and Vaccine Immunology 2014; 21(9): 1240-1245.
Publication II has been used as part of the thesis of author J.U. The original
publications have been reprinted with the permission of their copyright holders.
8
2
ACH
Acs
ADH
Adh
adh
AIDS
ALDH
Ald
ald
Als
AmB
ATP
BA
CaACS
CaADH
CaALD
CaPDC
CaPDH
cDNA
CO2
CLED
CMC
CVC
CYP
DC
Del-1
dH2O
DNA
dsDNA
ESCC
F-6-P
FAA
G-6-P
GI
HIV
IC
ICU
ABBREVIATIONS
acetaldehyde
acetyl-CoA synthetase (yeast)
alcohol dehydrogenase (human)
alcohol dehydrogenase (yeast)
alcohol dehydrogenase (bacteria)
acquired immune deficiency syndrome
aldehyde dehydrogenase (human)
aldehyde dehydrogenase (yeast)
aldehyde dehydrogenase (bacteria)
agglutinin-like sequences (yeast)
amphotericin B
adenosine triphosphate
lysed blood agar
acetyl-Coa synthetase gene (C. albicans)
alcohol dehydrogenase gene (C. albicans)
aldehyde dehydrogenase gene (C. albicans)
pyruvate decarboxylase gene (C. albicans)
pyruvate dehydrogenase enzyme complex (C. albicans)
complementary deoxyribonucleic acid
carbon dioxide
cysteine-, lactose- and electrolyte-deficient agar
chronic mucocutaneous candidosis
central venous catheter
cytochrome P450 (human)
dendritic cell
developmental endothelial locus 1
distilled water
deoxyribonucleic acid
double-stranded DNA
oesophageal squamous cell carcinoma
fructose-6-phosphate
fastidious anaerobe agar
glucose-6-phosphate
gastrointestinal
human immunodeficiency virus
invasive candidosis
intensive care unit
9
IL1
MEOS
MMP
MPO
NAC
NAD
NE
NLR
NV
OSCC
PAMP
PBS
PCA
PMN
PRR
ROS
RPMI
Xdh
XTT
SAB
Sap
ScADH
ScALD
SCC
SEM
TCA
TLR
TNFα
10
interleukin 1
microsomal ethanol oxidising system
matrix metalloproteinase
myeloperoxidase
non-albicans Candida
nicotinamide adenine dinucleotide
neutrophil elastase
nod-like receptor
neomycin-vancomycin blood agar
oral squamous cell carcinoma
pathogen associated molecular pattern
phosphate buffered saline
perchloric acid
polymorphonuclear neutrophil
pattern recognition receptor
reactive oxygen species
Roswell Park Memorial Institute -1640 media
xylitol dehydrogenase (yeast)
2,3-Bis(2-methoxy-4-nitro-5-sulfophenyl)-2H-tetrazolium-5carboxanilide inner salt
Sabouraud agar
secreted aspartic protease (yeast)
alcohol dehydrogenase gene (S. cerevisiae)
aldehyde dehydrogenase gene (S. cerevisiae)
squamous cell carcinoma
scanning electron microscopy
tricarboxylic acid cycle
Toll-like receptor
tumour necrosis factor alpha
3
ABSTRACT
Cancers of the upper digestive tract (oral, pharyngeal, laryngeal, oesophageal and
stomach) are among the ten most common cancers worldwide and are associated
with high mortality and morbidity. Despite improved treatment strategies, the
prognosis of these cancers remains poor. The main aetiological factors for the upper
digestive tract cancers are alcohol consumption and smoking. Dietary and genetic
factors can also contribute to an increased risk. Alcohol consumption, smoking and
poor oral hygiene all lead to exposure of the upper digestive tract mucosa to salivary
acetaldehyde. Ethanol is not carcinogenic, but acetaldehyde, its first metabolite, is.
The International Agency for Research on Cancer has classified acetaldehyde
associated with alcoholic consumption as a group I carcinogen. Cigarette smoke
contains acetaldehyde, as does saliva after alcohol intake. Microbially fermented food
products can also contain high levels of acetaldehyde. Therefore, salivary
acetaldehyde levels are a result of its direct intake as well as systemic, local and
microbial ethanol metabolism. The ability of the mucosa and microbes to remove
acetaldehyde has been shown to be limited.
Chronic Candida mucositis is associated with oro-oesophageal carcinogenesis.
Candida yeasts are common colonisers of the upper digestive tract, with an
extraordinary ability to form biofilms. Candida albicans is the dominant species in
health and disease. The biofilm structure protects fungal cells against hostile actions,
and infections are mainly biofilm-related. This is highlighted by high resistance to
antifungals by yeast biofilms. Fungal cells can become dispersed from a biofilm,
making it a constant source of infection. In immunocompromised patients, this
process can lead to fatal systemic infections. The similarity of fungal and human cells
makes the development of antifungals difficult, and treatment is often compromised.
The primary aim of the research presented in this thesis was to investigate the ability
of clinically important Candida yeasts to produce carcinogenic acetaldehyde in
planktonic and biofilm modes of growth in vitro. Fermented milk mursik constitutes a
major part of the diet in western Kenya, an area with a high local incidence of
oesophageal cancer and a low prevalence of common risk factors. The second aim was
to identify the microbiota of the starter cultures used for the production of mursik.
The ability of these mixed cultures of bacteria and yeast to produce acetaldehyde and
ethanol was assessed in a fermentation experiment. Finally, the potential anti-biofilm
and anti-inflammatory effect of a Lactobacillus metabolite, 2-hydroxyisocaproic acid
(HICA), was explored in vitro and in vivo. HICA has previously shown efficacy against
various planktonically grown microbes.
11
The results of the research presented in this thesis demonstrate that C. albicans
biofilms can produce mutagenic levels of acetaldehyde when incubated with ethanol
and glucose at clinically relevant concentrations. In addition, planktonic C. albicans
and non-albicans Candida (NAC) species produce significant levels of acetaldehyde in
the presence of ethanol. C. glabrata shows a unique ability among NAC species to
produce mutagenic acetaldehyde in the presence of sugars. Interestingly, the
production of acetaldehyde by Candida species can be significantly reduced by the
sugar alcohol xylitol. Highly mutagenic levels of acetaldehyde and ethanol can be
measured during the production of mursik. Candida yeasts are associated with high
acetaldehyde and ethanol production during fermentation. C. albicans biofilm
formation is significantly reduced in response to HICA treatment in vitro. Interestingly,
certain genes belonging to the fermentative pathway and acetaldehyde metabolism
are highly up-regulated under stress and during biofilm formation. An attenuated
inflammatory response was observed when a C. albicans biofilm was treated with
HICA in vivo.
Overall, these results underline the significant potential of clinically relevant Candida
yeasts to produce carcinogenic acetaldehyde. Colonisation of the upper digestive tract
with Candida spp. could expose the mucosa to mutagenic levels of acetaldehyde in
the presence of alcohol and dietary sugars. This could play a role in the carcinogenesis
associated with Candida infection. In addition, the consumption of fermented food
products such as mursik, with a high acetaldehyde and ethanol content, could elevate
the risk of cancer. HICA could provide a safe and efficient approach to treat biofilmrelated Candida infections and reduce the inflammation and mutagenicity associated
with C. albicans biofilms.
12
4
INTRODUCTION
The human microbiome, a complex community comprising all the bacterial, archaeal,
viral and fungal members, has a major impact on health and disease, and vice versa.
To fully understand the biological function of the human body, we also need to
understand the function and composition of our microbiome. Interestingly, the total
number of microbial genes is estimated to be 150 times larger than the number of
human genes, thus showing the genetic potential of the microbiome (Qin et al. 2010).
In health, the microbiome functions as a barrier against invaders and has a major role
in the development of vital innate and adaptive immune responses in early life. In
disease, the microbiome acts as a platform for the pathogenic growth of microorganisms and pathological immune responses that can lead to chronic inflammatory
disease in the long term and eventually to malignancy. In the past ten years, the
enormous diversity of the human microbiome has started to be revealed (Human
Microbiome Project 2012). These members interact with each other, and this
interplay has a specific role in maintaining the balance between the microbiome and
the host. Major steps have been taken towards understanding of the microbiome due
to major advances in detection techniques, particularly culture-independent
technologies and bioinformatics. Large projects, including the Human Microbiome
Project and Metagenomics of the Human Intestinal Tract (MetaHIT), have greatly
enhanced progress in the field (Human Microbiome Project 2012, Le Chatelier et al.
2013). Although there has been major progress, work has mainly focused on the
bacterial part of the microbiome, and little is still known about the fungal part of the
human microbiome, recently defined as the mycobiome (Cui et al. 2013, Ghannoum
et al. 2010).
The oral cavity is a major gateway for micro-organisms to enter the human body. As
an entry point to the gastrointestinal tract, the oral cavities of healthy individuals
harbour a diverse and fluctuating collection of micro-organisms (Aas et al. 2005, Bik et
al. 2010, Nasidze et al. 2009). Altogether, 1,179 taxa of bacteria have been identified
in the oral microbiome, also including uncultivable and unnamed taxa (Dewhirst et al.
2010). The most prevalent bacterial genera found are Streptococcus, Haemophilus,
Prevotella, Veillonella and Corynebacterium (Human Microbiome Project 2012). The
relative abundance of genera differs between the diverse environmental niches of the
oral cavity (Human Microbiome Project 2012). Anatomically, the oral cavity is
conjoined with the oesophagus, which further leads to the stomach. These three sites
comprise the upper digestive tract. However, the main focus of this thesis is on the
former two sites, as they share similar characteristics regarding mucosal fungal
infections. Although both the oral cavity and oesophagus have unique microbiomes,
13
they are connected by saliva and similar dominant genera are present (Bik et al. 2010,
Pei et al. 2004, Segata et al. 2012). Co-existing alongside the oral bacterial microbiome
is a fungal community that is dominated by the genus Candida, particularly Candida
albicans. Candida yeasts are opportunistic pathogens. Compromised immunity and
the loss of homeostasis within the microbiome predispose the human host to
increased colonisation by Candida spp. and induce the transition of Candida yeast
from commensal to pathogenic traits.
A biofilm lifestyle is preferred in the microbial communities along the GI tract
(Jenkinson and Lamont 2005, Macfarlane et al. 2011). This is particularly evident in
the oral cavity, where polymicrobial communities exist in multiple forms, such as
dental plaque (Marsh 2004). The majority of human microbial infections, including
Candida spp., are biofilm-related (Costerton et al. 1999, Donlan and Costerton 2002).
Good oral hygiene is required for the maintenance of a healthy balance at the orooesophageal host–microbiome interface. Long-term negligence of hygiene leads to
the aberrant growth of pathogenic bacteria and fungi (Marsh 2006, Sbordone and
Bortolaia 2004). The biofilm structure provides a well-protected environment for
microbes to grow, which is highlighted in Candida yeasts by an increase in resistance
against commonly used antifungals such as azoles. The frequent use of antifungals
and antibacterial compounds can lead to the selective growth of more resistant
strains and the loss of vital homeostasis as the spectrum of species narrows down
within the local microbiome (Jakobsson et al. 2010, Rautemaa et al. 2007b, Robinson
and Young 2010). This can aggravate the initial infection and related disease. As the
number of immune-compromised patients is increasing due to advances in medical
treatment, such as cancer therapies and organ transplantation surgery, the incidence
of opportunistic fungal infections is rising (Brown et al. 2012). Yeasts are eukaryotic
organisms and their similarity to human cells makes the development of antifungals
difficult. Current antifungals have poor efficacy against biofilms. Overall, more
advanced and effective treatment strategies are required. Prior the use of antifungal
medication, the traditional anti-infective actions to maintain oral hygiene, such as the
daily disruption of biofilms using a toothbrush and regular check-ups by dental
professionals should be highlighted in the prevention of oro-oesophageal biofilm
infections.
Members of the oral microbiome can be the primary cause of oral mucosal lesions or
a source for secondary intruders in non-infectious inflammatory mucosal conditions.
The most common mucosal infection caused by fungi in humans is oral candidosis. In
the long term, local inflammation induced by candidosis may lead to mucosal
dysplasia and other pre-malignant changes. Studies have shown that chronic candida
mucositis (CMC) can be carcinogenic in the upper digestive tract in vivo (Delsing et al.
14
2012, Rautemaa et al. 2007a). In addition, oral and oesophageal cancer lesions have
been associated with Candida, thus further raising a question about the link between
Candida yeasts and upper digestive tract carcinogenesis (Nagy et al. 1998, Scott and
Jenkins 1982).
Alcohol abuse and smoking are major and synergistic risk factors for upper digestive
tract cancers. In the human body, ethanol is mainly metabolised by the liver, but
ethanol metabolism at the mucosa–biofilm interface of the GI tract has been shown to
play an important role regarding its carcinogenicity. Acetaldehyde (ACH) is the first
metabolite of ethanol oxidation, and ACH exposure linked to chronic alcohol
consumption has been ranked as a group I carcinogen by the International Agency for
Research on Cancer (Secretan et al. 2009). ACH is one of the major carcinogens of
tobacco smoke. Both of these risk factors result in an increase in salivary ACH.
Previous studies have shown that the microbiome is the main determinant of local
ACH production in the upper digestive tract. The primary aim of the research
presented in this thesis was to investigate the capacity of Candida spp. or a mixed
bacteria/yeast cultures to produce mutagenic ACH and to provide new insights into
the carcinogenicity of Candida infections. Secondly, the goal was to find more
effective ways to inhibit the mutagenicity and pathogenicity of C. albicans biofilms
than currently available.
15
5
REVIEW OF THE LITERATURE
5.1
CANDIDA YEASTS AND THE HUMAN MYCOBIOME
In recent years, the extent and characteristics of fungal ecosystems in the human
body have begun to become apparent in conjunction with the progression of studies
on the human microbiome (Ghannoum et al. 2010, Huffnagle and Noverr 2013). The
spectrum and frequency of fungal species in the human body is strongly affected by
health and disease. The interplay between fungal and bacterial microbiomes affects
the balance, and both synergistic and competitive behaviour between the two have
been noted (Cui et al. 2013). The distribution and diversity of fungal species can vary
considerably between distant body sites, but similarities have been found between
neighbouring sites, such as the oral and nasal cavities (Ghannoum et al. 2010, SellartAltisent et al. 2007).
The most common commensal fungi in humans in both health and disease are
Candida yeasts, which are mostly isolated from the mucosal surfaces of the GI tract
and vagina (Calderone and Clancy 2011, Drell et al. 2013, Ghannoum et al. 2010).
Candida can also be found on the skin, but colonization is site-specific and not as
frequent as in the aforementioned niches (Findley et al. 2013). The genus Candida
falls into the ascomycetes in the fungal phylum, and more specifically into a
subphylum called Saccharomycotina. Approximately 160 species of Candida exist
(Calderone and Clancy 2011). However, only a small proportion of these species are
pathogenic and important in terms of human infections. Candida albicans is the most
common Candida species found in humans in both health and disease. Other
medically important Candida species include C. dubliniensis, C. famata, C. glabrata,
C. guilliermondii, C. inconspicua, C. kefyr, C. krusei, C. lusitaniae, C. norvegensis, C.
parapsilosis, C. rugosa and C. tropicalis, which are often referred to as non-albicans
Candida species (NAC; Table 1, Calderone and Clancy 2011, Pfaller et al. 2010). NAC
species are predominantly isolated from immunocompromised patients with
advanced medical conditions such as haematological malignancy or bone marrow
transplant (Krcmery and Barnes 2002).
16
Table 1. Major characteristics of C. albicans and non-albicans Candida yeasts, including their
frequency in health and disease (infection). The infection frequencies in patients with infections
are derived from a study that compiled statistics from 100 medical centres internationally.
Frequencies in health are based on data from a medical centre in the USA (aPfaller et al. 2011,
b
Ghannoum et al. 2010, cThompson et al. 2011, dButler et al. 2009, eSilva et al. 2009; nd. = not
described).
There are marked differences in the characteristics and behaviour of Candida species
due to their independent evolution and survival in unique habitats (Table 1). The
ability to adapt and endure stress considerably differs between species (Papon et al.
2013). Some of these differences can be explained by the large genetic distances
between species in the phylogenetic tree, some of which are larger than the distance
between humans and certain genera of fish (Figure 1; Dujon 2010). The subphylum
Saccharomycotina has two main branches or clades. Clinically important Candida
yeasts are primarily located in the Candida CTG clade, which was named due to the
alternative translation of the CTG codon as serine instead of leucine. C. glabrata
alongside S. cerevisiae belongs to a whole genome duplication (WGD) clade under the
Saccharomycetaceae, which also includes species such as C. krusei.
The mycobiome of the upper digestive tract is dominated by Candida spp. The mean
carriage rate of Candida spp. in the oral cavity of healthy humans varies (18% to 75%;
Cannon and Chaffin 1999, Ghannoum et al. 2010). C. albicans is the dominant species
in healthy subjects, but C. parapsilosis and C. tropicalis have also been found (Drell et
al. 2013, Ghannoum et al. 2010, Scanlan and Marchesi 2008). Other frequently
present fungi in the oral cavity include Malassezia, Cladosporium, Aureobasidium,
Saccharomyces, Aspergillus, Fusarium and Cryptococcus species (Dupuy et al. 2014,
Ghannoum et al. 2010). Previous studies have pointed out the similarities between
oesophageal and oral microbiomes (Gagliardi et al. 1998, Norder Grusell et al. 2013).
C. albicans has been isolated from samples taken along the upper and lower parts of
17
the oesophagus and is the dominant fungus and member of Candida spp. in the
oesophageal microbiome (Norder Grusell et al. 2013).
Figure 1. Phylogenetic tree illustrating genetic distances and clade distribution between the
clinically important Candida species in the subphylum Saccharomycotina. The distances shown
are arbitrary (reproduced from Papon et al. 2013).
5.2
PATHOGENESIS OF CANDIDA YEASTS
5.2.1
Virulence factors
The pathogenic properties of Candida yeasts are dominated by six major virulence
factors: polymorphism, phenotypic switching, biofilm formation, adhesion, invasion
and hydrolysis. These properties greatly differ between and within species, and the
pathogenicity of an isolate can consequently be drastically different from one to
another.
5.2.1.1
Polymorphism and phenotypic switching
Candida yeasts are polymorphic organisms. They can exist in a unicellular budding
form (yeast) or in two filamentous forms: as elongated ellipsoidal structures with
constrictions in septae (pseudohyphae) or as parallel lined filaments without
constrictions (hyphae). However, only C. albicans, C. dubliniensis and C. tropicalis can
form true hyphae (Table 1; Thompson et al. 2011). Hyphal formation has a major role
18
in tissue invasion, while the yeast form is more important for dissemination (Saville et
al. 2003). Candida can also transit between two phenotypes: i) white cells are round
and dome-shaped with a shiny cover, and ii) opaque cells form flatter and darker
colonies that can be elongated. The major transcriptional regulator for phenotypic
switching is the gene WOR1 (Huang et al. 2006). White-to-opaque switching has been
shown to occur during mating and biofilm formation. It has been suggested to play a
role in antifungal resistance and colonisation (Alby and Bennett 2009, Soll 2014).
Most Candida yeasts are diploid organisms, but some exceptions exist among NAC
species (Table 1; Butler et al. 2009). C. albicans was long considered as a strict diploid,
but a very recent study reported that C. albicans is able to transform its ploidy in
response to environmental challenges, and haploid or even polyploid forms also exist
(Hickman et al. 2013). This can be advantageous when the yeast is confronted by
various stress factors, such as antifungals (Harrison et al. 2014). The transition in
phenotype and ploidy highlights the genomic plasticity of C. albicans and its advanced
ability to adapt to different environmental niches.
5.2.1.2
Biofilm lifestyle
The main virulence factor underlined in the context of mucosal colonisation and
pathogenesis is the ability of Candida to grow as a biofilm. The earliest descriptions of
th
the biofilm lifestyle date back to the late 17 century, when Antonie van
Leeuwenhoek described dental plaque from his own teeth (van Leeuwenhoek 1684).
However, the general concept of a biofilm was first described by Bill Costerton in 1978
(Costerton et al. 1978). Approximately 80% of micro-organisms live in biofilm
communities and 65–80% of human infections are biofilm-related (Costerton et al.
1999, Donlan and Costerton 2002). Candida biofilms are structurally advanced
populations of fungal cells in yeast and/or hyphal forms adhered to either an abiotic
or biotic surface. These populations are embedded in a self-produced extracellular
carbohydrate polymeric matrix. Common abiotic surfaces in the human body
colonised with Candida include central venous catheters (CVC) and various prosthetic
structures such as dentures or tooth implants (Andes et al. 2004, Nett et al. 2010).
Candida frequently forms biofilms on biotic surfaces in the human body, particularly
on the mucosa of the GI tract and the vaginal epithelium (Dongari-Bagtzoglou et al.
2009, Harriott et al. 2010). Cell populations within biofilms phenotypically differ from
their planktonic counterparts due to transcriptional reprogramming, and this
highlights the unique nature of biofilm communities (Nett et al. 2009, Yeater et al.
2007).
19
The development of a biofilm is a multistep process (Finkel and Mitchell 2011). In the
early phase, Candida cells adhere to the surface and form micro-colonies. These cells
go through morphological switching and form hyphae, if able to do so. These hyphal
structures act as a rigid and solid mesh that supports the growth of a cell population
and resists mechanical stress (Figure 2). In the phase of maturation, fungal cells
multiply and the matrix structure starts to form to support and protect the cells
against environmental stress. In older biofilms, a dense and thick extracellular matrix
covers the cells and hyphae. β-Glucans, particularly β-1,3-glucan, and mannans are
the principle carbohydrates in the matrix structure (Nett et al. 2007). Fungal cells
begin to disperse in the final stages of biofilm development as the population starts to
regulate its growth by secreting quorum-sensing molecules, such as farnesol (Ramage
et al. 2002b). These dispersed cells have been shown to be more virulent in a mouse
model for disseminated candidosis (Uppuluri et al. 2010b). Therefore, biofilms can be
considered as a constant source of infection. Hyphal and biofilm formation are both
highly governed by various environmental cues, including pH, temperature, osmolarity
and the presence of endogenous or exogenous antimicrobial agents (Chandra et al.
2001, Ene et al. 2012, Fiori et al. 2012, Kucharikova et al. 2011).
Figure 2. Significant steps in biofilm formation: i) cellular adherence, ii) initial micro-colony
formation and hyphal growth, iii) cellular multiplication and matrix formation and iv) dispersal of
the cells from the biofilm due to quorum sensing. The SEM image above shows the ultrastructure
of a mature (24h) C. albicans biofilm characterised by budding ovoid yeast cells and a hyphal
network.
Due to the advances in genomics, major progress has been made in understanding the
plethora of genes that regulate C. albicans biofilm formation. A recent study revealed
20
six major transcription factors (TFs) that govern C. albicans biofilm formation: Bcr1p,
Brg1p, Efg1p, Ndt80p, Rob1p and Tec1p (Nobile et al. 2012). This transcriptional
circuitry includes over 1,000 genes, which represents 15% of all Candida genes.
Candida species have a varied ability to form biofilms (Table 1; Kuhn et al. 2002, Silva
et al. 2009). C. albicans, C. tropicalis, C. parapsilosis and C. glabrata can all form fully
mature biofilms. However, C. albicans and C. tropicalis biofilms show a more complex
and robust biofilm ultrastructure compared to the more uniform and thinner biofilms
produced by C. glabrata and C. parapsilosis. Biofilm formation and the different
properties are not only dependent on the species but also on the strain and
environmental conditions (Silva et al. 2009).
Mature biofilms are highly resistant against antimicrobial compounds, host immune
factors and antimicrobial peptides (Ramage et al. 2014). Biofilms are also highly
tolerant and adaptive to environmental cues, such as hypoxia, and oxidative and
mechanical stress (Bonhomme et al. 2011, Paramonova et al. 2009, Seneviratne et al.
2008, Stichternoth and Ernst 2009). The factors behind increased resistance include
the advanced matrix structure, the hyphal-cell ratio, cell density, transcriptional
rewiring, extracellular DNA, the occurrence of persister cells and increased expression
of efflux pumps (Martins et al. 2010, Nett et al. 2007, Perumal et al. 2007, Ramage et
al. 2002a, Sellam et al. 2009, Vediyappan et al. 2010).
5.2.1.3
Adhesion, invasion and hydrolysis
Overall, effective adherence to a biomaterial or the host surface is of major
importance to biofilm formation and survival. The agglutinin-like sequence (ALS) is a
major family of adhesion proteins or adhesins, and a total of eight ALS adhesins have
been identified in C. albicans (Hoyer et al. 2008). Als3p has been shown to be the
most vital adhesin, which mediates attachment to host cells and artificial surfaces
(Almeida et al. 2008). In addition to Als3p, Hwp1p has been recognised as an
important adhesin and virulence factor (Staab et al. 1999). Both Als3p and Hwp1p
contribute strongly to biofilm formation (Nobile et al. 2008).
Invasion of the host tissue is important for the establishment of infection and nutrient
acquisition. Candida expresses specific proteins called invasins that can link the cell to
host ligands and induce endocytosis (Phan et al. 2005, Phan et al. 2007). Hyphaassociated Als3p and Ssa1p have been recognized as the major invasins of C. albicans
(Phan et al. 2007, Sun et al. 2010). Hyphal formation is essential for both active
penetration of tissue and induced endocytosis (Wachtler et al. 2012, Gow and Hube
2012). Mechanical stress, nutrient sources, contact sensing and surface topology have
21
been shown to activate and regulate both hyphal and biofilm formation (Brand and
Gow 2009, Brand et al. 2008, Watts et al. 1998). These factors are important for
virulence in vivo (Brand et al. 2008). Hydrolytic enzymes secreted by Candida spp.,
including proteases, lipases and phospholipases, are major virulence factors and
important for effective invasion. Secreted aspartic proteases (SAPs) have been shown
to be the main family of proteases and have a dominant role in invasion and virulence.
Their expression differs depending on environmental factors such as the pH and
nutrient concentrations. The number of SAPs varies between Candida species. C.
albicans expresses a total of ten SAPs (Naglik et al. 2003).
5.2.2
Infection and disease
The incidence of fungal infections has been rising due to an increase in the number of
immunocompromised patients (Garcia-Vidal et al. 2013, Pfaller and Diekema 2007).
Candida is the most common cause of human fungal infection (Pfaller and Diekema
2007). In general, the infections caused by Candida yeasts vary from superficial
mucosal infections, such as vaginal or oral candidosis, to life-threatening invasive
infections. In healthy subjects, the colonisation and growth of Candida yeasts is
controlled by specific and non-specific defence mechanisms, such as salivary and
mucosal antimicrobial compounds, as well as by competitive inhibition by the
surrounding microbiome. There are multiple pre-disposing factors that solely or in
combination allow Candida to transform from a commensal organism into a pathogen
in the upper digestive tract (Table 2).
Table 2. Local, systemic and treatment-related predisposing factors for candidosis of the oral
cavity and oesophagus.
In general, an opportunity for infection arises when local or systemic defence
mechanisms are weakened, the homeostasis within the microbiome is altered, or oral
hygiene is poorly managed (Table 2; Laudenbach and Epstein 2009, Rautemaa and
Ramage 2011, Samaranayake et al. 2009). In addition, uncleaned and cracked
dentures, salivary hypofunction and smoking have adverse effects on mucosal health
22
and increase the risk of Candida overgrowth (Baboni et al. 2009, Fanello et al. 2006,
Grimoud et al. 2005, Närhi et al. 1993). Systemic factors should be well assessed, as
immunosuppressed patients are at major risk of developing candidosis. The use of
corticosteroids, high glucose levels due to imbalanced diabetes, long-term treatment
with broad-spectrum antibiotics, a low pH due to reflux disease and various immune
deficiencies all increase the risk of development of candidosis (Arendorf and Walker
1980, Buhl 2006, Darwazeh et al. 1990, Kiehne et al. 2005). Oral candidosis is the most
common oral complication in patients with human immunodeficiency virus (HIV)
(Patton et al. 2002, Thanyasrisung et al. 2014). Up to 90% of HIV and acquired
immune deficiency syndrome (AIDS) patients are colonised by Candida, and 27% of
HIV patients and one-third of patients with AIDS have been reported with
symptomatic oral candidosis (Calderone and Clancy 2011, Patel et al. 2012, Thompson
et al. 2010).
5.2.2.1
Oral candidosis
Oral candidosis is a superficial infection with either acute (thrush) or chronic
characteristics. The clinical manifestations include pseudomembranous and
erythematous forms, which are both typically described as acute infections and often
a result from a general cause such as antibiotic or corticosteroid treatment
(Samaranayake et al. 2009). The clinical picture of the pseudomembranous form often
includes inflamed mucosa with white patches on top. In erythematous candidosis, the
mucosa appears as red, dry and glazed. Erythematous candidosis is the most common
form diagnosed in HIV patients (Bodhade et al. 2011). Typical symptoms for both of
these forms include pain and/or a burning sensation. Chronic infections include
atrophic erythematous and hyperplastic variants. Hyperplastic candidosis can be
further divided into subtypes, including nodular, plaque-like or candida leukoplakia,
depending on the clinical picture (Samaranayake et al. 2009). Chronic infections are
characterised by both persistence and recurrence. In addition to these major variants
of candidosis, lesions with a multifactorial background are often associated with
Candida growth. These include median rhomboid glossitis, denture stomatitis, angular
cheilitis and linear gingival erythema (Samaranayake et al. 2009).
In comparison to various primary infections listed above, secondary candidosis is a
result of underlying immune deficiency or autoimmune disease. Diseases most
commonly found in the aetiology of the infections are HIV and haematological
malignancies combined with the use of cytotoxic agents (Pagano et al. 2006,
Samaranayake 1992). Chronic mucocutaneous candidosis (CMC) is a rare and
23
heterogeneous group of clinical syndromes, in which chronic Candida infections occur
in the skin and mucosal surfaces throughout the body (Liu and Hua 2007).
5.2.2.2
Oesophageal candidosis
Oesophageal candidosis, or Candida oesophagitis, is the most common infectious
disease of the oesophagus and highly prevalent in patients with HIV (8–53%; Bonacini
et al. 1991, Nishimura et al. 2013, Nkuize et al. 2010, Rolston and Rodriguez 1992).
The high variance in the reported prevalence is mainly due to the advances in antiviral
treatment in the past two decades. Candida oesophagitis is often classified into mild
and severe forms according to endoscopial findings. Severe forms are associated with
haemorrhage, oesophagotracheal fistulas and stenosis (Gaissert et al. 1999, Kanzaki et
al. 2009). The main symptom is pain during swallowing. Oral candidosis is often used
as a marker for the diagnosis of oesophageal candidosis (Samonis et al. 1998, Wilcox
et al. 1995). The majority of the predisposing factors for oral candidosis therefore
apply for oesophageal candidosis (Table 2; Baehr and McDonald 1994).
5.2.2.3
Invasive candidosis
If superficial candidosis does not respond to treatment, is left untreated or has
refractory features in immunocompromised patients, the infection can transform into
invasive candidosis (IC; Laudenbach and Epstein 2009). The term IC encompasses
various candida-related disorders with an invasive and more severe clinical picture:
candidaemia, disseminated candidosis, endocarditis, meningitis, endophthalmitis, or
deep organ involvement. In immunocompromised patients, candidal leakage into the
bloodstream is a result of high local Candida colonisation and mucosal damage in the
GI tract (Nucci and Anaissie 2001, Rosen et al. 2005, Pasqualotto et al. 2006). Patients
are commonly treated for bacteraemia prior to the onset of candidaemia. The use of
broad-spectrum antibiotics is a well-known predisposing factor for candidosis and
involves increased colonisation (Pasqualotto et al. 2006). Biofilm infections related to
medical devices, such as CVC, predispose patients to candidaemia (Kojic and
Darouiche 2004). Dispersing cells from the gut or CVC-related biofilm travel in the
bloodstream, and the highly adherent cells can disseminate into extra-intestinal
organs such as the liver or spleen. IC has been reported as the fourth leading cause of
nosocomial infections, with a yearly incidence globally varying from 2.4 to 29 per
100,000 (Pfaller and Diekema 2007). The crude mortality rates of these infections
have remained high globally (26–46%; Morgan 2005, Tortorano et al. 2006). C.
albicans is the dominant cause of infection, but the incidence of infections caused by
24
NAC species, particularly C. glabrata and C. parapsilosis, has been rising in the past
two decades (Krcmery and Barnes 2002).
5.2.3
Interaction with bacteria
As commensal organisms, Candida yeasts are in constant interaction with bacterial
counterparts in the upper digestive tract, and they can together form complex
polymicrobial biofilms on soft and hard tissues (Dongari-Bagtzoglou et al. 2009, Zijnge
et al. 2010). The outcome of this interaction may be positive or negative in terms of
human health. Denture stomatitis is an oral inflammatory disease highly associated
with Candida biofilms (Ramage et al. 2004). Mixed species biofilms of bacteria and
fungi have been shown to affect the inflammatory process in denture stomatitis and
periodontitis (Avon et al. 2007, Järvensivu et al. 2004). Streptococcus gordonii, a
common coloniser of the upper digestive tract mucosa, can induce C. albicans
filamentation and biofilm formation in a synergistic fashion (Bamford et al. 2009,
Silverman et al. 2010). Inter-species signalling affects the behaviour of bacteria and
fungi. The quorum-sensing molecule farnesol secreted by C. albicans can alter the
behaviour and virulence of Pseudomonas aeruginosa by reducing the production of
phenazine, a potent toxin (Cugini et al. 2007). In contrast, P. aeruginosa -secreted
phenazines inhibit yeast–hyphal transition and reduce biofilm formation (Morales et
al. 2013). The formation of polymicrobial biofilms can result in increased virulence of
both bacteria and yeast, and reduced susceptibility to antimicrobials as has been
shown in co-infection studies with Staphylococcus aureus or Streptococcus spp. and C.
albicans (Carlson 1983, Harriott and Noverr 2009, Shirtliff et al. 2009). In addition, it
has been recently suggested that Candida yeasts can play a role in the onset of dental
caries (Falsetta et al. 2014). Bacterial/fungal co-communities can create conditions
that can either promote or control the growth of one or the other. Lactobacillus spp.
can reduce the fungal load, and multiple studies have identified Lactobacillus
metabolites that inhibit the metabolism and growth of Candida (Atanassova et al.
2003, Kohler et al. 2012, Okkers et al. 1999, Villena et al. 2011, Wagner et al. 2000).
High vaginal colonisation by Lactobacilli inversely correlates with the incidence of
vaginitis in women (Boris and Barbes 2000).
5.2.4
Mucosal immunity
Local defence mechanisms are the primary line of immunity against mucosal
candidosis and include the epithelial barrier function, salivary and mucosal soluble
antimicrobial agents and the innate immune system. Mucosal trauma impairs the
barrier function and increases the risk of candidosis, as highlighted in denture
stomatitis (Samaranayake et al. 2009). Saliva plays an important role in antimicrobial
25
defence, and low salivary flow rates correlate with Candida colonisation (KamagataKiyoura et al. 2004, Radfar et al. 2003, Sreebny 2000). Various compounds in saliva
and mucus inhibit microbial growth, including histatins, secretory IgA, lysozyme,
mucins, defensins and lactoferrin (Li et al. 2006).
5.2.4.1
Innate immunity
The major players of the innate immune system are the epithelial cells and
phagocytotic cells, including polymorphonuclear neutrophils (PMNs), monocytes,
macrophages and dendritic cells. PMNs play a dominant role in anti-Candida
immunity. Their action in the targeted area is primed by the chemotaxis induced by
epithelial cell recognition, which leads to extravasation through the endothelial cell
layer in the capillaries. Neutropenia is a major risk factor for Candida infections, thus
highlighting the role of PMNs (Warnock 1998). Phagocytosis and the intercellular
signalling circuitry determine the outcome of the response and also initiate the
adaptive immune response to provide “memory”. Interestingly, recent studies have
shown that innate effector cells can also have “memory” when re-infection occurs
(Ifrim et al. 2014, Quintin et al. 2012).
The activation of innate immunity is primed by the recognition of Candida with
specific pattern recognition receptors (PRRs) located in epithelial or immune cells
(Cheng et al. 2012, Lionakis and Netea 2013, Naglik et al. 2011). This recognition can
lead to either suppression of the protective response (tolerance) or induction of an
immune response. An effective immune response includes complement, cytokine and
chemokine signalling, phagocytosis and antimicrobial peptide production. Human βdefensins and LL-37 have been recognised as important antimicrobial peptides in
mucosal immunity against candidosis (Li et al. 2011, Vylkova et al. 2007). In addition,
extracellular proteases, including matrix metalloproteinases (MMPs), are produced by
epithelial and immune effector cells to regulate the mucosal barrier function, and
have been shown to modulate the epithelium during infection (Claveau et al. 2004).
These factors are involved in the inflammatory response, which can eventually lead to
host damage. PRRs bind to specific pathogen-associated molecular patterns (PAMPs)
on the surface of yeast cells or hyphae. The hyphal cell surface is structurally different
from yeast cells and involves different downstream signalling and pathway activation
(Moyes et al. 2010). PRRs can be divided into three major classes: Toll-like receptors
(TLRs), lectins and intracellular Nod-like receptors (NLRs). The main targets for these
receptors are various cell wall components (differently linked mannans, glucans,
lipoproteins, lipopolysaccharides) and fungal RNA or DNA. Synergistic interactions
have been demonstrated between PRRs, such as Dectin-1, TLR-2 and TLR-4 (Ferwerda
26
et al. 2008, Gantner et al. 2003). PRR-PAMP binding leads to the stimulation of
signalling cascades and eventually to the activation of more or less two major
pathways: NF-κβ and Type I IFN (Lionakis and Netea 2013). In addition, candidal
immunity involves the inflammasome pathways: NLRP3-Caspase-1 and CARD9Caspase-8, which have been implicated to play a role in IL1β production (Gringhuis et
al. 2012, Kumar et al. 2009).
The cytokine- or chemokine profile can either be pro- or anti-inflammatory and is
determined by the sum of the pathways above. Pro-inflammatory cytokines IL-6, IL-8
and TNFα have a major role in PMN recruitment and activation (Balish et al. 1999,
Netea et al. 1999, Romani et al. 1996, van Enckevort et al. 1999). In addition to the
pro-inflammatory cytokines, hematopoietic growth factors granulocyte colonystimulation factor (G-CSF) and granulocyte-macrophage colony stimulation factor
(GM-CSF) are important for PMN recruitment and activation (Kullberg et al. 1999,
Lieschke and Burgess 1992). IL1α and ILβ are crucial cytokines in candidal immunity
(Vonk et al. 2006). PMNs and mononuclear cells, monocytes and macrophages, can kill
both intra- and extracellularly. Hydrolytic enzymes, antimicrobial peptides and
reactive oxygen radical are produced by PMNs. Myeloperoxidase (MPO) and
neutrophil elastase (NE) are the major hydrolytic enzymes produced by PMNs and are
important in the activation of monocytes/macrophages and extracellular proteases,
such as matrix metalloproteinase. In addition, neutrophils also produce extracellular
fibre structures named neutrophil extracellular traps (NETs) that can capture and kill
Candida (Urban et al. 2006). Considering the importance of the biofilm lifestyle in
pathogenesis, few reports have shown that a biofilm induces an attenuated immune
response with a very distinct cytokine profile in contrast to planktonic cells in vitro
(Chandra et al. 2007, Xie et al. 2012). Biofilms have been proposed as a possible
immune evasion mechanism (Dwivedi et al. 2011). A similar response has been noted
in a study with S. aureus biofilms in vivo (Thurlow et al. 2011).
5.2.4.2
Adaptive immunity
Dendritic cells (DCs) function as a major link between innate and adaptive immunity.
DCs phagocytose and process Candida for antigen presentation and activate T-cells in
lymph nodes (Newman and Holly 2001). Adaptive immunity against mucosal
candidosis is coordinated by Th1, Th2 and Th17 T-cell responses, which are mainly
regulated by CD4+ T-cells. However, a compensatory role has been described for CD8+
T-cells in HIV patients (Fidel 2011). Th1 T-cells activate phagocytotic cells by specific
signals against Candida, while in contrast, Th2 T-cells attenuate the candidicidal
measures by inhibiting effector cell actions. Th17 T-cells act in synergy with Th1 T-cells
27
and are the main source of IL-17. The Th17 response contributes to multiple facets of
both innate and adaptive immunity: PMN chemotaxis and the regulation of
antimicrobial peptide, cytokine and MMP production (Huppler et al. 2014). Deficiency
of IL-17 results in increased susceptibility to mucosal and systemic candidosis (Conti et
al. 2009, Ho et al. 2010, Huppler et al. 2014, Puel et al. 2011). Generally, no strict line
can be drawn between innate and adaptive responses. Effective candidal immunity is
underlined by the advanced integration of factors from both innate and adaptive
sides.
5.3
TREATMENT STRATEGIES
Considering the nature of oro-oesophageal candidosis as a biofilm infection, the
outcome of treatment depends on the ability to disrupt and remove the biofilm.
Antifungal resistance is an increasing problem and highlighted in biofilm infections.
The maintenance of good oral hygiene is the backbone for the successful treatment of
oral candidosis, particularly in patients with impaired immunity. The successful
treatment of candidosis involves the identification of impaired local and systemic
defence mechanisms. The treatment strategy should be chosen based on the extent
and nature of Candida infection, for example whether the infection is local/superficial
or systemic, acute or chronic, or whether it is a recurrent infection.
Currently, three major classes of antifungals are used in the treatment of candidosis:
polyenes, azoles (imidazoles and triazoles) and echinocandins. They all target the
function or structure of the cell membrane or wall (Figure 3). All three classes can be
administered intravenously, and topical formulations are available for certain
polyenes and azoles. In general, immunocompetent patients with oral candidosis
should be treated with topical antifungals, while a combination of topical and
systemic antifungals should be targeted at immunocompromised patients.
Oesophageal candidosis requires systemic antifungal therapy (Denning and Hope
2010). If candidosis is refractory, the risk for the development and selection of
resistant strains should be considered and minimised when choosing the appropriate
antifungal. The potential presence of NAC species should also be kept in mind in
recurrent and refractory infections. NAC species often require different medications
for successful treatment, as they exhibit decreased susceptibility, particularly towards
azoles. This underlines the importance of accurate microbial diagnostics in the
treatment of candidosis, especially in immunocompromised patients with recurrent
infections. Increasing attention has recently been drawn to the mechanisms by which
antifungals modulate host immune responses, which might be of importance when
28
constructing a rational treatment strategy against deep-seated infections in patients
with impaired immunity (Ben-Ami et al. 2008).
Figure 3. Cellular targets of major antifungals used for the topical and systemic treatment of
candidosis.
5.3.1
5.3.1.1
Treatment of infection
Topical treatment
Topical antifungals are used for the local treatment of oral candidosis, and high drug
levels can be obtained in the targeted area. Their absorption from the GI tract is
mainly poor, and the systemic levels will therefore remain low. Thus, side effects and
drug–drug interactions will not be as strong a compromising factor as with
intravenously administered antifungals, although local irritation, GI symptoms, poor
taste and inconvenience have been noted (Epstein et al. 2004).
5.3.1.1.1 Polyenes
Nystatin and amphotericin B (AmB) are polyene antifungals that bind to ergosterol in
the fungal cell membrane and increase membrane permeability (Figure 3). Both
nystatin and amphotericin B have been widely used topically, as they are poorly
absorbed from the GI tract and have various adverse effects in systemic therapy. They
have been considered as the primary choice for the local treatment of oral candidosis
and have broad-spectrum antimicrobial activity (Rautemaa and Ramage 2011).
29
5.3.1.1.2 Azoles
Clotrimazole and miconazole inhibit fungal ergosterol synthesis, which is vital for the
fungal cell plasma membrane. This results in an increase in cell membrane
permeability (Figure 3). Both compounds are readily absorbed from the GI tract, and
drug–drug interactions will thus occur. Clotrimazole is a moderate inhibitor of CYP3A4
on the GI cell wall and can increase the bioavailability of certain drugs, such as
tacrolimus, cyclosporine, bentzodiazepines and HMG-CoA reductase inhibitors
(Laudenbach and Epstein 2009, Vasquez et al. 2005). Miconazole treatment leads to
fungal cell death due to the accumulation of reactive oxygen species (ROS) and
oxidative stress (Barasch and Griffin 2008). Both drugs have long been used for the
treatment of superficial candidosis, as they have both antifungal, mainly fungistatic,
and antibacterial activity. Miconazole is active against various Candida spp., including
fluconazole resistance isolates (Isham and Ghannoum 2010).
5.3.1.1.3 Chlorhexidine
Chlorhexidine is a cationic polybisguanide compound that has broad-spectrum
antimicrobial activity. It impairs the function of the cell membrane by increasing its
fluidity (Figure 3). Chlorhexidine gluconate solution (0.2%) as an oral rinse has shown
efficacy in the treatment of oral candidosis when used in combination with tooth
brushing (Barasch and Griffin 2008, Grimoud et al. 2005). It has shown broadspectrum activity against various Candida spp., including fluconazole-resistant strains
(Salim et al. 2013a). Chlorhexidine and nystatin should not be applied simultaneously,
because interaction and loss of the anti-Candida effect will occur (Barkvoll and
Attramadal 1989). A 4% suspension of chlorhexidine has been successfully used as a
denture disinfectant (Mantri et al. 2013).
5.3.1.2
Systemic treatment
Systemic antifungal therapy should be considered in complicated refractory superficial
infections or in patients who are at risk of systemic Candida infection. First-generation
triazoles (fluconazole and itraconazole) are considered as a first line of antifungal
therapy if oral systemic treatment is required (Laudenbach and Epstein 2009,
Rautemaa and Ramage 2011). The use of second-generation triazoles (voriconazole
and posaconazole) should be restricted to immunocompromised patients with a
severe infection and resistance profile. In cases where triazole cross-resistance has
been diagnosed or infection is device- and biofilm-related, echinocandins are the best
option, followed by amB (Cornely et al. 2012).
30
5.3.1.2.1 Triazoles
Triazoles interrupt ergosterol biosynthesis by inhibiting the cytochrome P450 enzyme
14α-demethylase in the fungal cell wall, and toxic C-14α methylsterols start to
accumulate within the cell (Figure 3). Fluconazole is a well-tolerated antifungal and
widely used as a first line of systemic treatment, especially in HIV and AIDS patients
(Cha and Sobel 2004, Hamza et al. 2008). However, prolonged and repeated therapy
with a low dosage of fluconazole should be avoided, as the development of resistant
isolates and selection pressure towards NAC species have been reported (Rautemaa
et al. 2007b, Siikala et al. 2010). Itraconazole has a better susceptibility profile than
fluconazole (Denning and Hope 2010). It is absorbed well, particularly as an oral
solution, but sub-therapeutic drug concentrations are seen in patients with an
elevated gastric pH. It is secreted into mucous membranes, and a topical effect will
thus also be achieved. Multiple toxicities are associated with the use of itraconazole,
and this should be assessed before the therapy (Denning and Hope 2010).
The second-generation triazole antifungals voriconazole and posaconazole have a
broad spectrum of activity against a plethora of Candida spp. and moulds, including
fluconazole-resistant species and isolates (Pfaller et al. 2004). Voriconazole is
structurally related to fluconazole and posaconazole to itraconazole. Posaconazole is a
well-tolerated drug and oral bioavailability is especially enhanced by fatty food.
Voriconazole has various drug interactions and side effects that should be taken into
consideration. Immunomodulatory properties have been detected for triazoles:
fluconazole and voriconazole have shown synergistic activity with phagocytes against
C. albicans (Brummer and Stevens 1996, Garcha et al. 1995, Vora et al. 1998) and
fluconazole can drive protective T-cell responses in vivo (Cenci et al. 1997).
5.3.1.2.2 Echinocandins
Echinocandin class antifungals (caspofungin, micafungin, anidulafungin) can only be
administered intravenously. They act via non-competitive inhibition of β-1,3-D-glucan
synthetase in the fungal cell wall, which results in osmotic instability and cell lysis
(Figure 3). All three echinocandins are well tolerated, have broad-spectrum fungicidal
activity against Candida spp. and have few drug–drug interactions, which is a major
advantage over other antifungals (Cornely et al. 2012). However, resistance against
caspofungin has been reported, especially among C. glabrata isolates (Alexander et al.
2013). Caspofungin has shown promising efficacy against Candida biofilms, and has
thus been suggested as a therapeutic option in CVC-related infections (Andes et al.
2012). Caspofungin modulates immune responses. Exposure of C. albicans to
31
caspofungin results in β-glucan unmasking in the fungal cell wall, and this induces proinflammatory responses by macrophages (Lamaris et al. 2008, Wheeler et al. 2008).
5.3.1.2.3 Polyenes
There are currently three lipid formulations of conventional amphotericin B (AmB lipid
complex, ABLC, AmB colloidal dispersion, ABCD, and liposomal amphotericin B, LAmB), which were developed to enable systemic therapy and reduce toxicity. Lipid
AmBs have shown decreased renal toxicity compared to conventional AmB (AmB
deoxycholate) (Ashley et al. 2006). AmB antifungals have a very broad spectrum of
activity, with efficacy against Candida spp. and moulds. Conventional Amb and lipid
AmBs are suggested treatment options for device-related infections (Cornely et al.
2012). AmB and ABCD induce pro-inflammatory responses, but in contrast, ABLC and
L-AmB either attenuate or have no effect on inflammatory responses by immune
effector cells (Arning et al. 1995, Simitsopoulou et al. 2005).
5.3.1.2.4 Antifungal vaccines
One optional route for systemic therapy in infectious diseases is vaccination. The first
vaccines against bacterial and viral infections were developed over a few centuries
ago, and multiple vaccines currently exist against various bacterial and viral species.
No vaccines are currently on the market against fungal infections. Experimental
models have produced some promising results in this field. A Candida vaginitis model
using anti-β-glucan has shown success in mice (Pietrella et al. 2010). In addition, Als3p
adhesin and Sap2p invasin have been identified as important targets. Als3 antiCandida (rAls3-N) vaccine protected against oropharyngeal and systemic candidosis in
mice (Spellberg et al. 2006), while rSap2-N induced protection against vaginal
infection in rats (De Bernardis et al. 2012).
5.3.2
Suppressive therapy
In clinical scenarios where patients are at risk of developing a fungal infection, primary
or secondary antifungal prophylaxis may be an appropriate strategy to suppress
Candida overgrowth, transient colonisation and the risk of systemic disease
(Laudenbach and Epstein 2009, Rautemaa and Ramage 2011). The topical polyene
antifungals nystatin or AmB are recommended as the primary choice for prophylactic
treatment of oral candidosis in the prevention of recurrent infection. In HIV patients
with frequent and severe recurrent infections, prophylaxis can be considered (Kaplan
et al. 2009, Lodi et al. 2007). Fluconazole and itraconazole have been reported as
32
effective prophylactic antifungals for the prevention of systemic infection in intensive
care unit (ICU) patients with haematological or upper digestive tract malignancies or
AIDS (Cornely et al. 2003, Corvo et al. 2008, Playford et al. 2006, Vardakas et al. 2005).
Antifungal prophylaxis can be recommended for patients who have undergone
abdominal surgery and have perforations in the gut and anostomotic leakage (Cornely
et al. 2012). Posaconazole has been used as antifungal prophylaxis in autoimmune
polyendocrinopathy-candidosis-ectodermal dystrophy (APECED) patients with CMC.
Posaconazole was found to be more effective than fluconazole and itraconazole as a
prophylactic agent in neutropenic patients with leukaemia or in bone marrow
transplant recipients (Cornely et al. 2012, Ullmann et al. 2007).
Although promising outcomes have been observed with prophylactic treatment,
overall evidence for treatment efficacy remains limited and no drug has shown
superior potential. Controversial results have also been reported. Nystatin was not
effective in the prophylactic treatment of cancer patients (Clarkson et al. 2007,
Epstein et al. 2003). The choice of prophylactic agent is complicated by the multiple
adverse effects associated with the use of antifungal agents. If antifungal prophylaxis
is considered, the agent should be carefully chosen and the risk for the development
and selection of resistant strains should be kept in mind, especially when using
fluconazole. In addition, the predisposing factors for candidosis and oral hygiene
should be well addressed and managed beforehand.
5.3.3
Probiotics as an alternative approach
Candida colonisation in health is suppressed by competitive inhibition with
surrounding bacteria. The idea behind probiotic therapy is based on this model and
th
was first presented in the early 20 century by Elie Metchnikoff (Metchnikoff 1908).
Probiotics are defined as health-promoting live microbial supplements that are nonpathogenic and normally belong to the human microbiome. Probiotics can stabilise
the balance within the microbiome by reducing the overgrowth of pathogenic
microbes. Probiotic bacteria have been industrially applied to enhance processes such
as fermentation and to protect food products as preservatives. Micro-organisms used
as probiotics mainly belong to the group of lactic acid bacteria, such as Lactobacillus
spp. and Bifidobacterium spp., but other bacteria and also yeasts, such as
Saccharomyces boulardii, have additionally been used. Many beneficial effects have
been reported, particularly in the treatment of vulvovaginal candidosis and in the
suppressive therapy in the gut (Manzoni et al. 2006, Reid et al. 2003, Wagner et al.
1997). Probiotics can reduce Candida overgrowth and suppress candidal colonisation
of the oral cavity in vivo (Elahi et al. 2005, Hatakka et al. 2007). However, evidence
regarding the benefits of long-term use of probiotics in humans is still limited, and this
33
therapy can have adverse outcomes. The use of probiotic micro-organisms can lead to
bacteraemia and sepsis in immunocompromised patients, and caution is thus
warranted (Salminen et al. 2004).
5.3.4
Alternative anti-biofilm agents
The treatment of oro-oesophageal candidosis with current antifungals is often
challenged by the biofilm lifestyle and poor patient compliance due to multiple side
effects and/or toxicity. In addition to the mucosal surface structures, abiotic surfaces
in various medical devices provide an excellent surface for biofilm formation, as
highlighted in CVC-related infections. In the oral cavity, various biomaterial surfaces
exist in the form of dentures, fillings and implants. Multiple approaches are currently
under investigation to find more effective ways to eradicate biofilm infections.
Xylitol, a sugar alcohol, has shown beneficial effects on oral health and has been
widely used in oral health products and other consumables. It reduces dental plaque
formation and acid production by oral streptococci, particularly the cariogenic
pathogen Streptococcus mutans, and this inhibits the development of dental caries
(Iwata et al. 2003, Lee et al. 2009, Söderling and Hietala-Lenkkeri 2010). S. mutans and
C. albicans function synergistically in dental plaque formation, and increased
pathogenesis in the form of dental caries has been shown in vivo (Falsetta et al. 2014).
Xylitol could potentially act against this synergism. Chlorhexidine showed better in
vitro effect than fluconazole against C. albicans biofilms when the two compounds
were impregnated into denture acrylic in vitro (Salim et al. 2013b). Farnesol is a
quorum-sensing molecule produced by C. albicans that inhibits yeast–hyphal
transition (Hornby et al. 2001). It has shown efficacy against biofilms in vitro, but
contradictory outcomes have been observed in in vivo models (Hisajima et al. 2008,
Navarathna et al. 2007, Ramage et al. 2002b). Plants have effective strategies to fight
against fungi, and numerous plant compounds have been investigated for their
potential anti-Candida activity. Some of them, including tea tree oil, terpenoids and
eugenol, have shown promising efficacy against Candida biofilms in vitro (Khan and
Ahmad 2012, Ramage et al. 2012, Sudjana et al. 2012). However, the in vivo efficacy
remains unknown.
5.4
ETHANOL AND ACETALDEHYDE METABOLISM IN THE HUMAN
BODY
In general, the liver is the main organ responsible for ethanol metabolism.
Approximately 75–90% of ethanol oxidation is carried out by the liver in health, but
34
the level of oxidation can be substantially lower in liver disease, such as hepatic
cirrhosis (Agarwal and Goedde 1990). When alcohol is ingested, the bolus travels
through the upper digestive tract and stomach into the small intestine, where the
majority of ethanol (70–80%) is absorbed via simple diffusion without specific
transport mechanisms. It rapidly passes through the membranes because of its small
molecular size and high water solubility and enters the bloodstream, mainly the portal
vein carrying blood to the liver. Ethanol is evenly distributed via the circulation and
diffusion into the water content of human body, including saliva.
Ethanol metabolism also occurs in the digestive tract (extra-hepatic). Part of the
ingested ethanol is directly absorbed and metabolised by the mucosal and microbial
cells in the oral cavity, oesophagus and stomach. Studies comparing intravenous and
oral administration of ethanol have demonstrated that the blood alcohol level is
significantly lower after oral ingestion, thus proving the existence of first-pass
metabolism (Julkunen et al. 1985). Severe liver injury can induce first-pass metabolism
and account for as much as 40–50% of total ethanol metabolism (Dam et al. 2009,
Utne and Winkler 1980). Extra-hepatic metabolism is of considerable interest, as
ethanol is known to contribute to various pathological conditions in the upper
digestive tract.
Ethanol is oxidised in mucosal, microbial and hepatic cells, yielding acetaldehyde
(ACH) as the first and major metabolite. The majority of ACH is formed by various
alcohol dehydrogenase (ADH) enzymes in both human and microbial cells. In addition,
the cytochrome P450 2E1 (CYP2E1)-dependent microsomal ethanol oxidizing system
(MEOS) plays a role in ethanol oxidation in humans. Peroxisomal catalase can oxidise
ethanol, but the reaction is limited by the bioavailability of hydrogen peroxide, thus
contributing only minimally to ethanol oxidation in the liver (Oshino et al. 1973). Toxic
ACH is further metabolised by aldehyde dehydrogenase (ALDH) enzymes into a less
harmful compound, acetate. Finally, acetate is eliminated by the human body through
oxidation to CO2 in the heart, skeletal muscle and brain cells. Acetate can also be
further metabolised in human or microbial cells into acetyl coenzyme A (Acetyl-CoA)
and used as an energy source or utilised for the biosynthesis of cell components such
as lipids and cholesterol. In addition to the exogenous sources of ethanol, the human
body and mainly the GI mucosa can be exposed to endogenously produced ethanol
via microbial metabolism. At a low oxygen tension, microbes produce energy through
fermentation and ethanol is formed via ACH. Patients suffering from achlorhydric
atrophic gastritis have shown microbial overgrowth and endogenous ethanol and ACH
production (Väkeväinen et al. 2002). However, not all fermentative processes yield
ethanol as an end product. Some human cells produce lactate in anaerobiosis. Certain
35
bacterial species produce a spectrum of short-chain fatty acids, including butyrate,
acetate and succinate, in their fermentative metabolism.
5.4.1
5.4.1.1
Ethanol oxidation in humans
Alcohol dehydrogenase (ADH)
Human ADH enzymes are located in the cytosolic compartment of the cell. The
oxidative process catalysed by ADH involves intermediate carriers of electrons,
+
+
nicotine adenine dinucleotide or its phosphate analogue (NAD or NADP ), which are
reduced by two electrons to NADH or NADPH in the reaction. Human ADH enzymes
can be divided into five classes based on their main characteristics (Holmes 1994).
Class I ADH enzymes (ADH 1) have a low Km value for ethanol (<4 mM). Classes II
(ADH 2), IV (ADH 4) and V (ADH 5) have moderate Km values for ethanol (<40 mM),
and class III (ADH 3) is hardly saturated at all (Km > 3 M). A total of seven genes
(ADH1A, ADH1B, ADH1C, ADH2-5) have been identified to encode various ADH
enzymes (Osier et al. 2002). The three major isoenzymes of class I encoded by ADH1A,
ADH1B and ADH1C play a substantial role in ethanol oxidation in humans, and their
polymorphisms have been linked to disease (Bosron and Li 1986). In addition, ADH
enzymes have a significant role in the oxidation of other excess alcohol compounds
such as retinol, which is needed for retinoic acid biosynthesis (Duester et al. 1999).
5.4.1.2
Microsomal ethanol oxidising system (MEOS)
The MEOS is located in the endoplasmic reticulum and characterised by the presence
of cytochrome P450 (CYP) enzymes. In humans, multiple CYP enzymes oxidise ethanol,
including CYP2E1, -1A2 and -3A4 (Koop 1992, Niemelä et al. 1999). However, CYP2E1
has the highest affinity for ethanol of all CYP enzyme family members (Koop 1992).
CYP2E1 can also oxidise other low molecular weight molecules, such as carcinogenic
nitrosamines, aflatoxins and polycyclic hydrocarbons, and ethanol-induced CYP2E1
metabolism has therefore been linked to other toxic effects (Koop 1992). Normally,
CYP2E1-dependent MEOS accounts only for a small percentage of ethanol metabolism
due to its high Km (7–10 mM), but the activity can be induced by up to 10–20 fold in
chronic alcohol use (Oneta et al. 2002). However, individual variation exists in the
level of induction. During CYP2E1-mediated ethanol metabolism, ROS are generated
(Lieber 1997). This leads to the formation of lipid peroxidation products, which can be
genotoxic due to protein adduct formation and cause damage comparable to that of
ACH (Aleynik et al. 1998, Niemelä et al. 1999).
36
5.4.1.3
Aldehyde dehydrogenase (ALDH)
Ten different genes (ALDH1-10) encode ALDH enzymes in humans (Yoshida et al.
+
1998). ALDH requires NAD in the oxidation of ACH into acetate, thus further
decreasing the redox state of the cell. The cellular localisation of ALDH enzymes is
more varied compared to ADH enzymes. Mitochondrial, cytosolic and microsomal
isoforms have been identified. Different ALDH enzymes can be divided into groups
according to their Km values: ALDH1 and ALDH2 isoenzymes have low Km values (0.2–
180 µM) and ALDH3 and ALDH4 are isoenzymes with a high Km value (5–83 mM)
(Jelski and Szmitkowski 2008, Klyosov et al. 1996). Cytosolic ALDH1 and particularly
mitochondrial ALDH2 have a high affinity to ACH and play a major role in ACH
metabolism. The rest of the ALDH enzymes do not contribute substantially to ACH
elimination due to their high Km values and low affinity for ACH, and their role in ACHrelated pathogenesis is thus minimal (Crabb et al. 2004). They are mainly responsible
for the elimination of other toxic aldehydes.
5.4.2
Hepatic metabolism
ADH 1 and its isoenzymes play a substantial (95% of total activity) role in ethanol
oxidation in the liver. Their low Km values underline this property. The redox state, i.e.
+
NADH/NAD ratio, regulates metabolism in the liver. A reduction in the redox status
due to ethanol oxidation can impair other metabolic functions of the liver, such as
gluconeogenesis, fatty acid oxidation and the activity of the tricarboxylic acid (TCA)
cycle. Ethanol turnover into ACH, i.e. the rate of ethanol elimination, is mainly
regulated by the rate of NADH re-oxidation. Blood ACH levels and ethanol oxidation in
the liver can be significantly affected by polymorphisms. Major mutations appear in
two ADH genes: ADH1B and ADH1C (Crabb et al. 2004). The ADH1B*2 allele encodes a
40 times more active isoenzyme than ADH1B1*1, which leads to ACH accumulation in
the blood. A similar intolerance of alcohol has been described as in individuals
heterozygous for ALDH2*2 (Thomasson et al. 1991). Caucasians rarely have ADH1B*2,
but it is more frequent in Asians. ADH1C*1 increases ethanol metabolism by 2.5 fold
compared with the other allele, ADH1C*2 (Bosron and Li 1986). The major role of
ALDH2 in ACH metabolism has been supported by studies on ALDH2-deficient
individuals. Two allelic forms exist for the ALDH2 isoenzyme: the active ALDH2*1 and
highly inactive ALDH2*2. Individuals homozygous for ALDH2*2 have no activity and
persons heterozygous for ALDH2*2 retain approximately 6% of the activity (Crabb et
al. 1989). Between 40–50% of Asians have the deficient form of ALDH2. Homozygotes
cannot tolerate alcohol due to adverse effects (nausea, facial flushing and headache)
associated with high ACH concentrations in blood. Heterozygotes can tolerate alcohol
better, and they are consequently at greater risk of ACH-related pathogenesis.
37
5.4.3
5.4.3.1
Extra-hepatic metabolism
Metabolism in the upper digestive tract
Mucosal cells in the oral cavity and oesophagus are challenged with ethanol and ACH
via saliva and blood. A study by Dong et al. demonstrated the presence of high Km
ADH 4, but a lack of low Km ALDH2 and ALDH1 enzymes in gingival and tongue
biopsies (Table 3; Dong et al. 1996). Thus, it was suggested that cytotoxic ACH may be
involved in the aetiology of alcohol-related oral injury. Yin et al. compared the
expression and activity of ADH and ALDH enzymes of the oesophageal and gastric
mucosa and detected 4-fold higher activity of high Km ADH enzymes and 4-fold lower
activity of the low Km ALDH2 enzyme in the oesophageal mucosa compared to the
stomach (Yin et al. 1993, Yin et al. 1997). Both oral and oesophageal tissue samples
lacked highly active ALDH2 (Table 3; Chiang et al. 2012). CYP2E1 is highly present in
the oral and oesophageal mucosa (Godoy et al. 2002, Lechevrel et al. 1999, Liu et al.
2001b). Interestingly, CYP2E1 was shown to be expressed in pre-cancerous and cancer
lesions of the oral cavity (Warnakulasuriya et al. 2008a). Overall, these findings
indicate that oral, oesophageal and gastric mucosal cells are capable of oxidizing
ethanol, but their ability to eliminate ACH is either lacking or limited, especially among
subjects with a genetically deficient ALDH2 enzyme.
Table 3. Mean ADH and ALDH activity (mU) relative to the amount of tissue and protein
present in the oral cavity and oesophagus compared to the liver (modified from Chiang et al.
2012; nd. = non-detectable).
38
5.4.3.2
Microbial metabolism
In subjects with the normal ALDH2 enzyme, ACH levels in the peripheral blood remain
under the detection limit, even after high doses of ethanol (Stowell 1979).
Significantly elevated salivary ACH levels after alcohol intake can, however, be
detected in ALDH2-deficient individuals. After alcohol consumption, ethanol levels in
the saliva are equal to those in the blood (Jones 1979). Significant ACH levels are
detected in saliva after moderate alcohol intake, and these correlate with salivary
ethanol levels (Homann et al. 1997, Jokelainen et al. 1996). In ALDH2-deficient
subjects, a dose of alcohol results in 2–3 times higher ACH levels in saliva compared to
individuals with the normal ALDH2 enzyme (Väkeväinen et al. 2000, Yokoyama et al.
2008). This is partly caused by oxidation of ethanol to ACH in the parotid glands
(Väkeväinen et al. 2000).
Main determinant for salivary ACH is microbial metabolism (Homann et al. 1997,
Väkeväinen et al. 2001). Microbial ACH production has been shown to occur in the
oropharynx in vivo (Pikkarainen et al. 1981). ACH levels detected in saliva were
significantly lower after a chlorhexidine mouthwash rinse, thus providing strong
support for the microbial origin of ACH. No correlation was found between microbial
counts and salivary ACH levels, thus indicating that a qualitative change in microbial
composition has an effect on salivary ACH levels (Homann et al. 1997). The highest
salivary ACH levels are seen immediately after alcohol intake at very high local ethanol
concentrations (Linderborg et al. 2011). Poor oral hygiene correlates with high salivary
ACH levels, and a link between facultative bacteria and yeasts, particularly C. albicans,
with high salivary ACH levels has been shown (Homann et al. 2001, Yokoyama et al.
2007). Altogether, these findings suggest that the microbiome plays a substantial role
in local ACH exposure in the upper digestive tract.
5.4.3.2.1 Bacterial metabolism
Under a low oxygen tension, bacterial cells produce energy through fermentation and
ethanol is formed. Endogenous ethanol production by the local microbiome has been
measured in vivo (Cope et al. 2000). Bacterial alcohol dehydrogenase (adh) can
function in a reversible fashion, and in aerobic conditions bacterial cells oxidise
ethanol to ACH (Maconi et al. 1988). Aerobic oral Streptococcus, Corynebacterium and
Stomatococcus species have previously been linked to high ACH levels detected in
saliva (Homann et al. 2001, Yokoyama et al. 2007). Oral Streptococci spp. and
Neisseria spp. can produce high amounts of ACH from ethanol in vitro and possess a
highly functional adh with a low Km for ethanol (Kurkivuori et al. 2007, Muto et al.
39
2000, Pavlova et al. 2013, Väkeväinen et al. 2001). Pavlova et al. investigated ADH
expression in S. gordonii and demonstrated that three functional ethanol oxidising
enzymes exist and one, namely adhE, with a high activity (Pavlova et al. 2013). Other
enzymes were also identified that showed substrate specificity for other alcohols.
Interestingly, no functioning aldehyde dehydrogenase (ald) enzyme was identified for
any of the Streptococcal or Neisseria species investigated (Muto et al. 2000, Pavlova
et al. 2013). Fully functional alds have been identified in bacteria of the gut, but the
elimination of ACH by these enzymes was lower than alcohol dehydrogenasedependent ACH formation, which could contribute to ACH accumulation (Nosova et al.
1998, Nosova et al. 1996). Interestingly, probiotic Lactobacillus spp. and
Bifidobacterium spp. possess an active enzyme that efficiently removes ACH (Nosova
et al. 2000). This indicates that probiotics could have protective effects in addition to
the ones already observed. Catalase is found in aerobic and facultatively anaerobic
bacteria, and inhibition of catalase decreased the ACH level produced in vitro. Thus,
alternative pathways for oxidation exist in bacteria (Tillonen et al. 1998). The way of
life of various commensal and pathogenic bacteria in the oral and oesophageal
microbiome is highlighted by the community-oriented biofilm lifestyle. When a
metatranscriptomic approach was applied to investigate the metabolic patterns in the
oral microbiome, disease-associated communities showed similar metabolic profiles,
and genes encoding enzymes in pyruvate and ethanol metabolism were highly
expressed (Jorth et al. 2014). Both oesophageal and oral microbiomes, including the
salivary microbiome, are highly dominated by Streptococcus spp. (Segata et al. 2012,
Yang et al. 2009). Neisseria spp. are also among the highly prevalent species in these
microbiomes in healthy subjects. Therefore, the upper digestive tract is colonised by
bacteria that can efficiently oxidise ethanol, but their ability to remove ACH appears
limited. Thus, local ACH accumulation can occur. However, environmental factors
have to be taken into account. Nutrients and oxygen levels, together with the
composition of the bacterial community, vary considerably within each niche and
have a major effect on the metabolome and the potential ACH exposure locally.
5.4.3.2.2 Candida metabolism
Candidal metabolism is largely driven by carbohydrate metabolism. In general, yeasts
favour sugars as their primary and preferred energy source (Askew et al. 2009). Our
knowledge of the carbohydrate metabolism of eukaryotic cells mostly derives from
studies on Saccharomyces cerevisiae, referred to as conventional yeast. S. cerevisiae
has been widely used for industrial purposes due to its high tendency for
fermentation. A universal theme in yeast carbohydrate metabolism is the conversion
of 5- or 6-carbon sugars to either glucose-6-phosphate (G-6-P) or fructose-6-
40
phosphate (F-6-P), which are then converted to pyruvate in glycolysis (Figure 4;
Barnett 2003, Flores et al. 2000). The fate of pyruvate varies depending on the
availability of oxygen and the type of yeast.
Figure 4. A schematic diagram of central carbon metabolism and fermentative pathway in
Candida albicans (abbreviations as follows: F6P = fructose-6-phosphate, G6P = glucose-6phosphate). The solid line denotes the central carbon metabolism in aerobiosis and the dashed
line indicates the pathway of fermentative metabolism in anaerobiosis. The major genes
regarding ethanol metabolism are shown.
Eukaryotic cells typically prefer respiration to produce energy when oxygen is present
(normoxia). In respiration, acetyl-CoA derived from pyruvate is utilised by the TCA
cycle, where it is oxidised to carbon dioxide (CO2). NADH and FADH2 are formed in the
reductive process. Energy is produced in the form of adenosine triphosphate (ATP)
through oxidative phosphorylation, where the two electron carriers NADH and FADH2
are re-oxidised. At a low oxygen tension (hypoxia), eukaryotic cells produce energy
through fermentation. In yeasts, pyruvate is transformed into ethanol in the reductive
+
process along the fermentative pathway. Respiration is preferred, as more NAD and
ATP per glucose can be produced in oxidative phosphorylation than in fermentation.
S. cerevisiae is a Crabtree-positive yeast that prefers fermentation for energy
production and represses respiration even when carbohydrates and oxygen are
available. In contrast, C. albicans and other non-conventional yeasts are defined as
Crabtree-negative yeasts that oxidise pyruvate to CO2 to produce energy under
aerobic conditions via the TCA cycle, and fermentation is only used under hypoxic
conditions. To further underline these differences, highly distinct post-transcriptional
mechanisms regarding carbon assimilation exist in C. albicans and S. cerevisiae (Sandai
et al. 2013). Importantly, this section mainly focuses on the metabolism of C. albicans.
Multiple transcriptomic studies have highlighted the effect of oxygen levels on
metabolism. In C. albicans, genes belonging to the glycolytic pathway, ergosterol
41
synthesis and fermentation are up-regulated under hypoxic conditions, while the
expression of genes involved in oxidative metabolism is down-regulated (Ernst and
Tielker 2009, Setiadi et al. 2006, Synnott et al. 2010). Before entering the TCA cycle,
pyruvate is decarboxylated to acetyl-CoA in the mitochondrial pyruvate
dehydrogenase complex (CaPDH) (Figure 4; Flores et al. 2000, Pronk et al. 1996). The
equilibrium between respiration and fermentation is transcriptionally regulated by
CaGal4p and CaTye7p in C. albicans, depending on the nutrient source (Askew et al.
2009). These master regulators are vital for normal growth and virulence, and they
both activate genes linked to fermentation. However, the CaPDH-dependent
decarboxylation step is only controlled by Gal4p.
In the fermentative pathway, pyruvate is first converted to ACH by pyruvate
decarboxylase, which is encoded in C. albicans by CaPDC11. In yeasts, as well as in
humans, ACH is a metabolite that is located at the point where the fermentative
pathway and pyruvate bypass pathway cross (Figure 4; Flores et al. 2000). In
fermentation, ACH is metabolised to ethanol by ADH, and in the process NADH is
+
oxidised to NAD to maintain glycolysis. Glycerol production supports the redox
+
balance by the formation of NAD in anaerobiosis. Importantly, glycerol metabolism is
also associated with osmotic stress resistance and biofilm formation (Desai et al.
2013, Flores et al. 2000). CaAdh1p is a bi-directional enzyme in C. albicans. Thus,
ethanol can be used as both a carbon and an energy source (Bertram et al. 1996). In
pyruvate bypass, ACH is metabolised to acetate by the enzyme Ald, followed by the
metabolism of acetate to acetyl-CoA by the enzyme acetyl CoA synthetase (Acs).
Acetyl-CoA is transported into mitochondria by a carnitine-dependent mechanism,
+
and NADH is re-oxidised to NAD during the process. This route bypasses
mitochondrial acetyl-CoA formation by the PDH complex. Fermentation is upregulated in hyphal and biofilm formation and during colonisation (Hernaez et al.
2010, Martinez-Gomariz et al. 2009, Monteoliva et al. 2011, Pierce et al. 2013). In
various niches of the human host, fungal cells are often faced with hypoxic conditions.
Therefore, fermentative metabolism is vital for adaptation and survival.
A total of five alcohol dehydrogenase genes (CaADH1-5) exist in the C. albicans
genome (Inglis et al. 2012). CaADH1 and CaADH2 show significant homology to
corresponding genes in S. cerevisiae (75–77%; Inglis et al. 2012). CaAdh1p is the only
isoenzyme that has been studied in depth. The protein functions either in the cytosol
or cell wall (Crowe et al. 2003, Mukherjee et al. 2006). CaAdh1p mainly converts ACH
to ethanol and is expressed in the exponential and stationary growth phases of
planktonic growth (Bertram et al. 1996, Kusch et al. 2008, Swoboda et al. 1994).
mRNA levels of CaADH1 do not correlate with the protein levels, thus suggesting the
existence of post-transcriptional regulation (Bertram et al. 1996). Previous
42
transcriptomic and proteomic studies have indicated that CaADH1 or CaAdh1p has a
regulatory role in biofilm formation and growth (Martinez-Gomariz et al. 2009,
Mukherjee et al. 2006, Thomas et al. 2006). A limited amount of information is
available on the functions and role of other CaAdhs in C. albicans. S. cerevisiae
ScAdh2p functions under glucose-repressed conditions and preferably converts
ethanol to ACH (Maestre et al. 2008). In a transcriptomic study by Setiadi et al. (2006)
CaADH1, CaADH2 and CaADH5 were up-regulated in hypoxia. CaADH5 regulates
matrix formation during biofilm development (Nobile et al. 2009).
The number and functional role of CaAld enzymes in C. albicans is not well known.
There are five genes in the S. cerevisiae genome (Cherry et al. 2012). ScALD4 and
ScALD6 encode the major mitochondrial and cytosolic isoforms, respectively. ScALD5
encodes the minor mitochondrial isoform in S. cerevisiae and is homologous to
CaALD5 (61%; Inglis et al. 2012). ScALD2 and ScALD3 in S. cerevisiae encode cytosolic
isoenzymes that do not participate in acetate formation (Saint-Prix et al. 2004).
ScAld5p and ScAld6p have a major role in acetate formation in S. cerevisiae (Saint-Prix
et al. 2004). CaAld5p is expressed in C. albicans biofilms during oxidative stress and
hypoxia (Martinez-Gomariz et al. 2009, Seneviratne et al. 2008, Strijbis et al. 2009).
ACH accumulation has been reported to be regulated in planktonic cultures during
hypoxia by the downstream enzymes CaAld6p and CaAcs1p in C. albicans (Marttila et
al. 2013a). CaAld4p has a role in carnitine biosynthesis (Strijbis et al. 2008). Carnitinedependent transport of acetyl-CoA is required for growth on non-fermentable carbon
sources (ethanol, acetate) and for biofilm development (Strijbis et al. 2008). The role
of acetyl-CoA synthetases in C. albicans has been extensively studied, and two genes
have been identified: CaACS2 encodes the major isoform, which is vital for growth on
most carbon sources, and CaACS1 is required for the full utilisation of alternative
carbon sources (Carman et al. 2008).
Although carbohydrates are the preferred source, metabolic flexibility is needed for
survival and virulence. When carbohydrates are not available, utilisation of alternative
carbon sources, such as ethanol and acetate, is required for the biosynthesis of critical
cellular components. The glyoxylate cycle, a metabolic pathway partly overlapping
with the TCA cycle, enables the use of two-carbon sources in gluconeogenesis and is
critical for virulence in C. albicans (Figure 4; Ramirez and Lorenz 2007). During the
early steps of phagocytosis, glycolysis is repressed and the glyoxylate shunt and
gluconeogenesis are up-regulated. However, later survival is dependent on glycolysis
(Barelle et al. 2006). It has also been shown that the availability of glucose regulates
the stress responses in C. albicans, thus underlining the importance of central carbon
metabolism in fungal pathogenesis (Rodaki et al. 2009).
43
5.5
5.5.1
ACETALDEHYDE IN CARCINOGENESIS OF THE UPPER DIGESTIVE
TRACT
Ethanol-related toxicity
Ethanol (ethyl alcohol, CH3CH2OH) is a highly toxic compound when used in large
amounts. Alcohol abuse has been linked to various disorders and conditions, including
liver disease, pancreatic disorders, gastritis, cardiovascular diseases, mental disorders,
depression and impairment of metabolism and malnutrition (Room et al. 2005).
Ethanol can increase the permeability of the gut by modulating tight junctions
(Basuroy et al. 2005, Bjarnason et al. 1984, Robinson et al. 1981). In addition, it can
induce adverse changes in the gut microbiome, and these two factors combined can
lead to the translocation of pathogenic microbes into the circulation (Parlesak et al.
2000, Rao 2009). This can also result in extra-intestinal organ involvement and an
increase in inflammation. Although a clear association has been found between
chronic alcohol consumption and carcinogenesis of the upper digestive tract, liver,
colorectum and female breast, a debate exists concerning the mechanisms behind the
carcinogenicity of ethanol. Multiple potential mechanisms have been described,
including direct damage to mucosal cells (Bode et al. 1982), increased oxidative stress
and lipid peroxidation (Niemelä et al. 1999), enhanced solubility of carcinogens (Wight
and Ogden 1998), increased cell membrane permeability due to lipid dissolution
(Squier et al. 2003), induction of detoxifying enzymes and activation of procarcinogens (Stickel et al. 2002), immunomodulation (Szabo 1999), an increase in
oestrogen levels in women (Coutelle et al. 2004) and interference with folate and
retinoic acid metabolism and synthesis (Giovannucci et al. 1995, Halsted et al. 2002,
Liu et al. 2001a). Although these various harmful direct or indirect effects of ethanol
exist, ethanol itself is not genotoxic (Ellahuene et al. 2012, Phillips and Jenkinson
2001). However, its first metabolite, ACH, is.
5.5.2
Acetaldehyde-related toxicity and carcinogenicity
ACH (ethanal, CH3CHO) is a highly volatile, cytotoxic and mutagenic compound
(Dellarco 1988). A carcinogenic effect has been shown in animals (Feron et al. 1982,
Woutersen et al. 1986). Multiple epidemiological and biochemical studies have
supported its carcinogenicity in humans (Boccia et al. 2009, Homann et al. 1997).
Recently, ACH linked to alcohol abuse was ranked as a group I carcinogen by IARC
(Secretan et al. 2009). This was largely due to evidence derived from animal
experiments and particularly from epidemiologic studies, which have shown the
strong association between the GI tract cancer risk and ALDH2 deficiency in humans.
A small amount of ACH may be formed during threonine metabolism in the human
44
body, but the majority is derived from ethanol metabolism in the liver and gut (Lin
and Greenberg 1954). However, the highest levels of ACH can be found in saliva due
to microbial metabolism, and this might explain why ethanol-induced cancers are
located in the upper digestive tract (Väkeväinen et al. 2000, Yokoyama et al. 2008).
The mutagenic effect of ACH has been shown to occur at concentrations as low as 40–
100 µM, which can be found in saliva after moderate alcohol drinking (Brooks and
Theruvathu 2005, Homann et al. 1997, Lachenmeier and Monakhova 2011, Linderborg
et al. 2011, Seitz and Stickel 2007). In addition to its presence in the human body, it is
widely present in foodstuffs, beverages, industrial processes and the environment
(Feron et al. 1991, Salaspuro 2009). When consumed, the oral cavity and oesophagus
are the first sites in direct contact with exogenous ACH. ACH is the most abundant
carcinogen in tobacco smoke and it dissolves in saliva during smoking (Salaspuro and
Salaspuro 2004, Hoffmann et al. 2001). The ACH content of alcoholic and nonalcoholic beverages is dependent on the manufacturing process: fermentation,
distillation and preservation. The highest ACH content has been detected in calvados,
sherry and fruit spirits (Boffetta et al. 2011, Lachenmeier et al. 2009, Linderborg et al.
2008).
Multiple direct or indirect genotoxic effects related to ACH have been described
(Figure 5). ACH interferes with DNA synthesis and repair (Espina et al. 1988, Garro et
al. 1986). Cytogenetic changes observed in vitro include sister chromatid changes and
gross chromosomal aberrations (Maffei et al. 2000, Maffei et al. 2002, Obe et al.
1986). ACH forms stable DNA adducts, and significant levels have been measured in
multiple organs and conditions including oral pre-malignant and malignant lesions due
to chronic alcohol abuse (Fang and Vaca 1995, 1997, Warnakulasuriya et al. 2008a).
2
The most abundant and best-studied stable ACH adduct is N -ethyl-2’-deoxyguanosine
(N-EdG), which is found at damaging levels in the oral cavity within 4 to 6 hours of
alcohol exposure in humans (Balbo et al. 2012). ACH adducts can further generate
DNA-strand cross-links and cause DNA-strand breaks when repaired (Matsuda et al.
1998). ACH-related adduct formation has been shown to occur with lipid peroxidation
products, such as malondialdehyde, as a result of oxidative stress due to ROS (Niemelä
et al. 1999, Niemelä et al. 1995).
45
Figure 5. Schematic diagram of the major mechanisms of acetaldehyde-related mutagenesis.
The main enzymes related to acetaldehyde metabolism are shown: alcohol dehydrogenase (ADH),
aldehyde dehydrogenase (ALDH) and cytochrome P450 2E1 (CYP 2E1) within the microsomal
ethanol oxidising system (MEOS).
5.5.3
Aetiology and epidemiology of upper digestive tract cancers
Oral (lip, oral cavity, oro- and hypopharynx) and oesophageal carcinomas are among
the ten most common cancers globally (Ferlay et al. 2013). There is large geographic
and demographic variation in the incidence worldwide. In Finland, approximately 300
new cases of oral and pharyngeal cancers with an annual incidence of ~7/100,000 and
200 new cases of oesophageal cancer with an annual incidence of ~3/100,000 are
diagnosed every year among men (Finnish Cancer Registry). Among women, the
numbers are lower: 200 new cases of oral and pharyngeal cancers and 40 new cases
of oesophageal cancer are diagnosed in Finland every year (annual incidence of ~4
and ~1/100,000, respectively). In contrast, among the European Union countries,
Hungary has the highest incidence of oral cancer (40/100,000) and the United
Kingdom of oesophageal cancer (15/100,000) among men (Ferlay et al. 2013).
Squamous cell carcinoma (SCC) is the most common histological subtype of
oesophageal and oral cancer worldwide and the type mainly associated with ethanolrelated carcinogenesis (Scully and Bagan 2009, Steevens et al. 2010). However, the
incidence of oesophageal adenocarcinoma has rapidly increased and now exceeds the
incidence of oesophageal SCC (ESCC) in some Western countries, including Finland
(Pohl and Welch 2005). Despite advances in surgical techniques, radiation and
chemotherapy, mortality and morbidity rates remain high, with 5-year survival rates
of only 40–50% and 15–25% for oral and oesophageal cancer, respectively (Enzinger
and Mayer 2003, Warnakulasuriya 2009). This is mainly due to delayed diagnosis and
46
treatment. The incidence of oral and oesophageal cancer increases as a function of
age. However, an increased incidence of oral cancer has been lately been noted
among youths (van Monsjou et al. 2013).
5.5.3.1
Main risk factors
Squamous cell carcinomas of the oral cavity and oesophagus have very similar risk
factor profiles. Tobacco smoking and alcohol abuse are the major risk factors.
Approximately 30% of oral cancers and 25% of oesophageal cancers may be attributed
to the consumption of alcohol (Boffetta et al. 2006). Alcohol drinking and smoking
have a synergistic effect on upper digestive tract carcinogenesis (Lee et al. 2007,
Pelucchi et al. 2008). They both result in increased salivary ACH levels in vitro and in
vivo (Homann et al. 1997, Salaspuro and Salaspuro 2004, Tillonen et al. 1999). The risk
of carcinogenesis in the upper digestive tract is increased by poor oral health (Abnet
et al. 2008, Guha et al. 2007, Sepehr et al. 2005, Velly et al. 1998, Wei et al. 2005). As
stated above, additional support for the major role of alcohol abuse and ACH in
cancer aetiology is provided by mutations in the genes responsible for ethanol
metabolism (ADH1B, ADH1C, ADH2, ADH4, ALDH2, CYP2E1). These polymorphisms
have been associated with a high incidence of cancers of the oesophagus, oral cavity,
pharynx and larynx (Guo et al. 2008, Hashibe et al. 2006, Hashibe et al. 2008, McKay
et al. 2011, Oze et al. 2009, Tanaka et al. 2010). Other important risk factors for oral
cancer are a low socioeconomic status, nutritional deficiencies, ultraviolet and ionizing
radiation, sexual behaviour and HPV infection (Scully and Bagan 2009). Chewing of
betel quid increases the risk of both oral and oesophageal cancer, and a synergistic
effect exists with tobacco smoking (Petti 2009, Wen et al. 2010). In addition, regular
consumption of microbially fermented food products and beverages with a high ACH
and ethanol content has been linked to the aetiology of upper digestive tract cancers
(Lachenmeier and Sohnius 2008, Linderborg et al. 2008)
A dysbiotic microbiome has been linked to inflammation and malignant
transformation in the upper digestive tract (Hooper et al. 2009). Lower microbial
richness was found in lesions showing dysplasia in the oesophagus (Yu et al. 2014).
Specific Streptococcal and periodontopathogenic species have been frequently
identified in oesophageal cancer lesions (Narikiyo et al. 2004). Bacteria produce
metabolites that can modulate and damage the host genome (Schwabe and Jobin
2013). Oral carcinoma specimens showed similarities in their microbiome with certain
combinations of aerobic and anaerobic bacteria and C. albicans (Nagy et al. 1998).
Besides being a significant risk factor for cancer, smoking can alter the microbiome.
The microbiome of the oral cavity has been found to significantly differ between
47
smokers and non-smokers (Morris et al. 2013). Adherence, growth and biofilm
formation by S. aureus and C. albicans was enhanced after exposure to tobacco smoke
(Kulkarni et al. 2012, Semlali et al. 2014). In addition to their role in the production of
carcinogenic ACH, bacteria can activate carcinogenic nitrosamines of tobacco smoke
(Ahn et al. 2012, Shapiro et al. 1991).
5.5.3.2
Candida and cancer
The association of Candida with pre-malignant lesions and the potential cancerinducing role was already reported in the 1960s (Cawson 1969). More recently,
multiple studies have reported an association between Candida and mucosal dysplasia
(Barrett et al. 1998, McCullough et al. 2002, Nagy et al. 1998, O'Grady and Reade
1992, Rindum et al. 1994). Oesophageal carcinoma has been shown to be associated
with candidosis (Scott and Jenkins 1982). Several case reports have demonstrated the
association between CMC and oral and oesophageal carcinomas (Delsing et al. 2012,
Rautemaa et al. 2007a, Rosa et al. 2008). In a Finnish patient population, the majority
of the patients with autoimmune polyendocrinopathy-candidosis-ectodermal
dystrophy (APECED) syndrome had CMC, and 10% of the patients above the age of 25
developed oral or oesophageal carcinoma (Rautemaa et al. 2007a). Multiple studies
have shown that C. albicans isolated from oral cancer and APECED patients can
produce substantial levels of ACH (Marttila et al. 2013a, Marttila et al. 2013b, Uittamo
et al. 2009). High salivary ACH levels are associated with a high prevalence of Candida
yeasts, mainly C. albicans (Tillonen et al. 1999). Carcinogenic nitrosamine production
by Candida has been proposed as one possible mechanism underlying the malignant
transformation, and to support this, Hsia et al. reported an increased nitrosation
potential in strains isolated from pre-cancer lesions with more advanced dysplasia
(Hsia et al. 1981, Krogh 1990).
48
6
AIMS OF THE STUDY
The general aim of this study was to explore the ability of Candida yeasts to produce
carcinogenic acetaldehyde (ACH) during the planktonic and biofilm modes of growth,
and the effect of bacterial co-culture and metabolites on candidal growth, metabolism
and the host response.
The specific aims for this study were:
I
To determine the ability of Candida spp. to produce ACH from ethanol in the
planktonic mode of growth in vitro and examine the effect of glucose,
fructose and xylitol on ACH production solely or in combination with ethanol.
II
To examine the synergistic potential of bacteria and yeast to produce ACH
and ethanol during the production of fermented milk mursik. To study this
interplay, a simulated fermentation process was performed and the
microbiota of the mursik starter cultures used in the fermentation process
was identified. Starter cultures were collected from local families in western
Kenya, an area with an unusually high incidence of oesophageal cancer and
regular consumption of fermented food products.
III
To analyse the ability of C. albicans biofilms to produce ACH from ethanol
and glucose in vitro.
IV
To explore the potential of D,L-2-hydroxyisocaproic acid (HICA), a
Lactobacillus metabolite, to inhibit the growth and metabolism of C. albicans
biofilms in vitro as well as its impact on the degree of inflammation and
biofilm growth in vivo.
49
7
MATERIALS AND METHODS
7.1
STUDY DESIGN
The experiments performed in this study can be divided into two main parts, as shown
in Figures 6 and 7. In brief, the ability of planktonic Candida spp. to produce
acetaldehyde (ACH) was first investigated (I,II). Secondly, planktonic starter cultures
with a mixed bacteria-yeast flora were used to produce fermented milk. ACH and
ethanol levels were measured during the fermentation process and members of the
microbiota were identified (III). Thirdly, the effect of the Lactobacillus metabolite HICA
and its various comparators on the metabolic activity, biomass and ultrastructure of C.
albicans biofilms was explored using XTT and PicoGreen assays and scanning electron
microscopy (SEM), respectively, in vitro (IV). The production of ACH by biofilms was
measured using gas chromatography, and RT-qPCR was performed in order to analyse
the transcriptional changes in genes contributing to ethanol and ACH metabolism
during biofilm formation and HICA challenge (IV). Finally, the anti-biofilm and antiinflammatory activity of HICA was assessed in a C. albicans biofilm murine model in
vivo (V).
Figure 6.
I, II, III).
50
Study design for the experiments with planktonic microbes (RT = room temperature;
Figure 7.
7.2
7.2.1
Study design for the experiments with biofilms (IV,V).
MATERIALS
Candida strains and isolates
A total of three reference strains and one clinical isolate of C. albicans and total of
seven reference strains and 23 clinical isolates of non-albicans Candida (NAC) were
used for this study (Table 4). Reference strains were obtained from the American Type
Culture Collection (ATCC), the Culture Collection of the University of Gothenburg
(CCUG) and the United Kingdom National External Quality Assessment Service (UKNEQAS), except for C. albicans SC5314 and 3153A reference strains, which were
provided by Prof Gordon Ramage (University of Glasgow). Clinical isolates were
obtained from isolate depositories of the Clinical Microbiology Laboratory of Helsinki
University Central Hospital HUSLAB (T and HI- isolates) and the Institute of Dentistry,
University of Helsinki (D and G isolates). Candida spp. were identified based on colony
morphology on CHROMagar medium (CHROMagar, France) or API32C (Bio-merieux,
France) assimilation tests. All isolates and strains were stored in milk-glycerine at -70
o
C and plated at least twice on Sabouraud dextrose agar (SAB; Melford, UK or Lab M,
o
UK) and incubated under aerobic conditions at 37 C for 48 h before use to check
purity and viability.
51
Species
C. albicans
Isolate no.
ATCC 90029
SC5314
3153A
HI2580
C. dubliniensis UK NEQAS 2/07
G29
G130
G161
G174
C. glabrata
CCUG 32725
UK NEQAS 5/08
G212
G52
T1609
C. guilliermondii UK NEQAS 9/06
G66
G68
T1475
T1043
C. krusei
ATCC 6258
D206B
G65
T1336
T880
C. parapsilosis ATCC 22019
G95
G170
T1282
T1432
C. tropicalis
ATCC 750
D213
G9
G22
T1011
Table 4.
7.2.2
I
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
x
II
x
Study
III
IV
x
x
x
V
x
x
x
x
x
x
x
x
x
x
x
x
x
x
Summary of all Candida strains and clinical isolates used (n = 34).
Reagents
The different growth media, carbon substrates and antimicrobials used in the study
are listed in Table 5.
52
Table 5.
7.2.2.1
Different growth media, substrates and antimicrobials used in the experiments.
Starter cultures
A total of eight starter cultures were used to study ACH and ethanol production during
a simulated mursik milk fermentation process (III). Each starter culture was obtained
from separate families belonging to the Kalenjin tribe in the Bomet area, Rift Valley
Province, western Kenya by investigators from Tenwek Hospital (Bomet, Kenya).
7.2.2.2
Primers
All primers used in the study are listed in Table 6. Primers used for RT-qPCR
experiments were designed using Primer 3.0 Plus and were manufactured and
produced by Eurofins MWG Operon (Eurofins MWG Operon, Germany; Untergasser et
al. 2007). CaRIP1 was used as a reference gene for RT-qPCR experiments (Nailis et al.
2006). The PCR amplification efficiencies for all primers were optimised prior to the
analysis. ITS1 and NL4 primers spanning the ITS1-5.8S, rRNA-ITS2 and 26S rRNA
regions were used for the identification of yeast isolates and PA and PH primers
53
spanning the 16S rRNA were used for the identification of bacterial isolates (III;
Hutson et al. 1993, Kurtzman and Robnett 1997, White 1990).
Table 6.
7.2.2.3
Primers used in the experiments.
Antibodies
The antibodies used are listed in Table 7 (V).
Table 7.
54
Primary polyclonal antibodies used for immunohistochemistry (V).
7.3
METHODS
7.3.1
7.3.1.1
Candida growth conditions
Planktonic cells
o
Candida yeasts were incubated on SAB plates under aerobic conditions at 37 C for 48
h (I,II). One to two colonies were suspended in phosphate buffered saline (PBS, 6 M)
7
and the inoculum was adjusted spectrophotometrically to correspond to 1.0 x 10
colony forming units per millilitre (CFU/ml). The counts were checked by dilution
plating.
7.3.1.2
Biofilm formation
C. albicans reference strains 3153A, ATCC 90029 and SC5314 were used to study the
biofilm mode of growth. The ability of each strain to produce ACH and their
susceptibility to HICA and caspofungin were evaluated in the pilot experiment, and
the most representative strain was selected for the final experiments (Figure 7; IV,V).
One colony was suspended into 20 ml of yeast peptone dextrose (YPD) broth and
o
grown overnight at 37 C with continuous shaking. Cells were harvested and washed
twice with PBS, and were then re-suspended in RPMI-1640 medium (#6504) and
buffered to pH 7.4 with morpholinepropanesulfonic acid (MOPS) (Oxoid Ltd., UK). The
6
cell density was measured using a haemocytometer and adjusted to 1.0 x 10 CFU/ml
7
(study IV) or 1.0 x 10 CFU/ml (V) to produce the standardised biofilm inoculum. Cell
counts were verified by dilution plating.
7.3.1.2.1 In vitro biofilm
C. albicans biofilms were grown in vitro on Thermanox® coverslips (Thermoscientific,
USA) in RPMI #6504 medium at pH 7.4 in static 24-well plates (Corning Costar, USA)
o
under aerobic conditions at 37 C. After 4, 24 and 48 h of growth, the biofilms were
exposed to 5% (w/v) D,L-2-hydroxyisocaproic acid (HICA) and the comparators 10 mg/l
caspofungin, 5% (w/v) leucine, 0.05% (v/v; 11 mM) ethanol, PBS or RPMI-1640 #6504
for 24 h at pH 5.2 or pH 7.4. Substrates were dissolved in RPMI 1640 #6504, except for
the ethanol treatment, which was dissolved in RPMI #1383 without D-glucose or
sodium bicarbonate and then adjusted to pH 5.2 or pH 7.4.
55
7.3.1.2.2 Biofilm susceptibility to D,L-2-hydroxyisocaproic acid (HICA)
To test the susceptibility of C. albicans biofilm to HICA in vitro, a previously described
and standardised method was used (Pierce et al. 2008). Biofilms of three different
strains were pre-grown on 96-well plates in RPMI-1640 (#6504) at pH 7.4 under
o
aerobic conditions at 37 C. After 24 h of growth, biofilms were exposed to HICA by
injecting 100 µl of the two-fold dilution series (0.039–5%; v/v) into the well.
Caspofungin (10 mg/l) and RPMI were used as controls. The measurements were
performed in sextuplicate for each concentration. After treatment, the reduction in
metabolic activity was measured using the XTT assay, as described below.
7.3.1.2.3 In vivo biofilm
7
A volume of 100 µL of 10 CFU/ml C. albicans inoculum was injected percutaneously
6
into a diffusion chamber underneath the dorsal skin of a mouse (10 CFU/mouse; V).
The inoculum was well mixed by vortexing before injection. The counts were rechecked after injection by dilution plating on SAB. The viability of the biofilms formed
in vivo was checked by culture at the end of the experiment after treatment with
HICA, caspofungin or PBS.
7.3.2
Fermentation using starter cultures
The fermentation experiments simulated the protocol used by local population in
western Kenya to produce mursik milk (III). Finnish whole milk (Arla-Ingman, Finland)
was used in the experiments. Pilot fermentation was performed with non-boiled milk
and 0.1 ml of starter culture was mixed with 9.9 ml of the milk in a glass vial, sealed
o
and left to ferment for 0,1, 4, and 6 days at RT (22–25 C).
For the final experiment, the milk was first briefly boiled as suggested in the protocol
and then allowed to cool before mixing with the starter culture in similar proportions
to those used in the pilot experiment. The vials were sealed airtight and left to
ferment for 0, 24, 48 hours and 4, 7 and 14 days at RT. The pilot fermentation was
carried out once with three parallel samples. The final fermentation experiment was
carried out twice and the pH of the milk was measured at each time point.
56
7.3.3
7.3.3.1
Isolation and identification of the microbiota in starter
cultures
Assessment of microbial growth and isolation of microorganisms
Microbial growth (CFU/ml) was measured during fermentation at the time points 0,
24, 48 hours and 4, 7, 14 days by dilution plating on lysed blood agar (BA). Selective
and non-selective media under aerobic and anaerobic conditions were used to
determine the microbial flora in starter cultures. Fastidious anaerobe agar (FAA)
supplemented with 5% v/v horse blood was used to determine the total number of
cultivable bacteria. BA was used to enumerate the total number of aerobic bacteria.
Neomycin-vancomycin blood agar (NV; blood agar and neomycin sulphate)
supplemented with 7.5 µg/ml vancomycin, 0.5 µg/ml menadione and 5% v/v sheep
blood was used to determine the number of anaerobic Gram-negative bacteria.
Aerobic Gram-negative fermentative rods were isolated using cysteine-, lactose- and
electrolyte-deficient agar (CLED; C.L.E.D. medium). The growth of yeasts was
determined using SAB. The BA, CLED and SAB plates were incubated under aerobic
o
conditions and the FAA and NV plates under anaerobic conditions at 37 C for 48 h or
5 days, respectively. After incubation, multiple subcultures of bacteria and yeast
colonies were prepared. 16S or 18S PCR gene sequencing was used to detect bacterial
and fungal isolates from these pure cultures. Yeast isolates were also identified by API
32C auxanographic strips (bioMerieux, France) following the manufacturer’s protocol.
7.3.3.2
DNA extraction
Before gene sequencing, DNA was extracted from the bacterial and fungal pure
cultures (III). The MycXtra DNA extraction kit was used following the manufacturer’s
instructions to extract fungal DNA (Myconostica, UK). Bacterial DNA was initially
extracted using Chelex-100 beads (Walsh et al. 1991). If these two protocols were
unsuccessful, a modified cetyl trimethyl ammonium bromide (CTAB) method was
performed (Wilson 1987).
Briefly, HCl-washed beads (diameter 425-600 µm, Sigma G8772, Sigma-Aldrich, USA)
were mixed in a bead tube together with pre-heated extraction buffer containing 2%
CTAB and a few colonies from the pure culture. Suspensions were homogenized using
a Fastrep-24 instrument (MP Biomedical, USA). The vortex and heating steps were
repeated twice. Supernatants were collected into new tubes after centrifugation at
o
13 400 rpm for 2 min and followed by RNase treatment at 37 C for 30 min.
57
Chloroform-isoamyl alcohol was added, mixed well and centrifuged at 13 400 rpm for
2 min. The RNase and chloroform-isoamyl treatments were repeated once. The
aqueous phase was collected and mixed with isopropanol, followed by centrifugation
at 13 400 rpm for 2 min. The pellets were washed with ethanol and air-dried. Before
storage, the pellets were re-suspended in 50 µl of dH2O.
7.3.3.3
16S and 18S sequencing and data analysis
ITS1 and NL4 primers were used to identify yeast isolates from fungal DNA (Table 6;
III). PA and PH primers were used to identify bacterial isolates from bacterial DNA
(Table 6; III). Sequencing was conducted using a 16 capillary ABI Prism 3100 Genetic
Analyzer (Thermo Fisher Scientific, USA) and performed either at the University of
Manchester DNA Sequencing Facility or Beckman Coulter Genomics sequencing
service (Essex, UK). The sequences were aligned and matched to sequences stored in
GenBank using the NCBI-BLASTn matching tool (Altschul et al. 1990, Nilsson et al.
2006).
7.3.4
Gas chromatography
A Perkin Elmer Autosystem Gas Chromatograph and Headspace Sampler HS 40XL
(Perkin Elmer, USA) (I,II,III) or a Varian CP-3800 and CombiPal autosampler (Varian
Medical Systems, USA) (IV) were used to measure ACH and ethanol concentrations.
Chromatographs were equipped with Zebron WaxPlus or Supelco columns (CTC
Analytics GmbH, Germany or Sigma-Aldrich, USA). Peaks were detected by flame
ionisation.
7.3.4.1
Planktonic cells and biofilms
7
In studies using planktonic cultures of Candida spp., 350 µl (II) or 400 µl (I) of 1.0 x 10
CFU/ml yeast suspension was transferred into a gas chromatography vial. Thereafter,
50 µl of PBS buffer containing ethanol (I,II) and 50 µl of PBS containing D-glucose, Dfructose or xylitol (II) was added and the vial was immediately sealed airtight. The
ability of Candida to produce ACH solely from D-glucose, D-fructose or xylitol was
measured by adding 50 µl of one of these substrates and 50 µl of PBS into the vial
o
with yeast suspension (II). Vials were incubated at 37 C under aerobic conditions for
30 min. Metabolism was stopped by injecting 50 µl of 6 M PCA through the rubber
septum of the vial. To measure the baseline and artefactual ACH, 50 µl of PCA was
added to the vials immediately after adding the substrates and before incubation at
o
37 C.
58
The ability of three different C. albicans strains to produce ACH as biofilms was tested
and compared with planktonic cells in a pilot experiment. Biofilm was grown in RPMIo
1640 medium (#6504) for 24 h at 37 C, pH 7.4. The production of ACH by each strain
as planktonic cells was tested in parallel and the inoculum size was adjusted to
correspond the initial biofilm inoculum. In the pilot experiment, the biofilms and
planktonic cells were incubated with similar substrates to those used with planktonic
cultures above, except that fructose was not used. In the final biofilm experiment,
ACH production by pre-grown (4, 24, 48 h) biofilms exposed to 5% HICA, 11 mM
ethanol, 10 mg/l caspofungin or RPMI for 24 h was measured (IV).
In both pilot and final experiments with C. albicans biofilms, the biofilms were washed
with PBS before ACH analysis and transferred to 20 ml glass vials containing 1 ml of
ethanol, D-glucose or PBS. In the pilot experiment, the effect of xylitol on ACH
production from ethanol was tested by adding xylitol (1 ml) or ethanol (1 ml) to the
vial separately or in combination (500 µl + 500 µl). The vials were sealed with silicon
o
caps airtight and incubated under aerobic conditions at 37 C for 30 min. The reaction
was stopped by 100 µl of 6 M PCA. To measure the artefactual ACH production, 100 µl
of PCA was injected immediately after transferring the biofilm to the vials and prior to
adding the substrates. The final concentrations in incubations were 11 mM (I,IV) or 12
mM (II) for ethanol and 100 mM (I,IV) or 110 mM (II) for sugars.
7.3.4.2
Starter cultures and fermentation
Gas chromatography was performed at each time point of the pilot fermentation (0,
1, 4 or 6 days). In the final fermentation experiment, ethanol and ACH levels were
measured at the first three time points (0, 24 and 48 hours), which were selected
based on the microbial growth curves. For the ACH and ethanol measurements, 500 µl
aliquots of the suspension, including the milk and starter culture, were pipetted into
20 ml gas chromatography vials, closed airtight and left to ferment at RT. To measure
artefactual ACH and ethanol production, vials including the milk alone were
fermented and analysed in parallel to actual samples at each time point. The
fermentation process was stopped by injecting 50 µl of 6 M PCA through the rubber
septum of the vial. Artefactual concentrations of ACH and ethanol were subtracted as
background. In the final and pilot experiments, three independent and parallel
samples were used. The final experiments were performed twice.
To determine the microbial source for major ACH production during fermentation, a
second pilot experiment was performed in parallel to isolation of the microbial flora.
59
The starter culture with the highest ACH content measured in the pilot fermentation
was selected for the experiment. A volume of 0.1 ml of starter culture was added
onto non-selective (FAA or BA) or selective (SAB) growth media and incubated as
described above. After culturing, 3 ml of PBS was injected onto each growth media,
colonies were gently scrubbed into the liquid, and this was then transferred into a
sterile tube. The tube was homogenised by vortexing and 400 µl of this suspension
was used for ACH analysis. The suspension was incubated with 50 µl of 100 mM
o
glucose or 11 mM ethanol for 30 min at 37 C and the reaction stopped with 50 µl of
PCA. The baseline and artefactual ACH production was measured as earlier described
in the ACH experiments with planktonic cultures.
7.3.5
Analysis of Adh activity
A fluorometric approach was applied to study the Adh enzyme activity in Candida
yeasts as previously described (Kurkivuori et al. 2007). Planktonic cells were grown as
described earlier and then lysed with glass beads (diameter 1.0 mm) by vortexing in
5 x 1 min cycles in the presence of a protease inhibitor mixture (#P8340; SigmaAldrich, USA). Before each cycle, samples were cooled on ice. Cell lysates were
centrifuged for 5 min at 2,900 g (Hettich EBA 20, Hettich, UK). The supernatants were
o
collected and further centrifuged at 139,700 g for 65 min at 4 C (Beckman Optima Le80k Ultracentrifuge, Beckman Coulter, USA). The final supernatant was used for the
enzyme analysis. Cytosolic Adh activity was determined by measuring fluorescence
(excitation 340 nm, emission 440 nm) after the addition of ethanol or ethanol and
o
xylitol (final concentration 100 mM) and NAD (final concentration 2.5 mM) at 37 C in
0.1 M glycine buffer (pH 9.6). Ethanol was used at concentrations ranging from 0.68 to
2174 mM. Adh activity was detected using a Tecan Safire microplate detection system
and Magellan software ver. 6.05 (Tecan Trading AG, Switzerland). Lineweaver Burk
plots and GraphPad Prism software ver. 5.0 (GraphPad, USA) were used to determine
Vmax and Km values for the Adh enzyme.
7.3.6
7.3.6.1
Evaluation of Candida albicans biofilm formation
XTT reduction assay
The 2,3-bis-(2-methoxy-4-nitro-5-sulfophenyl)-2H-tetrazolium-5-carboxanilide (XTT)
reduction assay was performed to measure the metabolic activity of C. albicans
biofilms (IV). Before analysis, biofilms were washed thoroughly with sterile PBS.
Biofilms grown on coverslips were transferred to fresh 24-well plates after the wash
(Corning Costar, UK). Then, 100 or 200 µl of saturated XTT/1 µM menadione solution
was added to the 96- or 24-well plate. The plates were covered with aluminium foil
60
o
and incubated for 2 h at 37 C. After incubation, 100 µl of XTT supernatant was
pipetted into a fresh 96-well plate (Corning Costar, UK) and the colorimetric changes
were recorded spectrophotometrically (BMG Labtech, UK) at 492 nm.
7.3.6.2
PicoGreen assay
Fluorescent nucleic acid Quant-iT PicoGreen® reagent (Molecular Probes, USA) was
used to quantify double stranded DNA (dsDNA) in C. albicans biofilms (IV). The
amount of dsDNA has been shown to correlate with the number of viable cells, thus
describing the biomass of a biofilm (Marstorp and Witter 1999). DNA was extracted
from the biofilms using a modified CTAB method as described above. Extracted DNA
and PicoGreen® reagent were thoroughly mixed in the well and analysed
fluorometrically at 492 nm (BMG Labtech, UK). The lambda DNA within the Quant-iT
kit was used for the construction of the standard curve (concentration range 40–500
ng/ml) according to the manufacturer’s instructions and pipetted alongside samples
(100 µl per well; Corning Costar, UK).
7.3.6.3
Scanning electron microscopy (SEM)
Biofilms were pre-grown in RPMI as previously described for 24 h and then exposed to
5% HICA or RPMI (control) for 24h at pH 5.2 and pH 7.4 (IV). The biofilms were washed
once with PBS and then placed in a fixative overnight as previously described
(Erlandsen et al. 2004). After fixation, biofilms were rinsed in 0.1 M phosphate buffer
and air-dried. The samples were coated with gold/palladium (40/60% ratio) and
observed under a scanning electron microscope (Leo 435 VP) in high vacuum mode at
10 kV.
7.3.7
7.3.7.1
Gene expression analysis
RNA extraction
In vitro biofilms used for the gene expression analysis were flash-frozen in liquid
o
nitrogen post-exposure to HICA and comparators and stored at -80 C until RNA
extraction (IV). Biofilms were first detached by vortexing and then homogenized using
glass beads (425–600 µm in diameter; Sigma-Aldrich, USA) and a FastPrep-24
Instrument (MP Biomedical, USA). The samples were centrifuged at RT for 1 min at
12,000 g after the homogenization step. RNA was extracted using the Isolate I RNA
Mini kit according to the manufacturer’s instructions and genomic DNA was removed
using Ambion DNA-free DNase (Life Technologies, UK). The amount and quality of RNA
61
was assessed spectrophotometrically (A260). The purity was further assessed by RTo
qPCR and gel electrophoresis. RNA was stored at -80 C for further use.
7.3.7.2
cDNA synthesis
cDNA synthesis by reverse transcription was performed using an AffinityScript qPCR
cDNA synthesis kit (Agilent, UK) according to the manufacturer’s protocol (IV). Six
microlitres of RNA (10 ng/µl) was used as a template. In no-RT controls, reverse
transcriptase was replaced by dH2O.
7.3.7.3
RT-qPCR analysis
RT-qPCR reactions were conducted using a Brilliant II SYBR Green 2-step kit (Agilent,
UK) and as described in the manual (IV). The reaction mastermix was composed of
6.25 µl of SYBR Green fluorescent probe, 1 µl of forward (F) primer, 1 µl of reverse
primer, 2 µl of cDNA and 2.25 µl DNA/RNA-free molecular grade water (primers listed
in Table 6). RT-qPCR was performed with a Stratagene MX3005P instrument (Agilent,
UK).
7.3.8
7.3.8.1
In vivo biofilm murine model
Ethics statement
All mice in the study were handled in strict accordance with good animal practice and
conducted under an ethically reviewed license authorised by the Secretary of the
State at the University of Manchester, UK (license no. PPL 40/3101) (V).
7.3.8.2
Diffusion chamber model
A previously published chamber model was modified for the in vivo biofilm
experiment (Rajendran et al. 2011) (V). The biofilm diffusion chamber was structurally
based on a chamber kit (Millipore, Watford, UK). The chamber was made of a
semipermeable Durapore® membrane with a pore size of 0.45 µm fixed to a Plexiglas®
ring. A non-permeable silicon sheet was fixed onto the opposite side to close the
chamber. A total of 24 male CD1 mice weighing 21–24 g were included. Mice were
allowed to acclimatise before surgery for seven days. The surgical implantation of the
chambers was performed under isoflurane and xylocaine anaesthesia. The dorsal flank
of each mouse was shaved and a 2-cm incision was made. The diffusion chamber was
implanted subcutaneously, so that the semipermeable membrane faced the dorsal
62
flank of each mouse shaved and the non-permeable silicon sheet faced the skin. The
wound was closed with non-absorbable braided silk sutures (Ethicon, USA) and
meloxicam (3 mg/kg) was administered intraperitoneally daily for 3 days post-surgery.
One mouse was lost in surgery (n = 23). Mice were divided into two main groups:
biofilm (n = 15) and no-biofilm (n = 8). The chambers in the biofilm group were
6
injected percutaneously with C. albicans inoculum under isoflurane anaesthesia (10
CFU/mouse). The mice were allowed to recover for 5 days. A robust C. albicans biofilm
formed inside the chamber during the infection period. After five days of infection,
biofilms were treated for 12 h with 5% HICA (n = 8), and 10 mg/l caspofungin (n = 3) or
PBS (n = 4) were used as control treatments. No-biofilm chambers were similarly
treated with HICA (n = 2), caspofungin (n = 3) or PBS (n = 3). The mice were
euthanised post-treatment with an overdose of isoflurane. Chambers were collected
and biofilms detached and weighed. Tissues around the chambers were dissected and
embedded into 10% formaldehyde for storage and further analysis.
7.3.9
Immunohistochemistry
A previously described method was adapted for the immunohistochemistry and semiquantitative analyses (Kuula et al. 2009) (V). Tissue sections from mice were
embedded in paraffin. Paraffin-embedded specimens were cut, deparaffinised,
pretreated with 0.4% pepsin, and the endogenous peroxidase activity was blocked in
H2O2/methanol solution. Staining was performed with polyclonal Vectastain Elite
rabbit or goat ABC kits (Vector Laboratories, Burlingame, CA, USA). Samples were
blocked with goat or rabbit normal serum in 2% bovine serum albumin and incubated
with the following polyclonal primary antibodies: MMP-8, MMP-9, MPO, NE, IL1β,
TNFα and Del-1 (Table 7). The control sections were incubated with non-immune
rabbit or goat serum. The studied inflammatory markers were visualized with a
biotinylated anti-rabbit or anti-goat secondary antibody and avidin-biotin enzyme
complex. 3-Amino-9-ethyl-carbazole was used as a chromogen and Mayer’s
hematoxylin (Histolab Products AB, Sweden) as a counterstain. All sections were also
stained with hematoxylin and eosin (H&E) for routine histopathology. Slides were
mounted with Dako glycergel (Dako corp., USA).
The immunohistochemical staining was evaluated under an Olympus BX61 light
microscope and representative images were taken using an Olympus DP50 camera
attached to the microscope and AnalySIS software (AnalySIS ver. 3.2, Soft Imaging
System GmbH, Germany). The intensity of staining was semiquantified and graded as
no staining, mild, moderate or strong staining. The semiquantitative analyses were reconfirmed blindly by a second evaluator and a trained pathologist examined the basic
histopathology without information on the groupings. Distributions of staining
63
intensities within groups were visualised using GraphPad Prism ver. 5.0 software
(GraphPad Software Inc., USA).
7.3.10 Statistical analysis
PASW statistics ver. 15.0 - 21.0 (IBM, USA) and Graphpad Prism software ver. 5.0
(GraphPad Software Inc., USA) were used for statistical analyses in all studies.
Univariate analysis of variance (ANOVA) or a generalised estimating equations (GEE)
model was used for statistical comparisons (I, II, IV and V). Relative gene expressions
were statistically compared with a pair-wise fixed re-allocation test within REST 2009
ver. 1.0 software (Qiagen, USA) (Pfaffl et al. 2002) and basal transcriptions of each
-ΔCt
gene relative to the reference gene RIP1 were evaluated using equation 2
(ΔCt=Cttreatment-Ctreference), and these levels were used to calculate correlations (IV).
Spearman’s rho (rs) with a 95% confidence interval was used to calculate correlations
(I, II, III and IV). A p-value less than 0.05 was considered statistically significant.
64
8
RESULTS
8.1
PRODUCTION OF ACETALDEHYDE AND ETHANOL BY CANDIDA
YEASTS IN VITRO
This section summarises the results from all in vitro experiments, in which the ability
of Candida yeasts to produce acetaldehyde (ACH) as planktonic mono- or mixed
cultures or as biofilms was investigated (I, II, III, IV). ACH production is described as
high (>100 µM), moderate (40–100 µM) or low (<40 µM). All measurements with
planktonic monocultures and biofilms were performed after 30 min incubation with
o
substrate at 37 C. The ability of starter culture with a mixed planktonic bacteria-yeast
flora to produce ACH and ethanol was studied in a simulated fermentation
experiment that was carried out at room temperature (RT).
8.1.1
8.1.1.1
Planktonic Candida albicans and non-albicans Candida
cultures
Acetaldehyde production in ethanol incubation
All Candida isolates produced moderate or high levels of ACH from ethanol in vitro.
Marked variation was observed between and within species (Figure 8; I, II). C. albicans
and three NAC species, including C. parapsilosis, C. tropicalis and C. glabrata,
produced over 200 µM of ACH from ethanol. C. parapsilosis was the highest producer
(258.9 ± 48.6 µM). ACH production by C. guilliermondii and C. dubliniensis was lower,
but high levels were still detected (177.2 ± 62.9 µM and 165.0 ± 53.1 µM,
respectively). Moderate levels were only measured for C. krusei (56.0 ± 12.0 µM).
65
Figure 8. Acetaldehyde production by planktonic Candida spp. when incubated for 30 min at 37
C with ethanol, ethanol + xylitol, ethanol + glucose or glucose only, and ethanol + fructose or
fructose only. The strains and isolates used are presented in the Material and Methods section
(Table 4). Means (± SD) are shown.
o
8.1.1.2
Acetaldehyde production in glucose, fructose or xylitol
incubation or in co-incubation with ethanol
Major differences were observed in the ability of C. albicans and NAC species to
produce ACH when incubated with glucose or fructose (Figure 8; I,II). None of the
isolates were able to produce high levels. Moderate levels (40–100 µM) were mostly
produced by C. glabrata: 89.9 ± 18.8 µM in glucose and 94.2 ± 8.9 µM in fructose
incubation. One of the two C. albicans isolates produced moderate levels from
glucose. Almost no ACH was detected when Candida isolates were incubated with
xylitol alone (<1.8 µM). When glucose or fructose was co-incubated with ethanol, C.
glabrata was the only species that showed increased ACH production compared to
incubation with ethanol only (375.4 ± 60.7 µM vs. 220.5 ± 79.7 µM, p < 0.001). For
other species, glucose-ethanol or fructose-ethanol co-incubation resulted in a
decrease in ACH production when compared to incubation with ethanol alone (mean
23% and 29%, respectively). This change was not significant (p = ns). However, coincubation of ethanol and xylitol significantly reduced ACH production by a mean of
84% compared to incubation with ethanol alone in all species (p < 0.001).
66
8.1.1.3
Adh activity
Enzyme kinetics were investigated for Adh enzymes of three different Candida
species, which were chosen to represent the highest (C. glabrata CCUG 32725),
moderate (C. albicans ATCC 90029) and lowest (C. krusei ATCC 6258) ACH producers
-1
(II). C. glabrata Adh showed the highest activity (Vmax = 4.5 s , Km = 2.8 mM). This was
-1
followed by C. albicans Adh (Vmax = 3.4 s , Km = 0.2 mM). Adh with low activity was
-1
recorded for C. krusei (Vmax = 1.7 s , Km = 0.1 mM). Xylitol (100 mM) reduced the
enzyme activity in C. glabrata by 61–66% when co-incubated with 11 or 110 mM
ethanol. In C. albicans, the reduction in activity was 100%. No reduction was observed
in C. krusei.
8.1.2
Mixed bacteria/yeast starter culture during fermentation
The potential of a mixed bacteria/yeast culture to produce ACH was examined in a
simulated fermentation using mursik starter cultures. The first two pilot experiments
were carried out to test the potential ability of each starter culture (n = 8) and the
selected members of the starter culture microbiota to produce ACH and ethanol.
Starter cultures were mixed with whole milk and left to ferment at room temperature.
The microbiota of the starter cultures was identified. The changes in pH, microbial
counts, ACH and ethanol content were recorded during the final fermentation
process.
8.1.2.1
Pilot experiments
Figure 9. Results from two pilot experiments. (A) Each milk starter culture (n = 8) was mixed
with whole milk and left to ferment at room temperature for 6 d. The acetaldehyde and ethanol
content was measured at four time points during the fermentation. (B) Acetaldehyde production
by an anaerobic and aerobic bacterial flora and yeast. Selective and non-selective media were
used for isolation, as described in the Materials and Methods section. The colonies were scraped
off the media into the liquid and incubated with ethanol or glucose for 30 min at 37 oC. Means (±
SD) are shown.
67
The pilot fermentation demonstrated that significant ACH and ethanol production
occurs during fermentation at RT (Figure 9A). However, variation was observed in the
ability of each mursik starter culture to produce ethanol or ACH. The highest mean
ACH and ethanol levels were reached at 6 d (324.6 ± 295.3 µM and 47.5 ± 12.6 mM,
respectively). A second pilot was carried out in parallel to the identification of the
microbial flora of each starter culture. Microbial colonies on the selective and nonselective media belonging to the starter culture with highest ACH and ethanol levels in
the previous experiment were tested for their ability to produce ACH. Moderate or
high production of ACH was only detected when plates selective for yeasts or bacterial
aerobic flora were incubated with ethanol or glucose (Figure 9B). Yeasts produced the
highest levels of ACH in both ethanol and glucose incubations (130.4 ± 47.8 µM and
115.8 ± 15.6 µM, respectively).
8.1.2.2
Distribution of yeast and bacterial species in starter cultures
Figure 10. Microbial distribution and number of species in each starter culture (n = 8).
Yeasts were isolated from each starter culture. Their proportion of the total microbial
count (CFU/ml) varied considerably (Figure 10). Candida yeasts were isolated from
seven of the total of eight starter cultures, C. krusei being the most common (5/8
starter cultures). Other Candida species included C. kefyr and C. sphaerica.
Saccharomyces yeast was identified in one of the eight starter cultures. The bacterial
68
flora was dominated by Lactobacillus species. L. kefiri was the most frequently
isolated species (6/8 starter cultures). Furthermore, L. kefiri was commonly (5/8 milks)
found in combination with C. krusei. Other bacterial species were identified, but their
prevalence was lower.
8.1.2.3
Changes in microbial counts and pH during fermentation
An exponential increase was detected in microbial counts during the fermentation.
The counts first increased from the initial mean of 6.6 log10 CFU/ml to the peak value
of 10.4 log10 CFU/ml at 48 h, and then decreased slightly to 9.5 log10 CFU/ml at 14 d.
During the 14 d fermentation, the pH declined in a linear fashion from 6.6 to 3.5. No
significant variation in pH or microbial counts was detected between starter cultures.
8.1.2.4
Acetaldehyde and ethanol production during fermentation
and correlation with microbiota
ACH levels ranging from zero to 1809 µM were measured during fermentation. In
general, ACH and ethanol levels were higher and increased quicker compared to
pilots, and this may be due to the use of boiled whole milk as specified in the protocol.
Seven of the eight starter cultures produced high levels during first 24 h of
fermentation. At the 48 h time point, four milks reached an ACH content of over 1000
µM, with a range of 1150–1809 µM. High ethanol levels (>20 mM) were mostly
detected at 48 h. The highest ethanol levels were produced by the same four starter
cultures that also showed high ACH production (mean 79 mM, range 62–106 mM at
48 h). There was a significant correlation between ACH and ethanol production
(rs = 0.88, p < 0.001). The microbial composition correlated with high ACH production.
The four milks that produced high ACH levels also had similar combinations of yeast
and lactobacilli, mainly C. krusei and L. kefiri (milk no. 1–4; Figure 10). The prevalence
of other bacterial species correlated with low production of ACH.
8.1.3
8.1.3.1
Candida albicans biofilms
Pilot experiments
All three strains produced significantly higher levels of ACH from ethanol as a biofilm
(mean range 113.4–261.4 µM) compared to planktonic cultures (mean range 50.5–
71.3 µM), as demonstrated in Figure 11. ACH production by biofilms in ethanol
incubation was strain-dependent. Strain 3153A produced the highest levels of ACH
69
(261.4 ± 10.5 µM) and ATCC 90029 the lowest (113.4 ± 4.9 µM). However, SC5314 was
the only strain to produce moderate or high levels of ACH in both ethanol and glucose
incubations as a biofilm (164.1 ± 4.0 µM and 61.7 ± 2.8 µM, respectively). Xylitol coincubation with ethanol decreased the levels of ACH compared to incubation with
ethanol alone, and the reduction was significant for biofilms (p = 0.022; Figure 11).
Based on these results, strain SC5314 was chosen for further analysis (IV,V).
Figure 11. The ability of three different C. albicans strains to produce acetaldehyde from 11 mM
ethanol and 100 mM glucose was investigated in a pilot experiment. Cells were pre-grown in YPD
and counts were standardised (106 CFU/ml). Biofilms were allowed to form for 24 h in RPMI. The
effect of xylitol on ethanol oxidation was measured in co-incubation with ethanol, and
statistically significant differences are presented compared to incubation in ethanol alone (*
p < 0.05). Means (± SD) are shown.
8.1.3.2
Acetaldehyde production from ethanol and glucose at an
acidic and a neutral pH
All C. albicans SC5314 biofilms produced moderate or high levels of ACH in ethanol
incubation at both test pH values (Figure 12; mean range 49–135 µM at pH 5.2 and
40–130 µM at pH 7.4). The highest levels of ACH in ethanol incubation were produced
by biofilms grown in RPMI at both pH values: at pH 5.2 the peak ACH values were
detected for biofilms pre-grown for 24 h and at pH 7.4 for biofilms pre-grown for 48 h
(mean ± SEM: 135.2 ± 5.7 µM and 130.3 ± 12.3 µM). Biofilms that were pre-grown for
24 h and exposed to ethanol showed similarly high levels at the acidic pH (120.0 ±
12.7 µM).
When biofilms were incubated with glucose, ACH levels were markedly lower
compared to ethanol incubation (Figure 12). Biofilms produced more ACH at the
neutral pH compared to the acidic pH (range: mean 0–54 µM at pH 7.4 and mean 0–
41 µM at pH 5.2). Moderate levels were only produced by control biofilms pre-grown
for 48 h at neutral pH (53.8 ± 1.9 µM); otherwise, levels remained low.
70
Figure 12. Acetaldehyde production from (A) ethanol or (B) glucose by C. albicans SC5314
biofilms pre-grown for 4, 24 or 48 h in RPMI and then exposed for 24 h to HICA, ethanol and
caspofungin or RPMI (control) at pH 5.2 or 7.4. Means (± SEM) are shown.
8.2
8.2.1
INHIBITORY EFFECT OF D,L-2-HYDROXYISOCAPROIC ACID (HICA)
ON CANDIDA ALBICANS BIOFILM FORMATION AND
ACETALDEHYDE METABOLISM IN VITRO
Pilot experiment - HICA susceptibility
The susceptibility of C. albicans biofilms to HICA was tested using a concentration
range of 0.039–5% (w/v) and an antifungal control (caspofungin 10 mg/l). Three
reference strains were tested and HICA showed efficacy against all three strains
(Figure 13). No metabolic activity was detected for any of the three strains after
exposure to 5% (w/v) HICA. A major decrease in metabolic activity occurred when the
concentration was increased from 0.625% to 1.25%, and only 6–23% of the metabolic
activity remained compared to the control at the concentration of 1.25%. There was
considerable variation between strains in susceptibility to HICA and the caspofungin
control. Caspofungin showed no efficacy against the strain SC5314; moreover,
paradoxical growth was observed. In contrast, only 34% of metabolic activity
remained for 3153A strain after exposure to caspofungin. In addition, no reduction in
metabolic activity was shown for SC5314 in the concentration range of 0.039–
0.3125%, but 3153A retained only 56–65% of its metabolic activity in a similar range.
71
Figure 13. Percentage metabolic activity of mature (24 h) C. albicans biofilms formed by three
different strains (3153A, ATCC 90029 and SC5314) after 24 h HICA challenge relative to a control
biofilm in RPMI. To test the susceptibility, a concentration range of 0.039–5% (w/v) was used
(HICA diluted to RPMI medium at pH 5.2). Caspofungin, RPMI only and biofilm+RPMI controls were
included. Means (± SEM) are shown.
8.2.2
Effect of HICA on metabolic activity, biomass and
ultrastructure of biofilms
A major decrease in both metabolic activity and biomass was measured in all pregrown (4, 24, 48 h) biofilms after 24 h exposure to 5% HICA compared to RPMI at an
acidic pH (Figure 14A; mean -71% and -62%, respectively). At a neutral pH, the
efficacy of HICA was less significant: metabolic activity dropped by 22% and biomass
by 17% on average in all pre-grown biofilms. Caspofungin exposure only led to
significant changes in metabolic activity, and at a neutral pH the highest average
reduction of 29% of all conditions was measured. Interestingly, a reduction of 73% by
caspofungin was measured in 4 h pre-grown biofilms. The greatest reductions in
biomass at a neutral and acidic pH were observed post-exposure to cysteine (63% and
33%, respectively). Leucine, a metabolic comparator for HICA, increased the metabolic
activity by a mean of 3% and biomass by a mean of 14% at a neutral pH.
72
Figure 14. (A) Mean (± SEM) % metabolic activity and % biomass of all C. albicans pre-grown (4,
24 or 48 h) biofilms after 24 h exposure to 5% HICA and comparators relative to the control
(RPMI). (B) Scanning electron microscopy (SEM) images taken of 24 h pre-grown biofilms, the
upper biofilm treated with HICA and the lower one a control biofilm (RPMI) at pH 5.2.
Scanning electron microscopy (SEM) further supported the results concerning the
anti-biofilm activity of HICA. An SEM image revealed a collapsed network of hyphae
and aberrant cell wall structures in a 24 h pre-grown biofilm after HICA exposure at an
acidic pH (Figure 14B). At a neutral pH, an HICA-exposed biofilm showed no collapsing
hyphae, but a less dense mesh of hyphae and similar cell wall defects were observed.
A strong and intact mesh of hyphae was seen in images of RPMI controls at both pH
values.
8.2.3
Acetaldehyde production during challenge with HICA and
caspofungin
No ACH was detected in glucose incubation when biofilms were exposed to HICA at an
acidic pH (Figure 12). In addition, no production of ACH in glucose incubation was
seen with 4 h pre-grown biofilms exposed to caspofungin at a neutral pH. ACH
production by 24 and 48 h pre-grown biofilms exposed to HICA or caspofungin was
significantly lower compared to RPMI in glucose and ethanol incubation at a neutral
pH (p < 0.05). HICA exposure also resulted in lower ACH production in 4 and 24 h pregrown biofilms at an acidic pH. A similar significant reduction was observed in
caspofungin-treated biofilms at an acidic pH (p < 0.05). Interestingly, at an acidic pH,
moderate and the highest levels (>40 µM) were measured for 48 h pre-grown biofilms
exposed to caspofungin (41.4 ± 7.5 µM).
73
8.2.4
8.2.4.1
Transcriptional changes (CaADH1, CaADH2, CaALD4, CaALD5,
CaALD6, CaACS1 and CaACS2) in biofilms
Basal and relative gene expression
Transcriptional changes in pre-grown biofilms challenged with HICA or comparators
were first evaluated by analysing the basal expression levels at an acidic and a neutral
pH. Significant differences were detected in the expression of CaADH1, CaADH2 and
CaALD5 relative to the reference gene CaRIP1 (>1.0; Figure 15). ADH1 was most
highly expressed in 4 h and 48 h pre-grown biofilms at an acidic pH (range 2–18 fold).
The highest transcription rates for CaADH1 were detected after exposure to HICA,
cysteine or PBS. High CaADH1 expression was accompanied by CaADH2 expression
and a significant correlation was found (rs = 0.703, p < 0.001). However, CaADH2
expression always remained lower than CaADH1 expression (range 2–5 fold). No
similar up-regulation of CaADH1 or CaADH2 was recorded at a neutral pH. CaALD5
was the most highly up-regulated of all genes of interest in all conditions at both an
acidic and a neutral pH (range 2–24 fold). Interestingly, similar expression of CaADH1,
CaADH2 and CaALD5 was detected at an acidic pH in 48 h pre-grown biofilms after
exposure to HICA or cysteine. CaALD5 was the only gene up-regulated after exposure
to caspofungin (range 2–6 fold). This was mainly seen in 24 h pre-grown biofilms.
Figure 15. Basal gene expression of CaADH1, CaADH2 and CaALD5 in pre-grown (4, 24 or 48 h)
biofilms at pH 5.2 or 7.4 after 24 h exposure to HICA and comparators: caspofungin (CAS),
leucine (LEU), cysteine (CYS), ethanol (ETOH), and PBS and RPMI controls. The mean
transcription rate of each gene was calculated relative to the reference gene CaRIP1 under each
condition. The dotted line represents the transcription rate of 1.0 of the reference gene.
74
The expression levels were further standardised against the control condition (RPMI)
to examine the relative gene expression. Similar expression or significant upregulation of genes was observed after exposure to HICA at both pH levels relative to
RPMI, whereas caspofungin exposure mainly resulted in down-regulation of ethanol
metabolism. Interestingly, cysteine and leucine expression profiles showed similarities
with HICA. Ethanol exposure led to a similar expression profile to HICA at a neutral pH.
8.2.4.2
Correlation with acetaldehyde levels
In general, significant correlations were found between gene expression and ACH
levels. However, this was dependent on pH. At a neutral pH, CaALD5 expression
correlated negatively with ACH levels in ethanol incubation in all conditions tested (0.769 < rs < -0.661, 0.001< p < 0.003), but in glucose incubation this correlation was
only significant in the control condition (RPMI; rs = -0.763, p < 0.001). In contrast to a
neutral pH, no similar correlations were found at an acidic pH. The only significant
correlations were calculated under the control condition. ACH values correlated
negatively with CaADH1 and CaADH2 in ethanol incubation (rs = -0.558, p = 0.016 and
rs = -0.636, p = 0.005 respectively), but a positive correlation was found between
CaALD5 expression and ACH values (rs = 0.746, p < 0.001).
8.3
ANTI-BIOFILM AND ANTI-INFLAMMATORY PROPERTIES OF
HICA IN VIVO
HICA showed immunomodulatory properties when its efficacy was investigated in a
murine C. albicans biofilm infection model. No significant decrease in biofilm weight
was measured in vivo after 12 h treatment with HICA or the controls, caspofungin and
PBS (p = ns). A suppression of inflammation was observed in tissues surrounding
biofilm chambers treated with HICA compared to caspofungin or PBS controls. A
dense inflammatory infiltrate, including macrophages and PMNs, and a thick band of
granulation tissue were observed in the caspofungin and PBS biofilm groups. In
contrast, tissues in the HICA biofilm group showed characteristics more similar to noninfected, non-biofilm controls, as the infiltrate was predominantly of a mononuclear
cell type. However, in the HICA biofilm group, some PMN cells were also present.
Staining of MPO, MMP-8 and MMP-9 predominantly localised in inflammatory cells,
and the HICA biofilm group showed less intense staining compared to caspofungin and
PBS controls. When treatments were compared within biofilm groups, the most
prominent differences were observed in MPO and MMP-9 staining. In contrast, no
major differences were seen in the staining of MMP-8, MMP-9 or MPO in the nonbiofilm group. More intense staining for NE and IL1β was seen in inflammatory cells in
75
the biofilm group than in the non-biofilm group, with no differences between
treatments. Staining for TNFα was weak in all groups and treatments. Similar strong
staining for Del-1, produced by endothelial cells, was detected in both HICA biofilm
and all non-biofilm groups, whereas caspofungin and particularly PBS showed a milder
intensity.
76
9
DISCUSSION
When opportunity prevails, Candida yeasts can colonise the upper digestive tract in
high numbers and cause infection. Pathogenic Candida species prefer a biofilm
lifestyle in mucosal pathogenesis, which has been associated with oro-oesophageal
malignancies. The major risk factors for oral and oesophageal cancer are alcohol
abuse and smoking, and their effect on the risk is synergistic. When consumed, they
both result in an increased salivary acetaldehyde (ACH) content. The oral and
oesophageal microbiome has been shown as the main determinant for the local
production of ACH, the carcinogenic metabolite of ethanol (Homann et al. 1997).
Mucosal cells also possess the capacity to oxidise ethanol. However, their ability to
eliminate ACH is limited (Chiang et al. 2012). ACH-related mutagenesis has been
shown to occur at concentrations as low as 40 µM. Previous studies have
demonstrated that prevalence of yeast, particularly Candida yeasts, is associated with
high salivary ACH levels (Homann et al. 1997, Tillonen et al. 1999, Yokoyama et al.
2007).
9.1
PRINCIPLE FINDINGS
The results of this study provide strong evidence for the significant capacity and role
of Candida yeasts in the formation of carcinogenic ACH. This is the first study to show
that C. albicans biofilms are able to produce mutagenic amounts of ACH, as levels well
above 40 µM were measured after incubation with clinically relevant concentrations
of ethanol and glucose in vitro. The highest ACH concentrations were produced by 24
and 48 h pre-grown biofilms, which can be considered mature according to the
literature (Andes et al. 2004, Kaneko et al. 2013, Ricicová et al. 2010). This is
important, as mucosal biofilms are predominantly in a mature state. However, it is
relevant to note that mucosal biofilms are dynamic structures, and different stages
from initiation to dispersal can be observed in these biofilms (Uppuluri et al. 2010a,
Uppuluri et al. 2010b). Mutagenic levels of ACH were produced by biofilms at both a
neutral and an acidic pH. This is clinically relevant, as in the oral cavity and
oesophagus the pH can fluctuate daily in healthy subjects, and a more permanent
decrease in pH can occur in conditions such as reflux disease.
Furthermore, planktonic C. albicans cells were also able to produce mutagenic levels
of ACH from ethanol and glucose in vitro. In ethanol incubation, levels of up to 240
µM were measured. The mutagenic production of ACH by planktonic C. albicans
cultures has additionally been reported by other researchers (Tillonen et al. 1999,
Uittamo et al. 2009). When initial growth conditions and cell counts were
77
standardised in the pilot experiment of this study, biofilms produced significantly
higher levels of ACH compared to planktonic cells, again highlighting the impact of the
biofilm mode of growth on the production of ACH and providing insights into the
metabolic traits common for biofilms. All NAC species investigated were also able to
produce mutagenic amounts of ACH from ethanol in planktonic mode of growth in
vitro. When the ability to produce ACH was compared, large variation existed
between and within different Candida species. This can be explained by metabolic
rewiring due to their unique development through micro- and macroevolution in
various environmental niches (Butler et al. 2009, Ihmels et al. 2005). Interestingly,
only C. glabrata and C. albicans showed mutagenic production of ACH in glucose
incubation when grown planktonically.
Furthermore, C. glabrata was the only yeast that also produced ACH from fructose,
and co-incubation of ethanol with either of the two carbohydrates resulted in even
higher ACH production compared to incubation in ethanol alone. This suggests that
there is a high metabolic flux towards pyruvate bypass and ACH, although cells were
grown in aerobic conditions. Adh enzyme activity was also highest in C. glabrata,
which supports its high fermentative trait. These findings are supported by the
characteristics of C. glabrata, which is a Crabtree -positive yeast and more closely
located in the phylogenetic tree to another Crabtree -positive yeast, S. cerevisiae, than
other NAC species (Van Urk et al. 1990). High homology between the Adh1p enzyme
of C. glabrata and S. cerevisiae further strengthens the metabolic similarity (87%;
Inglis et al. 2012). For all Candida species, a significant reduction in ACH production
was detected when ethanol was co-incubated with sugar alcohol xylitol in vitro. Xylitol
is famous for its anti-cariogenic effect and inhibitory role in bacterial metabolism
(Persson et al. 1993). Adh and xylitol dehydrogenase (Xdh) enzymes show a close
resemblance in their structure, and both enzymes negatively affect to the redox
balance in the cell (Kotter et al. 1990, Persson et al. 1993). These factors could
contribute to the inhibitory action seen.
Carcinogenic exposure in the upper digestive tract may also occur by direct exposure
to ACH. Microbially fermented foodstuffs and beverages have been linked to the
aetiology of squamous cell carcinoma in the upper digestive tract (Lachenmeier and
Sohnius 2008, Linderborg et al. 2008). In this study, ACH and ethanol production
during fermentation was studied using original mursik milk starter cultures collected
from local families belonging to Kalenjin tribe in western Kenya. ESCC is the most
common malignancy among men and women, especially in members of Kalenjin tribe.
The common risk factors, smoking and alcohol consumption, are not highly prevalent
in the area (Duron et al. 2013, Parker et al. 2010, Wakhisi et al. 2005). A major portion
78
of the daily diet in Africans is comprised of fermented food products, and fermented
milk mursik is used daily by locals in western Kenya (Sanni 1993).
High levels of ACH (up to 1800 µM) and ethanol (up to 110 mM) were produced
during a simulated mursik fermentation process using original local starter cultures
and boiled whole milk. ACH is an important flavour compound in milk products, and
equally high levels of ACH and ethanol have been measured in previous studies
(Guzel-Seydim et al. 2000, Kok-Tas et al. 2013). The ACH and ethanol concentrations
correlated with the microbiota of the starter cultures. Candida yeasts and
Lactobacillus bacteria were the most prevalent microbial members isolated. These
microbial findings are in line with reports documenting the microbiota of industrial
and other indigenous fermentation processes (Narvhus and Gadaga 2003). When the
starter culture was grown on different selective media in a pilot experiment, colonies
from yeast-specific media were mainly responsible for the ACH production, again
highlighting the role of yeast as a source of significant ACH production. The highest
levels of ACH (1150–1809 µM) were produced by milks containing a C. krusei–L.kefiri
combination. When C. krusei was studied in planktonic mono-species cultures, it was
the lowest producer of ACH of all Candida species. The reduction in pH from mild to
highly acidic suggested robust lactate production by Lactobacillus. This is a wellknown property of Lactobacillus fermentations and beneficial for both yeasts and
Lactobacillus, as the majority of the competing or contaminating bacteria cannot
tolerate a low pH (Gadaga et al. 2001a, b, Mugula et al. 2003, Narvhus and Gadaga
2003). Thus, these findings imply that a metabolic synergy and symbiotic growth exist
between the bacteria and yeast. This can also be considered interesting in the context
of the human microbiome, where both species exist in combination and interact.
Competitive inhibition has been shown within the human microbiome in health and
disease. Probiotic lactobacilli and Lactobacillus metabolites can inhibit fungal growth
(Crowley et al. 2012, 2013b, Guo et al. 2012, Hatakka et al. 2007, Hietala et al. 1979,
Wagner et al. 2000). This study demonstrated that the Lactobacillus metabolite HICA,
a leucine derivative, can effectively reduce the growth and biofilm formation of C.
albicans in vitro. HICA has previously demonstrated broad-spectrum antimicrobial
activity against various planktonically grown bacterial and fungal species in vitro
(Sakko et al. 2013, Sakko et al. 2012). Both the metabolic activity and biomass of C.
albicans biofilms were significantly reduced by HICA in different phases of biofilm
growth, and defects in biofilm ultrastructure were confirmed by SEM. Major inhibitory
activity of HICA was measured at an acidic pH. This is common for Lactobacillus
metabolites. HICA also showed markedly improved efficacy over caspofungin, a
commonly used antifungal against Candida biofilms (Cornely et al. 2012, Pfaller et al.
2011).
79
Gene expression analysis revealed highly significant basal and relative expression of
CaADH1 and CaALD5 genes in biofilms due to challenge with particularly HICA and
cysteine, among others, in vitro. Considering the well-known toxicity of cysteine to
fungal cells at high concentrations, these results underline the importance of CaADH1
and CaALD5 in stress resistance and further support their broader role in biofilm
growth. CaALD5 encodes the mitochondrial isoenzyme in S. cerevisiae. Interestingly,
the C. albicans CaAld5p shows homology to mouse ALDH2, which is highly
homologous to human ALDH2 (Chang and Yoshida 1994). Human and mouse
mitochondrial ALDH2 has been associated with stress resistance. In addition, ACH
levels in biofilms significantly correlated with CaALD5 expression. These findings imply
that aldehyde dehydrogenases have similar roles in different eukaryotic organisms
(Baek et al. 2004, Ohsawa et al. 2003, Szocs et al. 2007). No ACH was produced by any
of the HICA-treated biofilms at an acidic pH, which may suggest an impairment of
central carbon metabolism and more specifically glycolysis. This has earlier been
shown as the potential mode of action for Lactobacillus metabolites (Crowley et al.
2013a, Kohler et al. 2012).
No anti-biofilm activity for HICA or caspofungin was measured in a murine infection
model. A longer treatment regimen may be required for an anti-biofilm effect, as
reported in other animal models (Walraven and Lee 2013). However, an attenuated
inflammatory response was observed in subcutaneous tissue sections surrounding an
HICA-treated biofilm chamber. Lower expression was described, particularly for MMP9 and MPO, a tissue destructive protease and its activator highly expressed in cascade
in various inflammatory conditions (Hajishengallis 2014, Sorsa et al. 2006). The
present study did not establish whether the anti-inflammatory effect observed was
direct or indirect, or a combination of these. In the case of an indirect effect,
attenuation of the inflammatory response could be secondary to the efficient
eradication of C. albicans biofilms. The potential direct anti-inflammatory effect of
HICA is supported by a few previous studies that have demonstrated attenuated
inflammation in response to treatment with Lactobacillus metabolites (Jones et al.
2009, Ramos et al. 2010). In addition, antifungals can possess direct anti-inflammatory
activity, as shown for clotrimazole during the treatment of vaginal candidosis (Wilson
et al. 2013). The direct effect is supported by the findings of a previous in vitro study
in which an anti-proteolytic effect was described for HICA (Westermarck et al. 1997).
Del-1 was strongly expressed in the non-infected, non-biofilm group, as expected for
an inhibitor of neutrophil extravasation. Interestingly, Del-1 was also strongly
expressed in the HICA biofilm group. Although Del-1 has been shown to play an
important role in the regulation of immune responses in various inflammatory
diseases, this study reports for the first time its expression in fungal infection (Baban
80
et al. 2013, Eskan et al. 2012, Shin et al. 2013, Zhao et al. 2014). Del-1 can also
regulate the expression of IL-17, which plays an important role in mucosal immunity
against candidosis (Conti et al. 2009, Eskan et al. 2012, Ho et al. 2010, Huppler et al.
2014, Puel et al. 2011). Thus, Del-1 could play a role in various Candida-related
inflammatory conditions.
9.2
LIMITATIONS OF THE STUDY
ACH production by Candida species was only investigated in vitro and mainly with
single-species cultures. These settings are drastically different from the in vivo
conditions faced by Candida in the upper digestive tract. The number of Candida cells
can be much lower at a single locus in vivo than used here in vitro. Multiple
environmental and inhibitory factors affect candidal metabolism and biofilm
formation. Competitive inhibition and synergy with bacteria also exist, further shaping
the metabolism and development of the fungal cell and biofilm in vivo.
The level of oxygen is a major factor affecting local ACH production. Oxygen levels
regulate the balance between respiration and fermentation in Candida spp. and other
members of the microbiome. The availability of oxygen can widely largely between
various niches of the oral cavity and oesophagus, and oxygen concentrations ranging
from 0.5% to 14% have been observed (Marcotte and Lavoie 1998, Theilade 1990).
High levels of oxygen can be measured in saliva (5–14% O2), and hypoxic levels in deep
gingival pockets or in the deeper layers of oral polymicrobial biofilms (<1% O2). In
comparison, the normal oxygen concentration in the air is around 21% . The low level
of oxygen at the core of the biofilm is a result of reduced oxygen diffusion and the
consumption of a major proportion of the available oxygen by facultative anaerobic
microbes. In the upper digestive microbiome, metabolically distinct microbial
members are present in terms of oxygen sensitivity, and members ranging from
obligate anaerobes to aerobes exist. Under anaerobic conditions, fermentative
metabolism is up-regulated and ethanol is produced from sugars via the reduction of
ACH. When oxygen is present, ethanol derived from endogenous or exogenous
sources is oxidised to ACH. Oxygen levels in saliva support this oxidative metabolism
of ethanol. An additional parameter driving microbial metabolism in hypoxia is the
higher level of CO2. In addition to microbial ethanol metabolism, mucosal cells also
have the enzymatic capacity to oxidise ethanol into ACH. Thus, multiple factors affect
the overall ACH exposure in vivo. Although the microbiota dominantly contributes to
local ACH exposure, no study has described the relative contribution of microbial and
host cells to the total ACH exposure. Overall, enhanced models simulating the in vivo
conditions are required.
81
Although cells were cultured under aerobic growth conditions, the levels of oxygen
faced by the planktonic cells or biofilms at the bottom of the plastic wells or glass vials
filled with media or buffer were more likely to be moderately hypoxic. This is
particularly the case when cultures are grown statically, as in this study (Odds et al.
1995). Moderately hypoxic conditions have been shown to stimulate high ACH
production by C. albicans (Marttila et al. 2013a). The strain and various growth
parameters, such as nutrient availability, time of biofilm development, pH and the
type of in vivo model, can have a major impact on the metabolic outcome of the
fungal cell and biofilm development in vitro and in vivo, as shown in the literature
(Kucharikova et al. 2011, Maccallum 2012, Martinez-Gomariz et al. 2009). Variation in
these growth parameters would have potentially resulted in more variation in the
results and in better understanding of the effects of various compounds tested in this
study.
The number of isolates per species under investigation was low, although both clinical
and reference strains were selected for the experiments with planktonic cultures.
Biofilm experiments were mainly carried out using one reference strain. Including
more Candida isolates from clinical samples would provide vital information and
produce a broader picture of how biofilm formation and central metabolic pathways,
including ACH metabolism, are affected by the microevolution of each isolate.
Biofilm biomass was calculated based on the concentration of double stranded DNA
(dsDNA), which has been shown to correlate with number of viable cells (Marstorp
and Witter 1999). The biofilm matrix can contain extracellular DNA (eDNA), which
facilitates the formation, resistance and inflammatory properties of fungal and
bacterial biofilms (Martins et al. 2010, Rajendran et al. 2013, Whitchurch et al. 2002).
Therefore, eDNA of the matrix could affect our results, although we tried to limit this
by including multiple washing steps. Reliable quantitation of Candida biofilms can be
problematic, as previously documented (Taff et al. 2012). Furthermore, it is important
to acknowledge the recent data presented by Harrison et al. (2014), which showed
that pathogenic Candida yeasts can shift the ploidy of the cell in response to certain
stress factors, such as antifungals, in vitro and in vivo. This shift increases the dsDNA
content of the cell. Thus, one can question how this phenomenon possibly affects the
results described in this study.
When gene expression levels were measured, only the main genes linked to ethanol
metabolism were analysed. Although both basal and relative expression was
calculated against a standard reference gene, these results remain restricted in terms
of the total transcriptomic and metabolic outcome of the biofilm. Therefore, more
82
extensive genomic and proteomic studies are warranted in the future. It should also
be taken into account that the availability of oxygen and nutrients can differ within
the different layers of a biofilm. Thus, the cellular metabolic and transcriptomic
responses can considerably differ between the core and surface of the biofilm. It is
common that the RNA is extracted from the whole biofilm. Therefore, the
transcriptional changes shown describe the average expression of the specific gene in
the biofilm and not the expression in a certain layer or niche. Inflammatory responses
were investigated in a murine infection model. The inflammatory responses can
greatly differ between humans and mice (Seok et al. 2013). Thus, the results cannot
be fully extrapolated to the human host. Furthermore, the number of mice,
particularly in the control groups, was low. Therefore, in vivo studies with larger
groups are warranted to strengthen these findings.
9.3
STRENGTHS OF THE STUDY
Overall, the results of this study demonstrate the considerable capacity of various
Candida species to oxidise ethanol to mutagenic ACH. High production of ACH by NAC
species is clinically important, as NAC species have become increasingly frequent
concurrently with the rise in the number of immunocompromised patients (Pfaller et
al. 2011). The concentrations of ethanol (11-12 mM) used in the experiments can be
considered clinically relevant, as they are normally found in saliva in an average male
(80 kg) after consuming 0.5–0.6 g of ethanol per kilogram of body weight.
Concentrations of glucose and fructose were adjusted to 18–20 g/l (100–110 mM),
commonly found in the Western diet. It was important to identify a compound with
high efficacy against C. albicans biofilms. New approaches to treat fungal infections
are required, as the current antifungals act poorly, especially against biofilms, and
harmful side effects often accompany their use. HICA is a safe compound, as it is
found in the human body and has been used as an animal food and by professional
athletes (Boebel and Baker 1982, Hietala et al. 1979, Mero et al. 2010). The antiinflammatory properties of HICA could provide additional benefits when combating
inflammatory disease associated with irreversible tissue damage (Westermarck et al.
1997). HICA could also be used in combination with other antifungals, such as
echinocandins or azoles, to strengthen the inhibitory effect. Future studies are
required to elucidate the potential effect of combined therapy. The properties of the
inflammatory response that a C. albicans biofilm induces in vivo are not well known.
Thus, this study contributed to the limited information currently existing and
demonstrated distinct cytokine expression, similar to what has been described in vitro
(Chandra et al. 2007, Xie et al. 2012).
83
9.4
RELEVANCE OF THE STUDY
Candida yeasts have been shown to be associated with pre-malignant lesions and
carcinomas in the oral cavity and oesophagus (McCullough et al. 2002, Nagy et al.
1998, Rautemaa et al. 2006, Warnakulasuriya et al. 2008b). Interestingly, carcinoma
often develops next to the site of biofilm-related Candida mucositis. Local infection
causes inflammation in neighbouring tissues, and in the long-term this process can
lead to dysplasia. Carcinogenesis is initiated when the threshold of DNA damage is
exceeded in a normal cell, although strong defence involving cellular repair
mechanisms exists. Increased production of mutagenic ACH in the presence of
suitable carbon sources due to colonisation with C. albicans or NAC species could
induce mutagenic changes and add to other stress factors, such as inflammation, in
mucosal cells, and eventually contribute to the initiation of cancer. Mutagenic ACH
could also affect the neighbouring flora and cause mutations leading to enhanced
virulence and particularly to the development of resistance, which is a major clinical
problem worldwide. Candida yeasts were also associated with the formation of
mutagenic levels of ACH during mursik fermentation. Regular consumption of
fermented products, such as mursik, with a high ACH content could further increase
the stress faced by mucosal cells. A high ethanol content present as a congener could
be oxidised to ACH by microbes, especially yeasts, thus further increasing the
exposure. HICA could provide a safe approach to inhibit biofilm growth by C. albicans
in mucosal infections and reduce the mutagenicity and inflammation associated with
the pathogenic biofilm lifestyle.
84
10
CONCLUSIONS
This study led to the following conclusions:
I
C. albicans can produce mutagenic levels of acetaldehyde from ethanol and
glucose in both planktonic and biofilm modes of growth in vitro. C. albicans is
the most common fungus in the upper digestive tract and its biofilm lifestyle
is associated with pathogenicity. Carcinoma often develops next to the site of
biofilm-related Candida mucositis. Thus, carcinogenic acetaldehyde produced
by C. albicans biofilm could have a major role in oral and oesophageal
carcinogenesis.
II
Experiments with planktonic cells demonstrated the significant ability of
clinically relevant non-albicans Candida species to produce acetaldehyde in
the presence of ethanol in vitro. However, significant differences were
detected between and within species. Interestingly, co-incubation with xylitol
significantly reduced acetaldehyde production from ethanol by all Candida
spp. C. glabrata shows a specific potential for acetaldehyde production, as
mutagenic levels could be measured in the presence of glucose and also
fructose. Acetaldehyde production by C. glabrata was increased in glucoseethanol and glucose-fructose co-incubations. These findings and also the high
Adh activity detected in C. glabrata demonstrate its high tendency for
fermentative metabolism. As the prevalence of non-albicans Candida species
in health and disease is increasing, colonisation with species such as C.
glabrata could significantly increase the production of carcinogenic
acetaldehyde locally when ethanol or sugars are present and add to the risk
of cancer in the upper digestive tract. Xylitol could be potentially used to
reduce local acetaldehyde exposure by Candida spp.
III
Regular consumption of fermented food products and beverages containing
acetaldehyde and ethanol has been associated with upper digestive tract
carcinogenesis. Consumption can lead to direct exposure of the mucosa to
carcinogenic acetaldehyde or indirect exposure through microbial ethanol
metabolism and salivary acetaldehyde. Highly mutagenic levels of
acetaldehyde and ethanol are produced during mursik milk fermentation, as
simulated in this study. Specific Candida–Lactobacillus combinations are
associated with the highest acetaldehyde and ethanol production, thus
showing the role of microbial synergy in the process. Yeasts have a major role
in the production of acetaldehyde. As mursik constitutes a major part of the
85
diet in western Kenya, an area with a high incidence of oesophageal cancer
and a low prevalence of common risk factors, the regular consumption of
mursik could contribute to the unusually high cancer incidence.
IV
86
Yeasts are eukaryotic organisms. Their similarity to human cells makes the
development of antifungals difficult, and patient compliance is often
compromised because of various side effects. Treating Candida infections is a
challenging task, and more effective antifungal agents are required. This
study demonstrated the major efficacy of Lactobacillus and the leucine
metabolite HICA against C. albicans biofilms in vitro. During challenge with
HICA and comparators, certain genes related to ethanol metabolism were upregulated, thus showing their important role in stress resistance and biofilm
formation in general. Furthermore, HICA attenuated the inflammatory
response in a murine C. albicans biofilm infection model. This combination of
anti-biofilm and anti-inflammatory activity can be highly beneficial when
treating biofilm-related chronic inflammatory disease. Effective eradication
of Candida biofilms with HICA could reduce the exposure of the neighbouring
mucosa and surrounding microbiome to mutagenic acetaldehyde and reduce
the biofilm-induced inflammatory burden on mucosal cells.
11
ACKNOWLEDGEMENTS
This PhD thesis was a collaborative project between the University of Helsinki, Finland
and the University of Manchester, United Kingdom during the years 2009-2014. The
experiments were performed at the Research Unit on Acetaldehyde and Cancer,
Biomedicum, University of Helsinki (I,II), at the Department of Bacteriology and
Immunology, Haartman Institute, University of Helsinki (III) and at Research lab of
Dental Institute, Biomedicum, University of Helsinki (I,II,V) and at University Hospital
of Southern Manchester, Wythenshawe Hospital, University of Manchester (III,IV,V).
In addition, Candida biofilm model was studied at the University of Glasgow under
supervision of Prof Gordon Ramage prior to construction of our own biofilm model at
Wythenshawe hospital. Study III was performed in collaboration with National Cancer
Institute of National Institutes of Health (NIH), USA and the samples were collected by
their field researchers in Western Kenya. I sincerely thank Dr Neal Freedman and his
co-workers at NIH for their contribution and effort.
First and foremost I am deeply grateful for my supervisors Riina Richardson, Mikko
Salaspuro and Timo Sorsa for their excellent guidance during this project.
Riina, I sincerely thank you for your endless enthusiasm, optimism and care. Your
determination and passion towards science has pushed me forward. The wisdom and
vast knowledge you possess has been something to admire. I also thank you for the
introduction to English culture together with your family, the accommodation during
my visits to Manchester and the spare time shared. Mikko, I kindly thank you for all
the support given especially in the beginning of my PhD studies. Your expertise,
innovative thinking and contribution to the field of GI tract carcinogenesis and
acetaldehyde-related toxicity has been of world-class and provided the framework for
this project. Timo, I greatly appreciate your knowledge and wide experience in
science, especially in the area of inflammation and tissue proteases. I also deeply
honour your integrity and open-mindedness. In addition, the discussions with you
inside and outside the field of science have been a great joy.
I wish to express my deep appreciation to my reviewers Prof Bernhard Hube and Prof
Markus Haapasalo for their valuable critique and in-depth commentary. I kindly thank
Roy Siddall for the excellent language revision.
I most sincerely thank Dr Paul Bowyer (Univ. of Manchester) for sharing his resources
and the world-class expertise in molecular microbiology. Dr Peter Warn (Univ. of
87
Manchester) is highly appreciated for the collaboration and work regarding animal
and infection models.
I am deeply indebted for Dr Lily Novak-Frazer (Univ. of Manchester) for all her hard
work and superior guidance during these years in the research lab of Wythenshawe
hospital. Your determination and precision were vital for the successful establishment
of the biofilm model and the progress made during the past three years. I deeply
honour your endless support and effort given. I also want to express my gratitude
towards other co-researchers in the lab, particularly Vilma Rautemaa, Rebecca Collins,
Mike Bromley and Marcin Fraczek.
Dr Julie Morris (Univ. of Manchester) is acknowledged for her expert guidance on
statistical analysis.
I highly honour the role of Prof Malcolm Richardson in sharing his expertise in
mycology. Your contribution to the manuscripts, particularly regarding their language,
cannot be thanked enough. The discussions around the diverse aspects of science and
technology have been highly amusing.
I kindly thank Prof Gordon Ramage and Dr Ranjith Rajendran from University of
Glasgow for teaching me the fundamentals of Candida biofilms in the beginning of my
research projects in UK. I am also deeply grateful for the resources and the excellent
collaboration and support shared.
Among the team members of Mikko Salaspuro, I especially want to acknowledge the
work of Maarit Raukola. Your conribution for the lab has been irreplaceable and the
discussions shared a great joy. The contribution and help of Pertti Kaihovaara in the
field of gas chromatography is highly acknowledged. I owe also special thanks to other
co-workers, particularly Johanna Uittamo, Satu Väkeväinen, Klas Linderborg and
Andreas Helminen for their support. In addition, I want to thank the Finnish
colleagues belonging to Riina’s team: Lotta Grönholm, Peter Rusanen, Esa Färkkilä,
Marjus Sakko and Emilia Siikala, it has been wonderful to work with you and share the
experiences around PhD studies.
Among the staff of the Research lab of Dental Institute of the University of Helsinki, I
want to acknowledge the work of Marjatta Kivekäs and Dr Taina Tervahartiala. I most
sincerely thank you for the contribution and knowledge shared regarding
immunohistochemistry. I kindly thank Dr Kirsti Kari for providing me excellent working
facilities. I also thank the other co-workers, particularly Dr Pirkko Pussinen and her
team members Elisa Kallio, Kati Hyvärinen, and Aino Salminen for the support and
88
joyful discussions. Collaborator of Timo Sorsa, Prof Marcela Hernandez from the
University of Chile is highly acknowledged for her advice and contribution regarding
immunohistochemistry and pathology.
This work is dedicated to my parents, Anna-Liisa and Markku, who have given their
endless support throughout my life and encouraged me to pursuit the dreams I have. I
am more than grateful for the rest of the family as well, especially my little sister and
brother Riitta and Martti.
I also owe the deepest gratitude to my dearest, Fanny, for all her love and friendship.
Your humour and endless support has given me tons of energy during this challenging,
but highly rewarding and enjoyable step of my career.
I am more than grateful to all my friends in Finland and abroad for pulling me out of
the hectic academic and work life once in a while, for putting things in life into
perspective and for all the amazing moments experienced.
This study was financially supported by Sigrid Juselius Foundation, Orion Research
Foundation, Helsinki University Central Hospital Research Fund, Dental Society of
Apollonia and Yrjö Jahnsson Foundation.
Mikko Nieminen
th
Helsinki, August 5 2014
89
12
REFERENCES
Aas JA, Paster BJ, Stokes LN, Olsen I, Dewhirst FE. Defining the normal bacterial flora of the oral
cavity. J Clin Microbiol 2005; 43: 5721-5732.
Abnet CC, Kamangar F, Islami F, Nasrollahzadeh D, Brennan P, Aghcheli K, Merat S, Pourshams
A, Marjani HA, Ebadati A, Sotoudeh M, Boffetta P, Malekzadeh R, Dawsey SM. Tooth loss
and lack of regular oral hygiene are associated with higher risk of esophageal squamous cell
carcinoma. Cancer Epidemiol Biomarkers Prev 2008; 17: 3062-3068.
Agarwal DP, Goedde HW. Pharmacogenetics of alcohol dehydrogenase (ADH). Pharmacol Ther
1990; 45: 69-83.
Ahn J, Chen CY, Hayes RB. Oral microbiome and oral and gastrointestinal cancer risk. Cancer
Causes Control 2012; 23: 399-404.
Alby K, Bennett RJ. Stress-induced phenotypic switching in Candida albicans. Mol Biol Cell 2009;
20: 3178-3191.
Alexander BD, Johnson MD, Pfeiffer CD, Jimenez-Ortigosa C, Catania J, Booker R, Castanheira
M, Messer SA, Perlin DS, Pfaller MA. Increasing echinocandin resistance in Candida glabrata:
clinical failure correlates with presence of FKS mutations and elevated minimum inhibitory
concentrations. Clin Infect Dis 2013; 56: 1724-1732.
Aleynik SI, Leo MA, Aleynik MK, Lieber CS. Increased circulating products of lipid peroxidation in
patients with alcoholic liver disease. Alcohol Clin Exp Res 1998; 22: 192-196.
Almeida RS, Brunke S, Albrecht A, Thewes S, Laue M, Edwards JE, Filler SG, Hube B. the hyphalassociated adhesin and invasin Als3 of Candida albicans mediates iron acquisition from host
ferritin. PLoS Pathog 2008; 4: e1000217.
Altschul SF, Gish W, Miller W, Myers EW, Lipman DJ. Basic local alignment search tool. J Mol Biol
1990; 215: 403-410.
Andes D, Nett J, Oschel P, Albrecht R, Marchillo K, Pitula A. Development and characterization of
an in vivo central venous catheter Candida albicans biofilm model. Infect Immun 2004; 72:
6023-6031.
Andes DR, Safdar N, Baddley JW, Playford G, Reboli AC, Rex JH, Sobel JD, Pappas PG, Kullberg
BJ. Impact of treatment strategy on outcomes in patients with candidemia and other forms of
invasive candidiasis: a patient-level quantitative review of randomized trials. Clin Infect Dis
2012; 54: 1110-1122.
Arendorf TM, Walker DM. The prevalence and intra-oral distribution of Candida albicans in man.
Arch Oral Biol 1980; 25: 1-10.
Arning M, Kliche KO, Heer-Sonderhoff AH, Wehmeier A. Infusion-related toxicity of three different
amphotericin B formulations and its relation to cytokine plasma levels. Mycoses 1995; 38: 459465.
Ashley ESD, Lewis R, Lewis JS, Martin C, Andes D. Pharmacology of systemic antifungal agents.
Clin Infect Dis 2006; 43: S28-S39.
90
Askew C, Sellam A, Epp E, Hogues H, Mullick A, Nantel A, Whiteway M. Transcriptional regulation
of carbohydrate metabolism in the human pathogen Candida albicans. PLoS pathogens 2009;
5: e1000612.
Atanassova M, Choiset Y, Dalgalarrondo M, Chobert JM, Dousset X, Ivanova I, Haertle T. Isolation
and partial biochemical characterization of a proteinaceous anti-bacteria and anti-yeast
compound produced by Lactobacillus paracasei subsp. paracasei strain M3. Int J Food
Microbiol 2003; 87: 63-73.
Avon SL, Goulet JP, Deslauriers N. Removable acrylic resin disk as a sampling system for the
study of denture biofilms in vivo. J Prosthet Dent 2007; 97: 32-38.
Baban B, Liu JY, Abdelsayed R, Mozaffari MS. Reciprocal relation between GADD153 and Del-1 in
regulation of salivary gland inflammation in Sjögren syndrome. Exp Mol Pathol 2013; 95: 288297.
Baboni FB, Barp D, Izidoro AC, Samaranayake LP, Rosa EA. Enhancement of Candida albicans
virulence after exposition to cigarette mainstream smoke. Mycopathologia 2009; 168: 227-235.
Baehr PH, McDonald GB. Esophageal infections: risk factors, presentation, diagnosis, and
treatment. Gastroenterology 1994; 106: 509-532.
Baek HY, Lim JW, Kim H, Kim JM, Kim JS, Jung HC, Kim KH. Oxidative-stress-related proteome
changes in Helicobacter pylori-infected human gastric mucosa. Biochem J 2004; 379: 291299.
Balbo S, Meng L, Bliss RL, Jensen JA, Hatsukami DK, Hecht SS. Kinetics of DNA adduct formation
in the oral cavity after drinking alcohol. Cancer Epidemiol Biomarkers Prev 2012; 21: 601-608.
Balish E, Wagner RD, Vazquez-Torres A, Jones-Carson J, Pierson C, Warner T. Mucosal and
systemic candidiasis in IL-8Rh-/- BALB/c mice. J Leukoc Biol 1999; 66: 144-150.
Bamford CV, d'Mello A, Nobbs AH, Dutton LC, Vickerman MM, Jenkinson HF. Streptococcus
gordonii modulates Candida albicans biofilm formation through intergeneric communication.
Infect Immun 2009; 77: 3696-3704.
Barasch A, Griffin AV. Miconazole revisited: new evidence of antifungal efficacy from laboratory
and clinical trials. Future Microbiol 2008; 3: 265-269.
Barelle CJ, Priest CL, Maccallum DM, Gow NA, Odds FC, Brown AJ. Niche-specific regulation of
central metabolic pathways in a fungal pathogen. Cell Microbiol 2006; 8: 961-971.
Barkvoll P, Attramadal A. Effect of nystatin and chlorhexidine digluconate on Candida albicans.
Oral Surg Oral Med Oral Pathol 1989; 67: 279-281.
Barnett JA. A history of research on yeasts 6: the main respiratory pathway. Yeast 2003; 20: 10151044.
Barrett AW, Kingsmill VJ, Speight PM. The frequency of fungal infection in biopsies of oral mucosal
lesions. Oral Dis 1998; 4: 26-31.
91
Basuroy S, Sheth P, Mansbach CM, Rao RK. Acetaldehyde disrupts tight junctions and adherens
junctions in human colonic mucosa: protection by EGF and L-glutamine. Am J Physiol
Gastrointest Liver Physiol 2005; 289: G367-375.
Ben-Ami R, Lewis RE, Kontoyiannis DP. Immunocompromised hosts: immunopharmacology of
modern antifungals. Clin Infect Dis 2008; 47: 226-235.
Bertram G, Swoboda RK, Gooday GW, Gow NA, Brown AJ. Structure and regulation of the
Candida albicans ADH1 gene encoding an immunogenic alcohol dehydrogenase. Yeast 1996;
12: 115-127.
Bik EM, Long CD, Armitage GC, Loomer P, Emerson J, Mongodin EF, Nelson KE, Gill SR, FraserLiggett CM, Relman DA. Bacterial diversity in the oral cavity of 10 healthy individuals. ISME J
2010; 4: 962-974.
Bjarnason I, Peters TJ, Wise RJ. The leaky gut of alcoholism: possible route of entry for toxic
compounds. Lancet 1984; 1: 179-182.
Boccia S, Hashibe M, Galli P, De Feo E, Asakage T, Hashimoto T, Hiraki A, Katoh T, Nomura T,
Yokoyama A, van Duijn CM, Ricciardi G, Boffetta P. Aldehyde dehydrogenase 2 and head and
neck cancer: a meta-analysis implementing a Mendelian randomization approach. Cancer
Epidemiol Biomarkers Prev 2009; 18: 248-254.
Bode JC, Knuppel H, Schwerk W, Lorenz-Meyer H, Durr HK. Quantitative histomorphometric study
of the jejunal mucosa in chronic alcoholics. Digestion 1982; 23: 265-270.
Bodhade AS, Ganvir SM, Hazarey VK. Oral manifestations of HIV infection and their correlation
with CD4 count. J Oral Sci 2011; 53: 203-211.
Boebel KP, Baker DH. Comparative utilization of the alpha-keto and D- and L-alpha-hydroxy
analogs of leucine, isoleucine and valine by chicks and rats. J Nutr 1982; 112: 1929-1939.
Boffetta P, Hashibe M, La Vecchia C, Zatonski W, Rehm J. The burden of cancer attributable to
alcohol drinking. Int J Cancer 2006; 119: 884-887.
Boffetta P, Kaihovaara P, Rudnai P, Znaor A, Lissowska J, Swiatkowska B, Mates D, Pandics T,
Salaspuro M. Acetaldehyde level in spirits from central European countries. Eur J Cancer Prev
2011; 20: 526-529.
Bonacini M, Young T, Laine L. The causes of esophageal symptoms in human immunodeficiency
virus infection. A prospective study of 110 patients. Arch Intern Med 1991; 151: 1567-1572.
Bonhomme J, Chauvel M, Goyard S, Roux P, Rossignol T, d'Enfert C. Contribution of the glycolytic
flux and hypoxia adaptation to efficient biofilm formation by Candida albicans. Mol Microbiol
2011; 80: 995-1013.
Boris S, Barbes C. Role played by lactobacilli in controlling the population of vaginal pathogens.
Microbes Infect 2000; 2: 543-546.
Bosron WF, Li TK. Genetic polymorphism of human liver alcohol and aldehyde dehydrogenases,
and their relationship to alcohol metabolism and alcoholism. Hepatology 1986; 6: 502-510.
92
Brand A, Gow NA. Mechanisms of hypha orientation of fungi. Curr Opin Microbiol 2009; 12: 350357.
Brand A, Vacharaksa A, Bendel C, Norton J, Haynes P, Henry-Stanley M, Wells C, Ross K, Gow
NA, Gale CA. An internal polarity landmark is important for externally induced hyphal
behaviors in Candida albicans. Eukaryot Cell 2008; 7: 712-720.
Brooks PJ, Theruvathu JA. DNA adducts from acetaldehyde: implications for alcohol-related
carcinogenesis. Alcohol 2005; 35: 187-193.
Brown GD, Denning DW, Gow NA, Levitz SM, Netea MG, White TC. Hidden killers: human fungal
infections. Sci Transl Med 2012; 4: 165rv113.
Brummer E, Stevens DA. Synergy of human neutrophils with fluconazole in killing Candida species.
Mycopathologia 1996; 134: 115-120.
Buhl R. Local oropharyngeal side effects of inhaled corticosteroids in patients with asthma. Allergy
2006; 61: 518-526.
Butler G, Rasmussen MD, Lin MF, Santos MA, Sakthikumar S, Munro CA, Rheinbay E, Grabherr
M, Forche A, Reedy JL, Agrafioti I, Arnaud MB, Bates S, Brown AJ, Brunke S, Costanzo MC,
Fitzpatrick DA, de Groot PW, Harris D, Hoyer LL, Hube B, Klis FM, Kodira C, Lennard N,
Logue ME, Martin R, Neiman AM, Nikolaou E, Quail MA, Quinn J, Santos MC, Schmitzberger
FF, Sherlock G, Shah P, Silverstein KA, Skrzypek MS, Soll D, Staggs R, Stansfield I, Stumpf
MP, Sudbery PE, Srikantha T, Zeng Q, Berman J, Berriman M, Heitman J, Gow NA, Lorenz
MC, Birren BW, Kellis M, Cuomo CA. Evolution of pathogenicity and sexual reproduction in
eight Candida genomes. Nature 2009; 459: 657-662.
Calderone RA, Clancy CJ. Candida and Candidiasis. 2nd ed: ASM Press; 2011.
Cannon RD, Chaffin WL. Oral colonization by Candida albicans. Crit Rev Oral Biol Med 1999; 10:
359-383.
Carlson E. Effect of strain of Staphylococcus aureus on synergism with Candida albicans resulting
in mouse mortality and morbidity. Infect Immun 1983; 42: 285-292.
Carman AJ, Vylkova S, Lorenz MC. Role of acetyl coenzyme A synthesis and breakdown in
alternative carbon source utilization in Candida albicans. Eukaryot Cell 2008; 7: 1733-1741.
Cawson RA. Leukoplakia and oral cancer. Proc R Soc Med 1969; 62: 610-615.
Cenci E, Mencacci A, Del Sero G, Bistoni F, Romani L. Induction of protective Th1 responses to
Candida albicans by antifungal therapy alone or in combination with an interleukin-4
antagonist. J Infect Dis 1997; 176: 217-226.
Cha R, Sobel JD. Fluconazole for the treatment of candidiasis: 15 years experience. Expert Rev
Anti Infect Ther 2004; 2: 357-366.
Chandra J, Kuhn DM, Mukherjee PK, Hoyer LL, McCormick T, Ghannoum MA. Biofilm formation by
the fungal pathogen Candida albicans: development, architecture, and drug resistance. J
Bacteriol 2001; 183: 5385-5394.
93
Chandra J, McCormick TS, Imamura Y, Mukherjee PK, Ghannoum MA. Interaction of Candida
albicans with adherent human peripheral blood mononuclear cells increases C. albicans
biofilm formation and results in differential expression of pro- and anti-inflammatory cytokines.
Infect Immun 2007; 75: 2612-2620.
Chang C, Yoshida A. Cloning and characterization of the gene encoding mouse mitochondrial
aldehyde dehydrogenase. Gene 1994; 148: 331-336.
Cheng SC, Joosten LA, Kullberg BJ, Netea MG. Interplay between Candida albicans and the
mammalian innate host defense. Infect Immun 2012; 80: 1304-1313.
Cherry JM, Hong EL, Amundsen C, Balakrishnan R, Binkley G, Chan ET, Christie KR, Costanzo
MC, Dwight SS, Engel SR, Fisk DG, Hirschman JE, Hitz BC, Karra K, Krieger CJ, Miyasato
SR, Nash RS, Park J, Skrzypek MS, Simison M, Weng S, Wong ED. Saccharomyces Genome
Database: the genomics resource of budding yeast. Nucleic Acids Res 2012; 40: D700-705.
Chiang CP, Jao SW, Lee SP, Chen PC, Chung CC, Lee SL, Nieh S, Yin SJ. Expression pattern,
ethanol-metabolizing activities, and cellular localization of alcohol and aldehyde
dehydrogenases in human large bowel: association of the functional polymorphisms of ADH
and ALDH genes with hemorrhoids and colorectal cancer. Alcohol 2012; 46: 37-49.
Clarkson JE, Worthington HV, Eden OB. Interventions for preventing oral candidiasis for patients
with cancer receiving treatment. Cochrane Database Syst Rev 2007; (1): CD003807.
Claveau I, Mostefaoui Y, Rouabhia M. Basement membrane protein and matrix metalloproteinase
deregulation in engineered human oral mucosa following infection with Candida albicans.
Matrix Biol 2004; 23: 477-486.
Conti HR, Shen F, Nayyar N, Stocum E, Sun JN, Lindemann MJ, Ho AW, Hai JH, Yu JJ, Jung JW,
Filler SG, Masso-Welch P, Edgerton M, Gaffen SL. Th17 cells and IL-17 receptor signaling are
essential for mucosal host defense against oral candidiasis. J Exp Med 2009; 206: 299-311.
Cope K, Risby T, Diehl AM. Increased gastrointestinal ethanol production in obese mice:
implications for fatty liver disease pathogenesis. Gastroenterology 2000; 119: 1340-1347.
Cornely OA, Bassetti M, Calandra T, Garbino J, Kullberg BJ, Lortholary O, Meersseman W, Akova
M, Arendrup MC, Arikan-Akdagli S, Bille J, Castagnola E, Cuenca-Estrella M, Donnelly JP,
Groll AH, Herbrecht R, Hope WW, Jensen HE, Lass-Florl C, Petrikkos G, Richardson MD,
Roilides E, Verweij PE, Viscoli C, Ullmann AJ, Group EFIS. ESCMID* guideline for the
diagnosis and management of Candida diseases 2012: non-neutropenic adult patients. Clin
Microbiol Infect 2012; 18 Suppl 7: 19-37.
Cornely OA, Ullmann AJ, Karthaus M. Evidence-based assessment of primary antifungal
prophylaxis in patients with hematologic malignancies. Blood 2003; 101: 3365-3372.
Corvo R, Amichetti M, Ascarelli A, Arcangeli G, Buffoli A, Cellini N, Cionini L, De Renzis C, Emiliani
E, Franchini P, Gabriele P, Gobitti C, Grillo Ruggieri F, Bertoni F, Magrini SM, Marmiroli L,
Orsatti M, Panizza GM, Tordiglione M, Ziccarelli L, Gava A, Zorat PL, Ghelfi R, Serra GF,
Vitale V. Effects of fluconazole in the prophylaxis of oropharyngeal candidiasis in patients
undergoing radiotherapy for head and neck tumour: results from a double-blind placebocontrolled trial. Eur J Cancer Care (Engl) 2008; 17: 270-277.
Costerton JW, Geesey GG, Cheng KJ. How bacteria stick. Sci Am 1978; 238: 86-95.
94
Costerton JW, Stewart PS, Greenberg EP. Bacterial biofilms: a common cause of persistent
infections. Science 1999; 284: 1318-1322.
Coutelle C, Hohn B, Benesova M, Oneta CM, Quattrochi P, Roth HJ, Schmidt-Gayk H,
Schneeweiss A, Bastert G, Seitz HK. Risk factors in alcohol associated breast cancer: alcohol
dehydrogenase polymorphism and estrogens. Int J Oncol 2004; 25: 1127-1132.
Crabb DW, Edenberg HJ, Bosron WF, Li TK. Genotypes for aldehyde dehydrogenase deficiency
and alcohol sensitivity. The inactive ALDH2(2) allele is dominant. J Clin Invest 1989; 83: 314316.
Crabb DW, Matsumoto M, Chang D, You M. Overview of the role of alcohol dehydrogenase and
aldehyde dehydrogenase and their variants in the genesis of alcohol-related pathology. Proc
Nutr Soc 2004; 63: 49-63.
Crowe JD, Sievwright IK, Auld GC, Moore NR, Gow NA, Booth NA. Candida albicans binds human
plasminogen: identification of eight plasminogen-binding proteins. Mol Microbiol 2003; 47:
1637-1651.
Crowley S, Mahony J, Morrissey JP, van Sinderen D. Transcriptomic and morphological profiling of
Aspergillus fumigatus Af293 in response to antifungal activity produced by Lactobacillus
plantarum 16. Microbiology 2013a; 159: 2014-2024.
Crowley S, Mahony J, van Sinderen D. Broad-spectrum antifungal-producing lactic acid bacteria
and their application in fruit models. Folia Microbiol (Praha) 2013b; 58: 291-299.
Crowley S, Mahony J, van Sinderen D. Comparative analysis of two antifungal Lactobacillus
plantarum isolates and their application as bioprotectants in refrigerated foods. J Appl
Microbiol 2012; 113: 1417-1427.
Cugini C, Calfee MW, Farrow JM, 3rd, Morales DK, Pesci EC, Hogan DA. Farnesol, a common
sesquiterpene, inhibits PQS production in Pseudomonas aeruginosa. Mol Microbiol 2007; 65:
896-906.
Cui L, Morris A, Ghedin E. The human mycobiome in health and disease. Genome Med 2013; 5:
63.
Dam G, Sorensen M, Munk OL, Keiding S. Hepatic ethanol elimination kinetics in patients with
cirrhosis. Scand J Gastroenterol 2009; 44: 867-871.
Darwazeh AM, Lamey PJ, Samaranayake LP, MacFarlane TW, Fisher BM, Macrury SM, MacCuish
AC. The relationship between colonisation, secretor status and in-vitro adhesion of Candida
albicans to buccal epithelial cells from diabetics. J Med Microbiol 1990; 33: 43-49.
De Bernardis F, Amacker M, Arancia S, Sandini S, Gremion C, Zurbriggen R, Moser C, Cassone A.
A virosomal vaccine against Candidal vaginitis: immunogenicity, efficacy and safety profile in
animal models. Vaccine 2012; 30: 4490-4498.
Dellarco VL. A mutagenicity assessment of acetaldehyde. Mutat Res 1988; 195: 1-20.
95
Delsing CE, Bleeker-Rovers CP, van de Veerdonk FL, Tol J, van der Meer JW, Kullberg BJ, Netea
MG. Association of esophageal candidiasis and squamous cell carcinoma. Med Mycol Case
Rep 2012; 1: 5-8.
Denning DW, Hope WW. Therapy for fungal diseases: opportunities and priorities. Trends Microbiol
2010; 18: 195-204.
Desai JV, Bruno VM, Ganguly S, Stamper RJ, Mitchell KF, Solis N, Hill EM, Xu W, Filler SG, Andes
DR, Fanning S, Lanni F, Mitchell AP. Regulatory role of glycerol in Candida albicans biofilm
formation. MBio 2013; 4: e00637-00612.
Dewhirst FE, Chen T, Izard J, Paster BJ, Tanner AC, Yu WH, Lakshmanan A, Wade WG. The
human oral microbiome. J Bacteriol 2010; 192: 5002-5017.
Dong YJ, Peng TK, Yin SJ. Expression and activities of class IV alcohol dehydrogenase and class
III aldehyde dehydrogenase in human mouth. Alcohol 1996; 13: 257-262.
Dongari-Bagtzoglou A, Kashleva H, Dwivedi P, Diaz P, Vasilakos J. Characterization of mucosal
Candida albicans biofilms. PLoS One 2009; 4: e7967.
Donlan RM, Costerton JW. Biofilms: survival mechanisms of clinically relevant microorganisms.
Clin Microbiol Rev 2002; 15: 167-193.
Drell T, Lillsaar T, Tummeleht L, Simm J, Aaspollu A, Vain E, Saarma I, Salumets A, Donders GG,
Metsis M. Characterization of the vaginal micro- and mycobiome in asymptomatic
reproductive-age Estonian women. PLoS One 2013; 8: e54379.
Duester G, Farres J, Felder MR, Holmes RS, Hoog JO, Pares X, Plapp BV, Yin SJ, Jornvall H.
Recommended nomenclature for the vertebrate alcohol dehydrogenase gene family. Biochem
Pharmacol 1999; 58: 389-395.
Dujon B. Yeast evolutionary genomics. Nat Rev Genet 2010; 11: 512-524.
Dupuy AK, David MS, Li L, Heider TN, Peterson JD, Montano EA, Dongari-Bagtzoglou A, Diaz PI,
Strausbaugh LD. Redefining the Human Oral Mycobiome with Improved Practices in
Amplicon-based Taxonomy: Discovery of Malassezia as a Prominent Commensal. PLoS One
2014; 9: e90899.
Duron V, Bii J, Mutai R, Ngetich J, Harrington D, Parker R, White R. Esophageal cancer awareness
in Bomet district, Kenya. Afr Health Sci 2013; 13: 122-128.
Dwivedi P, Thompson A, Xie Z, Kashleva H, Ganguly S, Mitchell AP, Dongari-Bagtzoglou A. Role
of Bcr1-activated genes Hwp1 and Hyr1 in Candida albicans oral mucosal biofilms and
neutrophil evasion. PLoS One 2011; 6: e16218.
Elahi S, Pang G, Ashman R, Clancy R. Enhanced clearance of Candida albicans from the oral
cavities of mice following oral administration of Lactobacillus acidophilus. Clin Exp Immunol
2005; 141: 29-36.
Ellahuene MF, Perez-Alzola LP, Olmedo MI. Chronic ethanol consumption in mice does not induce
DNA damage in somatic or germ cells, evaluated by the bone marrow micronucleous assay
and the dominant lethal mutation assay. Biol Res 2012; 45: 27-31.
96
Ene IV, Heilmann CJ, Sorgo AG, Walker LA, de Koster CG, Munro CA, Klis FM, Brown AJ. Carbon
source-induced reprogramming of the cell wall proteome and secretome modulates the
adherence and drug resistance of the fungal pathogen Candida albicans. Proteomics 2012;
12: 3164-3179.
Enzinger PC, Mayer RJ. Esophageal cancer. N Engl J Med 2003; 349: 2241-2252.
Epstein JB, Hancock PJ, Nantel S. Oral candidiasis in hematopoietic cell transplantation patients:
an outcome-based analysis. Oral Surg Oral Med Oral Pathol Oral Radiol Endod 2003; 96:
154-163.
Epstein JB, Truelove EL, Hanson-Huggins K, Mancl LA, Chen A, Press OW, Petersdorf SH,
Fritsche TR, Epstein JD. Topical polyene antifungals in hematopoietic cell transplant patients:
tolerability and efficacy. Support Care Cancer 2004; 12: 517-525.
Erlandsen SL, Kristich CJ, Dunny GM, Wells CL. High-resolution visualization of the microbial
glycocalyx with low-voltage scanning electron microscopy: dependence on cationic dyes. J
Histochem Cytochem 2004; 52: 1427-1435.
Ernst JF, Tielker D. Responses to hypoxia in fungal pathogens. Cell Microbiol 2009; 11: 183-190.
Eskan MA, Jotwani R, Abe T, Chmelar J, Lim JH, Liang S, Ciero PA, Krauss JL, Li F, Rauner M,
Hofbauer LC, Choi EY, Chung KJ, Hashim A, Curtis MA, Chavakis T, Hajishengallis G. The
leukocyte integrin antagonist Del-1 inhibits IL-17-mediated inflammatory bone loss. Nat
Immunol 2012; 13: 465-473.
Espina N, Lima V, Lieber CS, Garro AJ. In vitro and in vivo inhibitory effect of ethanol and
acetaldehyde on O6-methylguanine transferase. Carcinogenesis 1988; 9: 761-766.
Falsetta ML, Klein MI, Colonne PM, Scott-Anne K, Gregoire S, Pai CH, Gonzalez M, Watson G,
Krysan DJ, Bowen WH, Koo H. Symbiotic relationship between Streptococcus mutans and
Candida albicans synergizes the virulence of plaque-biofilms in vivo. Infect Immun 2014; 82:
1968-1981.
Fanello S, Bouchara JP, Sauteron M, Delbos V, Parot E, Marot-Leblond A, Moalic E, Le Flohicc
AM, Brangerd B. Predictive value of oral colonization by Candida yeasts for the onset of a
nosocomial infection in elderly hospitalized patients. J Med Microbiol 2006; 55: 223-228.
Fang JL, Vaca CE. Detection of DNA adducts of acetaldehyde in peripheral white blood cells of
alcohol abusers. Carcinogenesis 1997; 18: 627-632.
Fang JL, Vaca CE. Development of a 32P-postlabelling method for the analysis of adducts arising
through the reaction of acetaldehyde with 2'-deoxyguanosine-3'-monophosphate and DNA.
Carcinogenesis 1995; 16: 2177-2185.
Ferlay J, Steliarova-Foucher E, Lortet-Tieulent J, Rosso S, Coebergh JW, Comber H, Forman D,
Bray F. Cancer incidence and mortality patterns in Europe: estimates for 40 countries in 2012.
Eur J Cancer 2013; 49: 1374-1403.
Feron VJ, Kruysse A, Woutersen RA. Respiratory tract tumours in hamsters exposed to
acetaldehyde vapour alone or simultaneously to benzo(a)pyrene or diethylnitrosamine. Eur J
Cancer Clin Oncol 1982; 18: 13-31.
97
Feron VJ, Til HP, de Vrijer F, Woutersen RA, Cassee FR, van Bladeren PJ. Aldehydes:
occurrence, carcinogenic potential, mechanism of action and risk assessment. Mutat Res
1991; 259: 363-385.
Ferwerda G, Meyer-Wentrup F, Kullberg BJ, Netea MG, Adema GJ. Dectin-1 synergizes with TLR2
and TLR4 for cytokine production in human primary monocytes and macrophages. Cell
Microbiol 2008; 10: 2058-2066.
Fidel PL, Jr. Candida-host interactions in HIV disease: implications for oropharyngeal candidiasis.
Adv Dent Res 2011; 23: 45-49.
Findley K, Oh J, Yang J, Conlan S, Deming C, Meyer JA, Schoenfeld D, Nomicos E, Park M, Kong
HH, Segre JA. Topographic diversity of fungal and bacterial communities in human skin.
Nature 2013; 498: 367-370.
Finkel JS, Mitchell AP. Genetic control of Candida albicans biofilm development. Nat Rev Microbiol
2011; 9: 109-118.
Finnish Cancer Registry, Cancer Statistics 2013. www.cancerregistry.fi (accessed 25.4.2014
2014).
Fiori A, Kucharikova S, Govaert G, Cammue BP, Thevissen K, Van Dijck P. The heat-induced
molecular disaggregase Hsp104 of Candida albicans plays a role in biofilm formation and
pathogenicity in a worm infection model. Eukaryot Cell 2012; 11: 1012-1020.
Flores CL, Rodriguez C, Petit T, Gancedo C. Carbohydrate and energy-yielding metabolism in nonconventional yeasts. FEMS Microbiol Rev 2000; 24: 507-529.
Gadaga TH, Mutukumira AN, Narvhus JA. The growth and interaction of yeasts and lactic acid
bacteria isolated from Zimbabwean naturally fermented milk in UHT milk. Int J Food Microbiol
2001a; 68: 21-32.
Gadaga TH, Mutukumira AN, Narvhus JA. Growth characteristics of Candida kefyr and two strains
of Lactococcus lactis subsp lactis isolated from Zimbabwean naturally fermented milk. Int J
Food Microbiol 2001b; 70: 11-19.
Gagliardi D, Makihara S, Corsi PR, Viana Ade T, Wiczer MV, Nakakubo S, Mimica LM. Microbial
flora of the normal esophagus. Dis Esophagus 1998; 11: 248-250.
Gaissert HA, Breuer CK, Weissburg A, Mermel L. Surgical management of necrotizing Candida
esophagitis. Ann Thorac Surg 1999; 67: 231-233.
Gantner BN, Simmons RM, Canavera SJ, Akira S, Underhill DM. Collaborative induction of
inflammatory responses by dectin-1 and Toll-like receptor 2. J Exp Med 2003; 197: 11071117.
Garcha UK, Brummer E, Stevens DA. Synergy of fluconazole with human monocytes or monocytederived macrophages for killing of Candida species. J Infect Dis 1995; 172: 1620-1623.
Garcia-Vidal C, Viasus D, Carratala J. Pathogenesis of invasive fungal infections. Curr Opin Inf Dis
2013; 26: 270-276.
98
Garro AJ, Espina N, Farinati F, Salvagnini M. The effects of chronic ethanol consumption on
carcinogen metabolism and on O6-methylguanine transferase-mediated repair of alkylated
DNA. Alcohol Clin Exp Res 1986; 10: 73S-77S.
Ghannoum MA, Jurevic RJ, Mukherjee PK, Cui F, Sikaroodi M, Naqvi A, Gillevet PM.
Characterization of the oral fungal microbiome (mycobiome) in healthy individuals. PLoS
pathogens 2010; 6: e1000713.
Giovannucci E, Rimm EB, Ascherio A, Stampfer MJ, Colditz GA, Willett WC. Alcohol, lowmethionine--low-folate diets, and risk of colon cancer in men. J Natl Cancer Inst 1995; 87:
265-273.
Godoy W, Albano RM, Moraes EG, Pinho PR, Nunes RA, Saito EH, Higa C, Filho IM, Kruel CD,
Schirmer CC, Gurski R, Lang MA, Pinto LF. CYP2A6/2A7 and CYP2E1 expression in human
oesophageal mucosa: regional and inter-individual variation in expression and relevance to
nitrosamine metabolism. Carcinogenesis 2002; 23: 611-616.
Gow NA, Hube B. Importance of the Candida albicans cell wall during commensalism and infection.
Curr pin Microbiol 2012; 15: 406-412.
Grimoud AM, Lodter JP, Marty N, Andrieu S, Bocquet H, Linas MD, Rumeau M, Cazard JC.
Improved oral hygiene and Candida species colonization level in geriatric patients. Oral Dis
2005; 11: 163-169.
Gringhuis SI, Kaptein TM, Wevers BA, Theelen B, van der Vlist M, Boekhout T, Geijtenbeek TB.
Dectin-1 is an extracellular pathogen sensor for the induction and processing of IL-1beta via a
noncanonical caspase-8 inflammasome. Nat Immunol 2012; 13: 246-254.
Guha N, Boffetta P, Wunsch Filho V, Eluf Neto J, Shangina O, Zaridze D, Curado MP, Koifman S,
Matos E, Menezes A, Szeszenia-Dabrowska N, Fernandez L, Mates D, Daudt AW, Lissowska
J, Dikshit R, Brennan P. Oral health and risk of squamous cell carcinoma of the head and
neck and esophagus: results of two multicentric case-control studies. Am J Epidemiol 2007;
166: 1159-1173.
Guo J, Brosnan B, Furey A, Arendt E, Murphy P, Coffey A. Antifungal activity of Lactobacillus
against Microsporum canis, Microsporum gypseum and Epidermophyton floccosum. Bioeng
Bugs 2012; 3: 104-113.
Guo YM, Wang Q, Liu YZ, Chen HM, Qi Z, Guo QH. Genetic polymorphisms in cytochrome
P4502E1, alcohol and aldehyde dehydrogenases and the risk of esophageal squamous cell
carcinoma in Gansu Chinese males. World J Gastroenterol 2008; 14: 1444-1449.
Guzel-Seydim Z, Seydim AC, Greene AK. Organic acids and volatile flavor components evolved
during refrigerated storage of kefir. Journal of Dairy Science 2000; 83: 275-277.
Hajishengallis G. Immunomicrobial pathogenesis of periodontitis: keystones, pathobionts, and host
response. Trends Immunol 2014; 35: 3-11.
Halsted CH, Villanueva JA, Devlin AM, Chandler CJ. Metabolic interactions of alcohol and folate. J
Nutr 2002; 132: 2367S-2372S.
Hamza OJ, Matee MI, Bruggemann RJ, Moshi MJ, Simon EN, Mugusi F, Mikx FH, van der Lee HA,
Verweij PE, van der Ven AJ. Single-dose fluconazole versus standard 2-week therapy for
99
oropharyngeal candidiasis in HIV-infected patients: a randomized, double-blind, doubledummy trial. Clin Infect Dis 2008; 47: 1270-1276.
Harriott MM, Lilly EA, Rodriguez TE, Fidel PL, Jr., Noverr MC. Candida albicans forms biofilms on
the vaginal mucosa. Microbiology 2010; 156: 3635-3644.
Harriott MM, Noverr MC. Candida albicans and Staphylococcus aureus form polymicrobial biofilms:
effects on antimicrobial resistance. Antimicrob Agents Chemother 2009; 53: 3914-3922.
Harrison BD, Hashemi J, Bibi M, Pulver R, Bavli D, Nahmias Y, Wellington M, Sapiro G, Berman J.
A tetraploid intermediate precedes aneuploid formation in yeasts exposed to fluconazole.
PLoS Biol 2014; 12: e1001815.
Hashibe M, Boffetta P, Zaridze D, Shangina O, Szeszenia-Dabrowska N, Mates D, Janout V,
Fabianova E, Bencko V, Moullan N, Chabrier A, Hung R, Hall J, Canzian F, Brennan P.
Evidence for an important role of alcohol- and aldehyde-metabolizing genes in cancers of the
upper aerodigestive tract. Cancer Epidemiol Biomarkers Prev 2006; 15: 696-703.
Hashibe M, McKay JD, Curado MP, Oliveira JC, Koifman S, Koifman R, Zaridze D, Shangina O,
Wunsch-Filho V, Eluf-Neto J, Levi JE, Matos E, Lagiou P, Lagiou A, Benhamou S, Bouchardy
C, Szeszenia-Dabrowska N, Menezes A, Dall'Agnol MM, Merletti F, Richiardi L, Fernandez L,
Lence J, Talamini R, Barzan L, Mates D, Mates IN, Kjaerheim K, Macfarlane GJ, Macfarlane
TV, Simonato L, Canova C, Holcatova I, Agudo A, Castellsague X, Lowry R, Janout V,
Kollarova H, Conway DI, McKinney PA, Znaor A, Fabianova E, Bencko V, Lissowska J,
Chabrier A, Hung RJ, Gaborieau V, Boffetta P, Brennan P. Multiple ADH genes are associated
with upper aerodigestive cancers. Nat Genet 2008; 40: 707-709.
Hatakka K, Ahola AJ, Yli-Knuuttila H, Richardson M, Poussa T, Meurman JH, Korpela R. Probiotics
reduce the prevalence of oral Candida in the elderly--a randomized controlled trial. J Dent Res
2007; 86: 125-130.
Hernaez ML, Ximenez-Embun P, Martinez-Gomariz M, Gutierrez-Blazquez MD, Nombela C, Gil C.
Identification of Candida albicans exposed surface proteins in vivo by a rapid proteomic
approach. J Proteomics 2010; 73: 1404-1409.
Hickman MA, Zeng G, Forche A, Hirakawa MP, Abbey D, Harrison BD, Wang YM, Su CH, Bennett
RJ, Wang Y, Berman J. The 'obligate diploid' Candida albicans forms mating-competent
haploids. Nature 2013; 494: 55-59.
Hietala PK, Westermarck HW, Jaarma M. Identification of antimicrobial alpha-hydroxyacids in
Lactobacillus plantarum-fermented animal protein. Nutrition and metabolism 1979; 23: 227234.
Hisajima T, Maruyama N, Tanabe Y, Ishibashi H, Yamada T, Makimura K, Nishiyama Y, Funakoshi
K, Oshima H, Abe S. Protective effects of farnesol against oral candidiasis in mice. Microbiol
Immunol 2008; 52: 327-333.
Ho AW, Shen F, Conti HR, Patel N, Childs EE, Peterson AC, Hernandez-Santos N, Kolls JK, Kane
LP, Ouyang W, Gaffen SL. IL-17RC is required for immune signaling via an extended SEF/IL17R signaling domain in the cytoplasmic tail. J Immunol 2010; 185: 1063-1070.
Hoffmann D, Hoffmann I, El-Bayoumy K. The less harmful cigarette: a controversial issue. a tribute
to Ernst L. Wynder. Chem Res Toxicol 2001; 14: 767-790.
100
Holmes RS. Alcohol dehydrogenases: a family of isozymes with differential functions. Alcohol
Alcohol Suppl 1994; 2: 127-130.
Homann N, Jousimies-Somer H, Jokelainen K, Heine R, Salaspuro M. High acetaldehyde levels in
saliva after ethanol consumption: methodological aspects and pathogenetic implications.
Carcinogenesis 1997; 18: 1739-1743.
Homann N, Tillonen J, Rintamaki H, Salaspuro M, Lindqvist C, Meurman JH. Poor dental status
increases acetaldehyde production from ethanol in saliva: a possible link to increased oral
cancer risk among heavy drinkers. Oral Oncol 2001; 37: 153-158.
Hooper SJ, Wilson MJ, Crean SJ. Exploring the link between microorganisms and oral cancer: a
systematic review of the literature. Head Neck 2009; 31: 1228-1239.
Hornby JM, Jensen EC, Lisec AD, Tasto JJ, Jahnke B, Shoemaker R, Dussault P, Nickerson KW.
Quorum sensing in the dimorphic fungus Candida albicans is mediated by farnesol. Appl
Environ Microbiol 2001; 67: 2982-2992.
Hoyer LL, Green CB, Oh SH, Zhao X. Discovering the secrets of the Candida albicans agglutininlike sequence (ALS) gene family--a sticky pursuit. Med Mycol 2008; 46: 1-15.
Hsia CC, Sun TT, Wang YY, Anderson LM, Armstrong D, Good RA. Enhancement of formation of
the esophageal carcinogen benzylmethylnitrosamine from its precursors by Candida albicans.
Proc Natl Acad Sci U S A 1981; 78: 1878-1881.
Huang G, Wang H, Chou S, Nie X, Chen J, Liu H. Bistable expression of WOR1, a master
regulator of white-opaque switching in Candida albicans. Proc Natl Acad Sci U S A 2006; 103:
12813-12818.
Huffnagle GB, Noverr MC. The emerging world of the fungal microbiome. Trends Microbiol 2013;
21: 334-341.
Human Microbiome Project. Structure, function and diversity of the healthy human microbiome.
Nature 2012; 486: 207-214.
Huppler AR, Conti HR, Hernandez-Santos N, Darville T, Biswas PS, Gaffen SL. Role of neutrophils
in IL-17-dependent immunity to mucosal candidiasis. J Immunol 2014; 192: 1745-1752.
Hutson RA, Thompson DE, Collins MD. Genetic interrelationships of saccharolytic Clostridium
botulinum types B, E and F and related clostridia as revealed by small-subunit rRNA gene
sequences. FEMS Microbiol Lett 1993; 108: 103-110.
Ifrim DC, Quintin J, Joosten LA, Jacobs C, Jansen T, Jacobs L, Gow NA, Williams DL, van der
Meer JW, Netea MG. Trained immunity or tolerance: opposing functional programs induced in
human monocytes after engagement of various pattern recognition receptors. Clin Vaccine
Immunol 2014; 21: 534-545.
Ihmels J, Bergmann S, Gerami-Nejad M, Yanai I, McClellan M, Berman J, Barkai N. Rewiring of the
yeast transcriptional network through the evolution of motif usage. Science 2005; 309: 938940.
101
Inglis DO, Arnaud MB, Binkley J, Shah P, Skrzypek MS, Wymore F, Binkley G, Miyasato SR,
Simison M, Sherlock G. The Candida genome database incorporates multiple Candida
species: multispecies search and analysis tools with curated gene and protein information for
Candida albicans and Candida glabrata. Nucleic Acids Res 2012; 40: D667-674.
Isham N, Ghannoum MA. Antifungal activity of miconazole against recent Candida strains.
Mycoses 2010; 53: 434-437.
Iwata C, Nakagaki H, Morita I, Sekiya T, Goshima M, Abe T, Isogai A, Hanaki M, Kuwahara M,
Tatematsu M, Robinson C. Daily use of dentifrice with and without xylitol and fluoride: effect
on glucose retention in humans in vivo. Arch Oral Biol 2003; 48: 389-395.
Jakobsson HE, Jernberg C, Andersson AF, Sjölund-Karlsson M, Jansson JK, Engstrand L. Shortterm antibiotic treatment has differing long-term impacts on the human throat and gut
microbiome. PLoS One 2010; 5: e9836.
Jelski W, Szmitkowski M. Alcohol dehydrogenase (ADH) and aldehyde dehydrogenase (ALDH) in
the cancer diseases. Clin Chim Acta 2008; 395: 1-5.
Jenkinson HF, Lamont RJ. Oral microbial communities in sickness and in health. Trends Microbiol
2005; 13: 589-595.
Jokelainen K, Heikkonen E, Roine R, Lehtonen H, Salaspuro M. Increased acetaldehyde
production by mouthwashings from patients with oral cavity, laryngeal, or pharyngeal cancer.
Alcohol Clin Exp Res 1996; 20: 1206-1210.
Jones AW. Distribution of Ethanol between Saliva and Blood in Man. Clinical and Experimental
Pharmacology and Physiology 1979; 6: 53-59.
Jones SE, Versalovic J. Probiotic Lactobacillus reuteri biofilms produce antimicrobial and antiinflammatory factors. BMC Microbiol 2009; 9: 35.
Jorth P, Turner KH, Gumus P, Nizam N, Buduneli N, Whiteley M. Metatranscriptomics of the
human oral microbiome during health and disease. MBio 2014; 5: e01012-01014.
Julkunen RJ, Tannenbaum L, Baraona E, Lieber CS. First pass metabolism of ethanol: an
important determinant of blood levels after alcohol consumption. Alcohol 1985; 2: 437-441.
Järvensivu A, Hietanen J, Rautemaa R, Sorsa T, Richardson M. Candida yeasts in chronic
periodontitis tissues and subgingival microbial biofilms in vivo. Oral Dis 2004; 10: 106-112.
Kamagata-Kiyoura Y, Abe S, Yamaguchi H, Nitta T. Protective effects of human saliva on
experimental murine oral candidiasis. J Infect Chemother 2004; 10: 253-255.
Kanzaki R, Yano M, Takachi K, Ishiguro S, Motoori M, Kishi K, Miyashiro I, Ishikawa O, Imaoka S.
Candida esophagitis complicated by an esophago-airway fistula: report of a case. Surg Today
2009; 39: 972-978.
Kaneko Y, Miyagawa S, Takeda O, Hakariya M, Matsumoto S, Ohno H, Miyazaki Y. Real-time
microscopic observation of Candida biofilm development and effects due to micafungin and
fluconazole. Antimicrob Agents Chemother 2013; 57: 2226-2230.
102
Kaplan JE, Benson C, Holmes KK, Brooks JT, Pau A, Masur H. Guidelines for prevention and
treatment of opportunistic infections in HIV-infected adults and adolescents: recommendations
from CDC, the National Institutes of Health, and the HIV Medicine Association of the Infectious
Diseases Society of America. MMWR Recomm Rep 2009; 58: 1-207; quiz CE201-204.
Khan MS, Ahmad I. Antibiofilm activity of certain phytocompounds and their synergy with
fluconazole against Candida albicans biofilms. J Antimicrob Chemother 2012; 67: 618-621.
Kiehne K, Brunke G, Meyer D, Harder J, Herzig KH. Oesophageal defensin expression during
Candida infection and reflux disease. Scand J Gastroenterol 2005; 40: 501-507.
Klyosov AA, Rashkovetsky LG, Tahir MK, Keung WM. Possible role of liver cytosolic and
mitochondrial aldehyde dehydrogenases in acetaldehyde metabolism. Biochemistry 1996; 35:
4445-4456.
Kohler GA, Assefa S, Reid G. Probiotic interference of Lactobacillus rhamnosus GR-1 and
Lactobacillus reuteri RC-14 with the opportunistic fungal pathogen Candida albicans. Infect
Dis Obstet Gynecol 2012; 2012: 636474.
Kojic EM, Darouiche RO. Candida infections of medical devices. Clin Microbiol Rev 2004; 17: 255267.
Kok-Tas T, Seydim AC, Ozer B, Guzel-Seydim ZB. Effects of different fermentation parameters on
quality characteristics of kefir. J Dairy Sci 2013; 96: 780-789.
Koop DR. Oxidative and reductive metabolism by cytochrome P450 2E1. FASEB J 1992; 6: 724730.
Kotter P, Amore R, Hollenberg CP, Ciriacy M. Isolation and Characterization of the Pichia stipitis
Xylitol Dehydrogenase Gene, Xyl2, and Construction of a Xylose-Utilizing Saccharomyces
cerevisiae Transformant. Current Genetics 1990; 18: 493-500.
Krcmery V, Barnes AJ. Non-albicans Candida spp. causing fungaemia: pathogenicity and
antifungal resistance. J Hosp Infect 2002; 50: 243-260.
Krogh P. The role of yeasts in oral cancer by means of endogenous nitrosation. Acta Odontol
Scand 1990; 48: 85-88.
Kucharikova S, Tournu H, Lagrou K, Van Dijck P, Bujdakova H. Detailed comparison of Candida
albicans and Candida glabrata biofilms under different conditions and their susceptibility to
caspofungin and anidulafungin. J Med Microbiol 2011; 60: 1261-1269.
Kuhn DM, Chandra J, Mukherjee PK, Ghannoum MA. Comparison of biofilms formed by Candida
albicans and Candida parapsilosis on bioprosthetic surfaces. Infect Immun 2002; 70: 878-888.
Kulkarni R, Antala S, Wang A, Amaral FE, Rampersaud R, Larussa SJ, Planet PJ, Ratner AJ.
Cigarette smoke increases Staphylococcus aureus biofilm formation via oxidative stress. Infect
Immun 2012; 80: 3804-3811.
Kullberg BJ, Netea MG, Vonk AG, van der Meer JW. Modulation of neutrophil function in host
defense against disseminated Candida albicans infection in mice. FEMS Immunol Med
Microbiol 1999; 26: 299-307.
103
Kumar H, Kumagai Y, Tsuchida T, Koenig PA, Satoh T, Guo Z, Jang MH, Saitoh T, Akira S, Kawai
T. Involvement of the NLRP3 inflammasome in innate and humoral adaptive immune
responses to fungal beta-glucan. J Immunol 2009; 183: 8061-8067.
Kurkivuori J, Salaspuro V, Kaihovaara P, Kari K, Rautemaa R, Grönroos L, Meurman JH,
Salaspuro M. Acetaldehyde production from ethanol by oral streptococci. Oral Oncol 2007; 43:
181-186.
Kurtzman CP, Robnett CJ. Identification of clinically important ascomycetous yeasts based on
nucleotide divergence in the 5' end of the large-subunit (26S) ribosomal DNA gene. J Clin
Microbiol 1997; 35: 1216-1223.
Kusch H, Engelmann S, Bode R, Albrecht D, Morschhauser J, Hecker M. A proteomic view of
Candida albicans yeast cell metabolism in exponential and stationary growth phases. Int J
Med Microbiol 2008; 298: 291-318.
Kuula H, Salo T, Pirilä E, Tuomainen AM, Jauhiainen M, Uitto VJ, Tjäderhane L, Pussinen PJ,
Sorsa T. Local and systemic responses in matrix metalloproteinase 8-deficient mice during
Porphyromonas gingivalis-induced periodontitis. Infect Immun 2009; 77: 850-859.
Lachenmeier DW, Kanteres F, Rehm J. Carcinogenicity of acetaldehyde in alcoholic beverages:
risk assessment outside ethanol metabolism. Addiction 2009; 104: 533-550.
Lachenmeier DW, Monakhova YB. Short-term salivary acetaldehyde increase due to direct
exposure to alcoholic beverages as an additional cancer risk factor beyond ethanol
metabolism. J Exp Clin Cancer Res 2011; 30: 3.
Lachenmeier DW, Sohnius EM. The role of acetaldehyde outside ethanol metabolism in the
carcinogenicity of alcoholic beverages: evidence from a large chemical survey. Food Chem
Toxicol 2008; 46: 2903-2911.
Lamaris GA, Lewis RE, Chamilos G, May GS, Safdar A, Walsh TJ, Raad, II, Kontoyiannis DP.
Caspofungin-mediated
beta-glucan
unmasking
and
enhancement
of
human
polymorphonuclear neutrophil activity against Aspergillus and non-Aspergillus hyphae. J Infect
Dis 2008; 198: 186-192.
Laudenbach JM, Epstein JB. Treatment strategies for oropharyngeal candidiasis. Expert Opin
Pharmacother 2009; 10: 1413-1421.
Le Chatelier E, Nielsen T, Qin J, Prifti E, Hildebrand F, Falony G, Almeida M, Arumugam M, Batto
JM, Kennedy S, Leonard P, Li J, Burgdorf K, Grarup N, Jorgensen T, Brandslund I, Nielsen
HB, Juncker AS, Bertalan M, Levenez F, Pons N, Rasmussen S, Sunagawa S, Tap J, Tims S,
Zoetendal EG, Brunak S, Clement K, Dore J, Kleerebezem M, Kristiansen K, Renault P,
Sicheritz-Ponten T, de Vos WM, Zucker JD, Raes J, Hansen T, Bork P, Wang J, Ehrlich SD,
Pedersen O. Richness of human gut microbiome correlates with metabolic markers. Nature
2013; 500: 541-546.
Lechevrel M, Casson AG, Wolf CR, Hardie LJ, Flinterman MB, Montesano R, Wild CP.
Characterization of cytochrome P450 expression in human oesophageal mucosa.
Carcinogenesis 1999; 20: 243-248.
Lee CH, Wu DC, Lee JM, Wu IC, Goan YG, Kao EL, Huang HL, Chan TF, Chou SH, Chou YP, Lee
CY, Chen PS, Ho CK, He J, Wu MT. Carcinogenetic impact of alcohol intake on squamous
104
cell carcinoma risk of the oesophagus in relation to tobacco smoking. Eur J Cancer 2007; 43:
1188-1199.
Lee YE, Choi YH, Jeong SH, Kim HS, Lee SH, Song KB. Morphological changes in Streptococcus
mutans after chewing gum containing xylitol for twelve months. Curr Microbiol 2009; 58: 332337.
Li M, Chen Q, Tang R, Shen Y, Liu WD. The expression of beta-defensin-2, 3 and LL-37 induced
by Candida albicans phospholipomannan in human keratinocytes. J Dermatol Sci 2011; 61:
72-75.
Li XS, Sun JN, Okamoto-Shibayama K, Edgerton M. Candida albicans cell wall ssa proteins bind
and facilitate import of salivary histatin 5 required for toxicity. J Biol Chem 2006; 281: 2245322463.
Lieber CS. Cytochrome P-4502E1: its physiological and pathological role. Physiol Rev 1997; 77:
517-544.
Lieschke GJ, Burgess AW. Granulocyte colony-stimulating factor and granulocyte-macrophage
colony-stimulating factor (1). N Engl J Med 1992; 327: 28-35.
Lin SC, Greenberg DM. Enzymatic breakdown of threonine by threonine aldolase. J Gen Physiol
1954; 38: 181-196.
Linderborg K, Joly JP, Visapää JP, Salaspuro M. Potential mechanism for Calvados-related
oesophageal cancer. Food Chem Toxicol 2008; 46: 476-479.
Linderborg K, Salaspuro M, Väkeväinen S. A single sip of a strong alcoholic beverage causes
exposure to carcinogenic concentrations of acetaldehyde in the oral cavity. Food Chem
Toxicol 2011; 49: 2103-2106.
Lionakis MS, Netea MG. Candida and host determinants of susceptibility to invasive candidiasis.
PLoS Pathog 2013; 9: e1003079.
Liu C, Russell RM, Seitz HK, Wang XD. Ethanol enhances retinoic acid metabolism into polar
metabolites in rat liver via induction of cytochrome P4502E1. Gastroenterology 2001a; 120:
179-189.
Liu S, Park JY, Schantz SP, Stern JC, Lazarus P. Elucidation of CYP2E1 5' regulatory RsaI/Pstl
allelic variants and their role in risk for oral cancer. Oral Oncol 2001b; 37: 437-445.
Liu X, Hua H. Oral manifestation of chronic mucocutaneous candidiasis: seven case reports. J Oral
Pathol Med 2007; 36: 528-532.
Lodi G, Tarozzi M, Sardella A, Demarosi F, Canegallo L, Di Benedetto D, Carrassi A. Miconazole
as adjuvant therapy for oral lichen planus: a double-blind randomized controlled trial. Br J
Dermatol 2007; 156: 1336-1341.
Maccallum DM. Hosting infection: experimental models to assay Candida virulence. Int J Microbiol
2012; 2012: 363764.
Macfarlane S, Bahrami B, Macfarlane GT. Mucosal biofilm communities in the human intestinal
tract. Adv Appl Microbiol 2011; 75: 111-143.
105
Maconi E, Griffini A, Cavazzoni V, Aragozzini F. Reduction of acetaldehyde to ethanol by some
micro-organisms and its stereospecificity. Biochem J 1988; 250: 929-932.
Maestre O, Garcia-Martinez T, Peinado RA, Mauricio JC. Effects of ADH2 overexpression in
Saccharomyces bayanus during alcoholic fermentation. Appl Environ Microbiol 2008; 74: 702707.
Maffei F, Fimognari C, Castelli E, Stefanini GF, Forti GC, Hrelia P. Increased cytogenetic damage
detected by FISH analysis on micronuclei in peripheral lymphocytes from alcoholics.
Mutagenesis 2000; 15: 517-523.
Maffei F, Forti GC, Castelli E, Stefanini GF, Mattioli S, Hrelia P. Biomarkers to assess the genetic
damage induced by alcohol abuse in human lymphocytes. Mutat Res 2002; 514: 49-58.
Mantri SS, Parkhedkar RD, Mantri SP. Candida colonisation and the efficacy of chlorhexidine
gluconate on soft silicone-lined dentures of diabetic and non-diabetic patients. Gerodontology
2013; 30: 288-295.
Manzoni P, Mostert M, Leonessa ML, Priolo C, Farina D, Monetti C, Latino MA, Gomirato G. Oral
supplementation with Lactobacillus casei subspecies rhamnosus prevents enteric colonization
by Candida species in preterm neonates: a randomized study. Clin Infect Dis 2006; 42: 17351742.
Marcotte H, Lavoie MC. Oral microbial ecology and the role of salivary immunoglobulin A. Microbiol
Mol Biol Rev 1998; 62: 71-109.
Marsh PD. Dental plaque as a biofilm and a microbial community - implications for health and
disease. BMC Oral Health 2006; 6: Suppl 1:S14.
Marsh PD. Dental plaque as a microbial biofilm. Caries Res 2004; 38: 204-211.
Marstorp H, Witter E. Extractable dsDNA and product formation as measures of microbial growth in
soil upon substrate addition. Soil Biol Biochem 1999; 31: 1443-1453.
Martinez-Gomariz M, Perumal P, Mekala S, Nombela C, Chaffin WL, Gil C. Proteomic analysis of
cytoplasmic and surface proteins from yeast cells, hyphae, and biofilms of Candida albicans.
Proteomics 2009; 9: 2230-2252.
Martins M, Uppuluri P, Thomas DP, Cleary IA, Henriques M, Lopez-Ribot JL, Oliveira R. Presence
of extracellular DNA in the Candida albicans biofilm matrix and its contribution to biofilms.
Mycopathologia 2010; 169: 323-331.
Marttila E, Bowyer P, Sanglard D, Uittamo J, Kaihovaara P, Salaspuro M, Richardson M,
Rautemaa R. Fermentative 2-carbon metabolism produces carcinogenic levels of
acetaldehyde in Candida albicans. Mol Oral Microbiol 2013a; 28: 281-291.
Marttila E, Uittamo J, Rusanen P, Lindqvist C, Salaspuro M, Rautemaa R. Acetaldehyde production
and microbial colonization in oral squamous cell carcinoma and oral lichenoid disease. Oral
Surg Oral Med Oral Pathol Oral Radiol 2013b; 116: 61-68.
106
Matsuda T, Kawanishi M, Yagi T, Matsui S, Takebe H. Specific tandem GG to TT base
substitutions induced by acetaldehyde are due to intra-strand crosslinks between adjacent
guanine bases. Nucleic Acids Res 1998; 26: 1769-1774.
McCullough M, Jaber M, Barrett AW, Bain L, Speight PM, Porter SR. Oral yeast carriage correlates
with presence of oral epithelial dysplasia. Oral Oncol 2002; 38: 391-393.
McKay JD, Truong T, Gaborieau V, Chabrier A, Chuang SC, Byrnes G, Zaridze D, Shangina O,
Szeszenia-Dabrowska N, Lissowska J, Rudnai P, Fabianova E, Bucur A, Bencko V, Holcatova
I, Janout V, Foretova L, Lagiou P, Trichopoulos D, Benhamou S, Bouchardy C, Ahrens W,
Merletti F, Richiardi L, Talamini R, Barzan L, Kjaerheim K, Macfarlane GJ, Macfarlane TV,
Simonato L, Canova C, Agudo A, Castellsague X, Lowry R, Conway DI, McKinney PA, Healy
CM, Toner ME, Znaor A, Curado MP, Koifman S, Menezes A, Wunsch-Filho V, Neto JE,
Garrote LF, Boccia S, Cadoni G, Arzani D, Olshan AF, Weissler MC, Funkhouser WK, Luo J,
Lubinski J, Trubicka J, Lener M, Oszutowska D, Schwartz SM, Chen C, Fish S, Doody DR,
Muscat JE, Lazarus P, Gallagher CJ, Chang SC, Zhang ZF, Wei Q, Sturgis EM, Wang LE,
Franceschi S, Herrero R, Kelsey KT, McClean MD, Marsit CJ, Nelson HH, Romkes M, Buch S,
Nukui T, Zhong S, Lacko M, Manni JJ, Peters WH, Hung RJ, McLaughlin J, Vatten L, Njolstad
I, Goodman GE, Field JK, Liloglou T, Vineis P, Clavel-Chapelon F, Palli D, Tumino R, Krogh
V, Panico S, Gonzalez CA, Quiros JR, Martinez C, Navarro C, Ardanaz E, Larranaga N, Khaw
KT, Key T, Bueno-de-Mesquita HB, Peeters PH, Trichopoulou A, Linseisen J, Boeing H,
Hallmans G, Overvad K, Tjonneland A, Kumle M, Riboli E, Valk K, Vooder T, Metspalu A,
Zelenika D, Boland A, Delepine M, Foglio M, Lechner D, Blanche H, Gut IG, Galan P, Heath
S, Hashibe M, Hayes RB, Boffetta P, Lathrop M, Brennan P. A genome-wide association study
of upper aerodigestive tract cancers conducted within the INHANCE consortium. PLoS Genet
2011; 7: e1001333.
Mero AA, Ojala T, Hulmi JJ, Puurtinen R, Karila TA, Seppälä T. Effects of alfa-hydroxy-isocaproic
acid on body composition, DOMS and performance in athletes. J Int Soc Sports Nutr 2010; 7:
1.
Metchnikoff E. The prolongation of life: optimistic studies. English transl. ed. by Mitchell PC. New
York: GP Putnam's Sons; 1908.
Monteoliva L, Martinez-Lopez R, Pitarch A, Hernaez ML, Serna A, Nombela C, Albar JP, Gil C.
Quantitative proteome and acidic subproteome profiling of Candida albicans yeast-to-hypha
transition. J Proteome Res 2011; 10: 502-517.
Morales DK, Grahl N, Okegbe C, Dietrich LE, Jacobs NJ, Hogan DA. Control of Candida albicans
metabolism and biofilm formation by Pseudomonas aeruginosa phenazines. MBio 2013; 4:
e00526-00512.
Morgan J. Global trends in candidemia: review of reports from 1995-2005. Curr Infect Dis Rep
2005; 7: 429-439.
Morris A, Beck JM, Schloss PD, Campbell TB, Crothers K, Curtis JL, Flores SC, Fontenot AP,
Ghedin E, Huang L, Jablonski K, Kleerup E, Lynch SV, Sodergren E, Twigg H, Young VB,
Bassis CM, Venkataraman A, Schmidt TM, Weinstock GM. Comparison of the respiratory
microbiome in healthy nonsmokers and smokers. Am J Respir Crit Care Med 2013; 187: 10671075.
Moyes DL, Runglall M, Murciano C, Shen C, Nayar D, Thavaraj S, Kohli A, Islam A, Mora-Montes
H, Challacombe SJ, Naglik JR. A biphasic innate immune MAPK response discriminates
107
between the yeast and hyphal forms of Candida albicans in epithelial cells. Cell Host Microbe
2010; 8: 225-235.
Mugula JK, Narvhus JA, Sorhaug T. Use of starter cultures of lactic acid bacteria and yeasts in the
preparation of togwa, a Tanzanian fermented food. Int J Food Microbiol 2003; 83: 307-318.
Mukherjee PK, Mohamed S, Chandra J, Kuhn D, Liu S, Antar OS, Munyon R, Mitchell AP, Andes
D, Chance MR, Rouabhia M, Ghannoum MA. Alcohol dehydrogenase restricts the ability of
the pathogen Candida albicans to form a biofilm on catheter surfaces through an ethanolbased mechanism. Infect Immun 2006; 74: 3804-3816.
Muto M, Hitomi Y, Ohtsu A, Shimada H, Kashiwase Y, Sasaki H, Yoshida S, Esumi H.
Acetaldehyde production by non-pathogenic Neisseria in human oral microflora: implications
for carcinogenesis in upper aerodigestive tract. Int J Cancer 2000; 88: 342-350.
Naglik JR, Challacombe SJ, Hube B. Candida albicans secreted aspartyl proteinases in virulence
and pathogenesis. Microbiol Mol Biol Rev 2003; 67: 400-428.
Naglik JR, Moyes DL, Wachtler B, Hube B. Candida albicans interactions with epithelial cells and
mucosal immunity. Microbes Infect 2011; 13: 963-976.
Nagy KN, Sonkodi I, Szoke I, Nagy E, Newman HN. The microflora associated with human oral
carcinomas. Oral Oncol 1998; 34: 304-308.
Nailis H, Coenye T, Van Nieuwerburgh F, Deforce D, Nelis HJ. Development and evaluation of
different normalization strategies for gene expression studies in Candida albicans biofilms by
real-time PCR. BMC Mol Biol 2006; 7: 25.
Närhi TO, Ainamo A, Meurman JH. Salivary yeasts, saliva, and oral mucosa in the elderly. J Dent
Res 1993; 72: 1009-1014.
Narikiyo M, Tanabe C, Yamada Y, Igaki H, Tachimori Y, Kato H, Muto M, Montesano R, Sakamoto
H, Nakajima Y, Sasaki H. Frequent and preferential infection of Treponema denticola,
Streptococcus mitis, and Streptococcus anginosus in esophageal cancers. Cancer Sci 2004;
95: 569-574.
Narvhus JA, Gadaga TH. The role of interaction between yeasts and lactic acid bacteria in African
fermented milks: a review. Int J Food Microbiol 2003; 86: 51-60.
Nasidze I, Li J, Quinque D, Tang K, Stoneking M. Global diversity in the human salivary
microbiome. Genome Res 2009; 19: 636-643.
Navarathna DH, Hornby JM, Krishnan N, Parkhurst A, Duhamel GE, Nickerson KW. Effect of
farnesol on a mouse model of systemic candidiasis, determined by use of a DPP3 knockout
mutant of Candida albicans. Infect Immun 2007; 75: 1609-1618.
Netea MG, Kullberg BJ, Boerman OC, Verschueren I, Dinarello CA, Van der Meer JW. Soluble
murine IL-1 receptor type I induces release of constitutive IL-1 alpha. J Immunol 1999; 162:
4876-4881.
Nett J, Lincoln L, Marchillo K, Massey R, Holoyda K, Hoff B, VanHandel M, Andes D. Putative role
of beta-1,3 glucans in Candida albicans biofilm resistance. Antimicrob Agents Chemother
2007; 51: 510-520.
108
Nett JE, Lepak AJ, Marchillo K, Andes DR. Time course global gene expression analysis of an in
vivo Candida biofilm. J Infect Dis 2009; 200: 307-313.
Nett JE, Marchillo K, Spiegel CA, Andes DR. Development and validation of an in vivo Candida
albicans biofilm denture model. Infect Immun 2010; 78: 3650-3659.
Newman SL, Holly A. Candida albicans is phagocytosed, killed, and processed for antigen
presentation by human dendritic cells. Infect Immun 2001; 69: 6813-6822.
Niemelä O, Parkkila S, Pasanen M, Viitala K, Villanueva JA, Halsted CH. Induction of cytochrome
P450 enzymes and generation of protein-aldehyde adducts are associated with sexdependent sensitivity to alcohol-induced liver disease in micropigs. Hepatology 1999; 30:
1011-1017.
Niemelä O, Parkkila S, Yla-Herttuala S, Villanueva J, Ruebner B, Halsted CH. Sequential
acetaldehyde production, lipid peroxidation, and fibrogenesis in micropig model of alcoholinduced liver disease. Hepatology 1995; 22: 1208-1214.
Nilsson RH, Ryberg M, Kristiansson E, Abarenkov K, Larsson KH, Koljalg U. Taxonomic reliability
of DNA sequences in public sequence databases: a fungal perspective. PLoS One 2006; 1:
e59.
Nishimura S, Nagata N, Shimbo T, Asayama N, Akiyama J, Ohmagari N, Yazaki H, Oka S, Uemura
N. Factors associated with esophageal candidiasis and its endoscopic severity in the era of
antiretroviral therapy. PLoS One 2013; 8: e58217.
Nkuize M, De Wit S, Muls V, Arvanitakis M, Buset M. Upper gastrointestinal endoscopic findings in
the era of highly active antiretroviral therapy. HIV Med 2010; 11: 412-417.
Nobile CJ, Fox EP, Nett JE, Sorrells TR, Mitrovich QM, Hernday AD, Tuch BB, Andes DR, Johnson
AD. A recently evolved transcriptional network controls biofilm development in Candida
albicans. Cell 2012; 148: 126-138.
Nobile CJ, Nett JE, Hernday AD, Homann OR, Deneault JS, Nantel A, Andes DR, Johnson AD,
Mitchell AP. Biofilm matrix regulation by Candida albicans Zap1. PLoS Biol 2009; 7:
e1000133.
Nobile CJ, Schneider HA, Nett JE, Sheppard DC, Filler SG, Andes DR, Mitchell AP.
Complementary adhesin function in C. albicans biofilm formation. Curr Biol 2008; 18: 10171024.
Norder Grusell E, Dahlen G, Ruth M, Ny L, Quiding-Jarbrink M, Bergquist H, Bove M. Bacterial
flora of the human oral cavity, and the upper and lower esophagus. Dis Esophagus 2013; 26:
84-90.
Nosova T, Jokelainen K, Kaihovaara P, Heine R, Jousimies-Somer H, Salaspuro M. Characteristics
of aldehyde dehydrogenases of certain aerobic bacteria representing human colonic flora.
Alcohol Alcohol 1998; 33: 273-280.
Nosova T, Jokelainen K, Kaihovaara P, Jousimies-Somer H, Siitonen A, Heine R, Salaspuro M.
Aldehyde dehydrogenase activity and acetate production by aerobic bacteria representing the
normal flora of human large intestine. Alcohol Alcohol 1996; 31: 555-564.
109
Nosova T, Jousimies-Somer H, Jokelainen K, Heine R, Salaspuro M. Acetaldehyde production and
metabolism by human indigenous and probiotic Lactobacillus and Bifidobacterium strains.
Alcohol Alcohol 2000; 35: 561-568.
Nucci M, Anaissie E. Revisiting the source of candidemia: skin or gut? Clin Infect Dis 2001; 33:
1959-1967.
O'Grady JF, Reade PC. Candida albicans as a promoter of oral mucosal neoplasia.
Carcinogenesis 1992; 13: 783-786.
Obe G, Jonas R, Schmidt S. Metabolism of ethanol in vitro produces a compound which induces
sister-chromatid exchanges in human peripheral lymphocytes in vitro: acetaldehyde not
ethanol is mutagenic. Mutat Res 1986; 174: 47-51.
Odds FC, De Backer T, Dams G, Vranckx L, Woestenborghs F. Oxygen as limiting nutrient for
growth of Cryptococcus neoformans. J Clin Microbiol 1995; 33: 995-997.
Ohsawa I, Nishimaki K, Yasuda C, Kamino K, Ohta S. Deficiency in a mitochondrial aldehyde
dehydrogenase increases vulnerability to oxidative stress in PC12 cells. J Neurochem 2003;
84: 1110-1117.
Okkers DJ, Dicks LM, Silvester M, Joubert JJ, Odendaal HJ. Characterization of pentocin TV35b, a
bacteriocin-like peptide isolated from Lactobacillus pentosus with a fungistatic effect on
Candida albicans. J Appl Microbiol 1999; 87: 726-734.
Oneta CM, Lieber CS, Li J, Ruttimann S, Schmid B, Lattmann J, Rosman AS, Seitz HK. Dynamics
of cytochrome P4502E1 activity in man: induction by ethanol and disappearance during
withdrawal phase. J Hepatol 2002; 36: 47-52.
Oshino N, Oshino R, Chance B. The characteristics of the "peroxidatic" reaction of catalase in
ethanol oxidation. Biochem J 1973; 131: 555-563.
Osier MV, Pakstis AJ, Soodyall H, Comas D, Goldman D, Odunsi A, Okonofua F, Parnas J, Schulz
LO, Bertranpetit J, Bonne-Tamir B, Lu RB, Kidd JR, Kidd KK. A global perspective on genetic
variation at the ADH genes reveals unusual patterns of linkage disequilibrium and diversity.
Am J Hum Genet 2002; 71: 84-99.
Oze I, Matsuo K, Suzuki T, Kawase T, Watanabe M, Hiraki A, Ito H, Hosono S, Ozawa T, Hatooka
S, Yatabe Y, Hasegawa Y, Shinoda M, Kiura K, Tajima K, Tanimoto M, Tanaka H. Impact of
multiple alcohol dehydrogenase gene polymorphisms on risk of upper aerodigestive tract
cancers in a Japanese population. Cancer Epidemiol Biomarkers Prev 2009; 18: 3097-3102.
Pagano L, Caira M, Candoni A, Offidani M, Fianchi L, Martino B, Pastore D, Picardi M, Bonini A,
Chierichini A, Fanci R, Caramatti C, Invernizzi R, Mattei D, Mitra ME, Melillo L, Aversa F, Van
Lint MT, Falcucci P, Valentini CG, Girmenia C, Nosari A. The epidemiology of fungal infections
in patients with hematologic malignancies: the SEIFEM-2004 study. Haematologica 2006; 91:
1068-1075.
Papon N, Courdavault V, Clastre M, Bennett RJ. Emerging and emerged pathogenic Candida
species: beyond the Candida albicans paradigm. PLoS Pathog 2013; 9: e1003550.
110
Paramonova E, Krom BP, van der Mei HC, Busscher HJ, Sharma PK. Hyphal content determines
the compression strength of Candida albicans biofilms. Microbiology 2009; 155: 1997-2003.
Parker RK, Dawsey SM, Abnet CC, White RE. Frequent occurrence of esophageal cancer in young
people in western Kenya. Dis Esophagus 2010; 23: 128-135.
Parlesak A, Schafer C, Schutz T, Bode JC, Bode C. Increased intestinal permeability to
macromolecules and endotoxemia in patients with chronic alcohol abuse in different stages of
alcohol-induced liver disease. J Hepatol 2000; 32: 742-747.
Pasqualotto AC, Nedel WL, Machado TS, Severo LC. Risk factors and outcome for nosocomial
breakthrough Candidaemia. J Infect 2006; 52: 216-222.
Patel PK, Erlandsen JE, Kirkpatrick WR, Berg DK, Westbrook SD, Louden C, Cornell JE,
Thompson GR, Vallor AC, Wickes BL, Wiederhold NP, Redding SW, Patterson TF. The
Changing Epidemiology of Oropharyngeal Candidiasis in Patients with HIV/AIDS in the Era of
Antiretroviral Therapy. AIDS Res Treat 2012; 2012: 262471.
Patton LL, Phelan JA, Ramos-Gomez FJ, Nittayananta W, Shiboski CH, Mbuguye TL. Prevalence
and classification of HIV-associated oral lesions. Oral Dis 2002; 8 Suppl 2: 98-109.
Pavlova SI, Jin L, Gasparovich SR, Tao L. Multiple alcohol dehydrogenases but no functional
acetaldehyde dehydrogenase causing excessive acetaldehyde production from ethanol by oral
streptococci. Microbiology 2013; 159: 1437-1446.
Pei Z, Bini EJ, Yang L, Zhou M, Francois F, Blaser MJ. Bacterial biota in the human distal
esophagus. Proc Natl Acad Sci U S A 2004; 101: 4250-4255.
Pelucchi C, Gallus S, Garavello W, Bosetti C, La Vecchia C. Alcohol and tobacco use, and cancer
risk for upper aerodigestive tract and liver. Eur J Cancer Prev 2008; 17: 340-344.
Persson B, Hallborn J, Walfridsson M, Hahnhägerdal B, Keränen S, Penttilä M, Jörnvall H. Dual
Relationships of Xylitol and Alcohol Dehydrogenases in Families of 2 Protein Types. Febs
Letters 1993; 324: 9-14.
Perumal P, Mekala S, Chaffin WL. Role for cell density in antifungal drug resistance in Candida
albicans biofilms. Antimicrob Agents Chemother 2007; 51: 2454-2463.
Petti S. Lifestyle risk factors for oral cancer. Oral Oncol 2009; 45: 340-350.
Pfaffl MW, Horgan GW, Dempfle L. Relative expression software tool (REST) for group-wise
comparison and statistical analysis of relative expression results in real-time PCR. Nucleic
Acids Res 2002; 30: e36.
Pfaller M, Boyken L, Hollis R, Kroeger J, Messer S, Tendolkar S, Diekema D. Use of
epidemiological cutoff values to examine 9-year trends in susceptibility of Candida species to
anidulafungin, caspofungin, and micafungin. J Clin Microbiol 2011; 49: 624-629.
Pfaller MA, Diekema DJ. Epidemiology of invasive candidiasis: a persistent public health problem.
Clin Microbiol Rev 2007; 20: 133-163.
111
Pfaller MA, Diekema DJ, Gibbs DL, Newell VA, Ellis D, Tullio V, Rodloff A, Fu W, Ling TA. Results
from the ARTEMIS DISK Global Antifungal Surveillance Study, 1997 to 2007: a 10.5-year
analysis of susceptibilities of Candida Species to fluconazole and voriconazole as determined
by CLSI standardized disk diffusion. J Clin Microbiol 2010; 48: 1366-1377.
Pfaller MA, Messer SA, Boyken L, Hollis RJ, Rice C, Tendolkar S, Diekema DJ. In vitro activities of
voriconazole, posaconazole, and fluconazole against 4,169 clinical isolates of Candida spp.
and Cryptococcus neoformans collected during 2001 and 2002 in the ARTEMIS global
antifungal surveillance program. Diagn Microbiol Infect Dis 2004; 48: 201-205.
Phan QT, Fratti RA, Prasadarao NV, Edwards JE, Jr., Filler SG. N-cadherin mediates endocytosis
of Candida albicans by endothelial cells. J Biol Chem 2005; 280: 10455-10461.
Phan QT, Myers CL, Fu Y, Sheppard DC, Yeaman MR, Welch WH, Ibrahim AS, Edwards JE, Jr.,
Filler SG. Als3 is a Candida albicans invasin that binds to cadherins and induces endocytosis
by host cells. PLoS Biol 2007; 5: e64.
Phillips BJ, Jenkinson P. Is ethanol genotoxic? A review of the published data. Mutagenesis 2001;
16: 91-101.
Pierce CG, Uppuluri P, Tristan AR, Wormley FL, Jr., Mowat E, Ramage G, Lopez-Ribot JL. A
simple and reproducible 96-well plate-based method for the formation of fungal biofilms and its
application to antifungal susceptibility testing. Nature protocols 2008; 3: 1494-1500.
Pierce JV, Dignard D, Whiteway M, Kumamoto CA. Normal adaptation of Candida albicans to the
murine gastrointestinal tract requires Efg1p-dependent regulation of metabolic and host
defense genes. Eukaryot Cell 2013; 12: 37-49.
Pietrella D, Rachini A, Torosantucci A, Chiani P, Brown AJ, Bistoni F, Costantino P, Mosci P,
d'Enfert C, Rappuoli R, Cassone A, Vecchiarelli A. A beta-glucan-conjugate vaccine and antibeta-glucan antibodies are effective against murine vaginal candidiasis as assessed by a
novel in vivo imaging technique. Vaccine 2010; 28: 1717-1725.
Pikkarainen PH, Baraona E, Jauhonen P, Seitz HK, Lieber CS. Contribution of oropharynx
microflora and of lung microsomes to acetaldehyde in expired air after alcohol ingestion. J Lab
Clin Med 1981; 97: 631-636.
Playford EG, Webster AC, Sorrell TC, Craig JC. Antifungal agents for preventing fungal infections
in non-neutropenic critically ill and surgical patients: systematic review and meta-analysis of
randomized clinical trials. J Antimicrob Chemother 2006; 57: 628-638.
Pohl H, Welch HG. The role of overdiagnosis and reclassification in the marked increase of
esophageal adenocarcinoma incidence. J Natl Cancer Inst 2005; 97: 142-146.
Pronk JT, Yde Steensma H, Van Dijken JP. Pyruvate metabolism in Saccharomyces cerevisiae.
Yeast 1996; 12: 1607-1633.
Puel A, Cypowyj S, Bustamante J, Wright JF, Liu L, Lim HK, Migaud M, Israel L, Chrabieh M,
Audry M, Gumbleton M, Toulon A, Bodemer C, El-Baghdadi J, Whitters M, Paradis T, Brooks
J, Collins M, Wolfman NM, Al-Muhsen S, Galicchio M, Abel L, Picard C, Casanova JL. Chronic
mucocutaneous candidiasis in humans with inborn errors of interleukin-17 immunity. Science
2011; 332: 65-68.
112
Qin J, Li R, Raes J, Arumugam M, Burgdorf KS, Manichanh C, Nielsen T, Pons N, Levenez F,
Yamada T, Mende DR, Li J, Xu J, Li S, Li D, Cao J, Wang B, Liang H, Zheng H, Xie Y, Tap J,
Lepage P, Bertalan M, Batto JM, Hansen T, Le Paslier D, Linneberg A, Nielsen HB, Pelletier
E, Renault P, Sicheritz-Ponten T, Turner K, Zhu H, Yu C, Li S, Jian M, Zhou Y, Li Y, Zhang X,
Li S, Qin N, Yang H, Wang J, Brunak S, Dore J, Guarner F, Kristiansen K, Pedersen O,
Parkhill J, Weissenbach J, Bork P, Ehrlich SD, Wang J. A human gut microbial gene catalogue
established by metagenomic sequencing. Nature 2010; 464: 59-65.
Quintin J, Saeed S, Martens JH, Giamarellos-Bourboulis EJ, Ifrim DC, Logie C, Jacobs L, Jansen
T, Kullberg BJ, Wijmenga C, Joosten LA, Xavier RJ, van der Meer JW, Stunnenberg HG,
Netea MG. Candida albicans infection affords protection against reinfection via functional
reprogramming of monocytes. Cell Host Microbe 2012; 12: 223-232.
Radfar L, Shea Y, Fischer SH, Sankar V, Leakan RA, Baum BJ, Pillemer SR. Fungal load and
candidiasis in Sjögren's syndrome. Oral Surg Oral Med Oral Pathol Oral Radiol Endod 2003;
96: 283-287.
Rajendran R, Mowat E, McCulloch E, Lappin DF, Jones B, Lang S, Majithiya JB, Warn P, Williams
C, Ramage G. Azole resistance of Aspergillus fumigatus biofilms is partly associated with
efflux pump activity. Antimicrob Agents Chemother 2011; 55: 2092-2097.
Rajendran R, Williams C, Lappin DF, Millington O, Martins M, Ramage G. Extracellular DNA
release acts as an antifungal resistance mechanism in mature Aspergillus fumigatus biofilms.
Eukaryot Cell 2013; 12: 420-429.
Ramage G, Bachmann S, Patterson TF, Wickes BL, Lopez-Ribot JL. Investigation of multidrug
efflux pumps in relation to fluconazole resistance in Candida albicans biofilms. J Antimicrob
Chemother 2002a; 49: 973-980.
Ramage G, Milligan S, Lappin DF, Sherry L, Sweeney P, Williams C, Bagg J, Culshaw S.
Antifungal, cytotoxic, and immunomodulatory properties of tea tree oil and its derivative
components: potential role in management of oral candidosis in cancer patients. Front
Microbiol 2012; 3: 220.
Ramage G, Robertson SN, Williams C. Strength in numbers: antifungal strategies against fungal
biofilms. Int J Antimicrob Agents 2014; 43: 114-120.
Ramage G, Saville SP, Wickes BL, Lopez-Ribot JL. Inhibition of Candida albicans biofilm formation
by farnesol, a quorum-sensing molecule. Appl Environ Microbiol 2002b; 68: 5459-5463.
Ramage G, Tomsett K, Wickes BL, Lopez-Ribot JL, Redding SW. Denture stomatitis: a role for
Candida biofilms. Oral Surg Oral Med Oral Pathol Oral Radiol Endod 2004; 98: 53-59.
Ramirez MA, Lorenz MC. Mutations in alternative carbon utilization pathways in Candida albicans
attenuate virulence and confer pleiotropic phenotypes. Eukaryot Cell 2007; 6: 280-290.
Ramos AN, Gobbato N, Rachid M, González L, Yantorno O, Valdez JC. Effect of Lactobacillus
plantarum and Pseudomonas aeruginosa culture supernatants on polymorphonuclear damage
and inflammatory response. Int Immunopharmacol 2010; 10: 247-251.
Rao R. Endotoxemia and gut barrier dysfunction in alcoholic liver disease. Hepatology 2009; 50:
638-644.
113
Rautemaa R, Hietanen J, Niissalo S, Pirinen S, Perheentupa J. Oral and oesophageal squamous
cell carcinoma--a complication or component of autoimmune polyendocrinopathy-candidiasisectodermal dystrophy (APECED, APS-I). Oral oncology 2007a; 43: 607-613.
Rautemaa R, Ramage G. Oral candidosis--clinical challenges of a biofilm disease. Crit Rev
Microbiol 2011; 37: 328-336.
Rautemaa R, Richardson M, Pfaller M, Koukila-Kahkola P, Perheentupa J, Saxen H. Decreased
susceptibility of Candida albicans to azole antifungals: a complication of long-term treatment in
autoimmune polyendocrinopathy-candidiasis-ectodermal dystrophy (APECED) patients. J
Antimicrob Chemother 2007b; 60: 889-892.
Rautemaa R, Rusanen P, Richardson M, Meurman JH. Optimal sampling site for mucosal
candidosis in oral cancer patients is the labial sulcus. J Med Microbiol 2006; 55: 1447-1451.
Reid G, Charbonneau D, Erb J, Kochanowski B, Beuerman D, Poehner R, Bruce AW. Oral use of
Lactobacillus rhamnosus GR-1 and L. fermentum RC-14 significantly alters vaginal flora:
randomized, placebo-controlled trial in 64 healthy women. FEMS Immunol Med Microbiol
2003; 35: 131-134.
Rindum JL, Stenderup A, Holmstrup P. Identification of Candida albicans types related to healthy
and pathological oral mucosa. J Oral Pathol Med 1994; 23: 406-412.
Ricicová M, Kucharíková S, Tournu H, Hendrix J, Bujdáková H, Van Eldere J, Lagrou K, Van Dijck
P. Candida albicans biofilm formation in a new in vivo rat model. Microbiology 2010; 156: 909919.
Robinson CJ, Young VB. Antibiotic administration alters the community structure of the
gastrointestinal microbiota. Gut Microbes 2010; 1: 279-284.
Robinson GM, Orrego H, Israel Y, Devenyi P, Kapur BM. Low-molecular-weight polyethylene glycol
as a probe of gastrointestinal permeability after alcohol ingestion. Dig Dis Sci 1981; 26: 971977.
Rodaki A, Bohovych IM, Enjalbert B, Young T, Odds FC, Gow NA, Brown AJ. Glucose promotes
stress resistance in the fungal pathogen Candida albicans. Mol Biol Cell 2009; 20: 4845-4855.
Rolston KV, Rodriguez S. Upper gastrointestinal disease in human immunodeficiency virusinfected individuals. Arch Intern Med 1992; 152: 881-882.
Romani L, Mencacci A, Cenci E, Spaccapelo R, Toniatti C, Puccetti P, Bistoni F, Poli V. Impaired
neutrophil response and CD4+ T helper cell 1 development in interleukin 6-deficient mice
infected with Candida albicans. J Exp Med 1996; 183: 1345-1355.
Room R, Babor T, Rehm J. Alcohol and public health. Lancet 2005; 365: 519-530.
Rosa DD, Pasqualotto AC, Denning DW. Chronic mucocutaneous candidiasis and oesophageal
cancer. Med Mycol 2008; 46: 85-91.
Rosen GP, Nielsen K, Glenn S, Abelson J, Deville J, Moore TB. Invasive fungal infections in
pediatric oncology patients: 11-year experience at a single institution. J Pediatr Hematol Oncol
2005; 27: 135-140.
114
Saint-Prix F, Bonquist L, Dequin S. Functional analysis of the ALD gene family of Saccharomyces
cerevisiae during anaerobic growth on glucose: the NADP+-dependent Ald6p and Ald5p
isoforms play a major role in acetate formation. Microbiology 2004; 150: 2209-2220.
Sakko M, Moore C, Novak-Frazer L, Rautemaa V, Sorsa T, Hietala P, Järvinen A, Bowyer P,
Tjäderhane L, Rautemaa R. 2-hydroxyisocaproic acid is fungicidal for Candida and Aspergillus
species. Mycoses 2013.
Sakko M, Tjäderhane L, Sorsa T, Hietala P, Järvinen A, Bowyer P, Rautemaa R. 2Hydroxyisocaproic acid (HICA): a new potential topical antibacterial agent. Int J Antimicrob
Agents 2012; 39: 539-540.
Salaspuro M. Acetaldehyde: a cumulative carcinogen in humans. Addiction 2009; 104: 551-553.
Salaspuro V, Salaspuro M. Synergistic effect of alcohol drinking and smoking on in vivo
acetaldehyde concentration in saliva. Int J Cancer 2004; 111: 480-483.
Salim N, Moore C, Silikas N, Satterthwaite J, Rautemaa R. Chlorhexidine is a highly effective
topical broad-spectrum agent against Candida spp. Int J Antimicrob Agents 2013a; 41: 65-69.
Salim N, Silikas N, Satterthwaite JD, Moore C, Ramage G, Rautemaa R. Chlorhexidineimpregnated PEM/THFM polymer exhibits superior activity to fluconazole-impregnated
polymer against Candida albicans biofilm formation. Int J Antimicrob Agents 2013b; 41: 193196.
Salminen MK, Rautelin H, Tynkkynen S, Poussa T, Saxelin M, Valtonen V, Järvinen A.
Lactobacillus bacteremia, clinical significance, and patient outcome, with special focus on
probiotic L. rhamnosus GG. Clin Infect Dis 2004; 38: 62-69.
Samaranayake LP. Oral mycoses in HIV infection. Oral Surg Oral Med Oral Pathol 1992; 73: 171–
180.
Samaranayake LP, Keung Leung W, Jin L. Oral mucosal fungal infections. Periodontology 2000
2009; 49: 39-59.
Samonis G, Skordilis P, Maraki S, Datseris G, Toloudis P, Chatzinikolaou I, Georgoulias V, Bodey
GP. Oropharyngeal candidiasis as a marker for esophageal candidiasis in patients with
cancer. Clin Infect Dis 1998; 27: 283-286.
Sandai D, Yin Z, Selway L, Stead D, Walker J, Leach MD, Bohovych I, Ene IV, Kastora S, Budge
S, Munro CA, Odds FC, Gow NA, Brown AJ. The evolutionary rewiring of ubiquitination targets
has reprogrammed the regulation of carbon assimilation in the pathogenic yeast Candida
albicans. MBio 2012; 11: 3.
Sanni AI. The need for process optimization of African fermented foods and beverages. Int J Food
Microbiol 1993; 18: 85-95.
Saville SP, Lazzell AL, Monteagudo C, Lopez-Ribot JL. Engineered control of cell morphology in
vivo reveals distinct roles for yeast and filamentous forms of Candida albicans during infection.
Eukaryot Cell 2003; 2: 1053-1060.
115
Sbordone L, Bortolaia C. Oral microbial biofilms and plaque-related diseases: microbial
communities and their role in the shift from oral health to disease. Clin Oral Investig 2003; 7:
181-188.
Scanlan PD, Marchesi JR. Micro-eukaryotic diversity of the human distal gut microbiota: qualitative
assessment using culture-dependent and -independent analysis of faeces. ISME J 2008; 2:
1183-1193.
Schwabe RF, Jobin C. The microbiome and cancer. Nat Rev Cancer 2013; 13: 800-812.
Scott BB, Jenkins D. Gastro-oesophageal candidiasis. Gut 1982; 23: 137-139.
Scully C, Bagan J. Oral squamous cell carcinoma overview. Oral Oncol 2009; 45: 301-308.
Secretan B, Straif K, Baan R, Grosse Y, El Ghissassi F, Bouvard V, Benbrahim-Tallaa L, Guha N,
Freeman C, Galichet L, Cogliano V. A review of human carcinogens--Part E: tobacco, areca
nut, alcohol, coal smoke, and salted fish. Lancet Oncol 2009; 10: 1033-1034.
Segata N, Haake SK, Mannon P, Lemon KP, Waldron L, Gevers D, Huttenhower C, Izard J.
Composition of the adult digestive tract bacterial microbiome based on seven mouth surfaces,
tonsils, throat and stool samples. Genome Biol 2012; 13: R42.
Seitz HK, Stickel F. Molecular mechanisms of alcohol-mediated carcinogenesis. Nat Rev Cancer
2007; 7: 599-612.
Sellam A, Tebbji F, Nantel A. Role of Ndt80p in sterol metabolism regulation and azole resistance
in Candida albicans. Eukaryot Cell 2009; 8: 1174-1183.
Sellart-Altisent M, Torres-Rodriguez JM, Gomez de Ana S, Alvarado-Ramirez E. [Nasal fungal
microbiota in allergic and healthy subjects]. Rev Iberoam Micol 2007; 24: 125-130.
Semlali A, Killer K, Alanazi H, Chmielewski W, Rouabhia M. Cigarette smoke condensate
increases C. albicans adhesion, growth, biofilm formation, and EAP1, HWP1 and SAP2 gene
expression. BMC Microbiol 2014; 14: 61.
Seneviratne CJ, Wang Y, Jin L, Abiko Y, Samaranayake LP. Candida albicans biofilm formation is
associated with increased anti-oxidative capacities. Proteomics 2008; 8: 2936-2947.
Seok J, Warren HS, Cuenca AG, Mindrinos MN, Baker HV, Xu W, Richards DR, McDonald-Smith
GP, Gao H, Hennessy L, Finnerty CC, Lopez CM, Honari S, Moore EE, Minei JP, Cuschieri J,
Bankey PE, Johnson JL, Sperry J, Nathens AB, Billiar TR, West MA, Jeschke MG, Klein MB,
Gamelli RL, Gibran NS, Brownstein BH, Miller-Graziano C, Calvano SE, Mason PH, Cobb JP,
Rahme LG, Lowry SF, Maier RV, Moldawer LL, Herndon DN, Davis RW, Xiao W, Tompkins
RG. Genomic responses in mouse models poorly mimic human inflammatory diseases. Proc
Natl Acad Sci USA 2013; 110: 3507-3512.
Sepehr A, Kamangar F, Fahimi S, Saidi F, Abnet CC, Dawsey SM. Poor oral health as a risk factor
for esophageal squamous dysplasia in northeastern Iran. Anticancer Res 2005; 25: 543-546.
Setiadi ER, Doedt T, Cottier F, Noffz C, Ernst JF. Transcriptional response of Candida albicans to
hypoxia: linkage of oxygen sensing and Efg1p-regulatory networks. J Mol Biol 2006; 361: 399411.
116
Shapiro KB, Hotchkiss JH, Roe DA. Quantitative relationship between oral nitrate-reducing activity
and the endogenous formation of N-nitrosoamino acids in humans. Food Chem Toxicol 1991;
29: 751-755.
Shin J, Hosur KB, Pyaram K, Jotwani R, Liang S, Chavakis T, Hajishengallis G. Expression and
function of the homeostatic molecule Del-1 in endothelial cells and the periodontal tissue. Clin
Dev Immunol 2013; 2013: 617809.
Shirtliff ME, Peters BM, Jabra-Rizk MA. Cross-kingdom interactions: Candida albicans and
bacteria. FEMS Microbiol Lett 2009; 299: 1-8.
Siikala E, Rautemaa R, Richardson M, Saxen H, Bowyer P, Sanglard D. Persistent Candida
albicans colonization and molecular mechanisms of azole resistance in autoimmune
polyendocrinopathy-candidiasis-ectodermal dystrophy (APECED) patients. J Antimicrob
Chemother 2010; 65: 2505-2513.
Silva S, Henriques M, Martins A, Oliveira R, Williams D, Azeredo J. Biofilms of non-Candida
albicans Candida species: quantification, structure and matrix composition. Med Mycol 2009;
47: 681-689.
Silverman RJ, Nobbs AH, Vickerman MM, Barbour ME, Jenkinson HF. Interaction of Candida
albicans cell wall Als3 protein with Streptococcus gordonii SspB adhesin promotes
development of mixed-species communities. Infect Immun 2010; 78: 4644-4652.
Simitsopoulou M, Roilides E, Dotis J, Dalakiouridou M, Dudkova F, Andreadou E, Walsh TJ.
Differential expression of cytokines and chemokines in human monocytes induced by lipid
formulations of amphotericin B. Antimicrob Agents Chemother 2005; 49: 1397-1403.
Soll DR. The role of phenotypic switching in the basic biology and pathogenesis of Candida
albicans. J Oral Microbiol 2014; 6.
Sorsa T, Tjäderhane L, Konttinen YT, Lauhio A, Salo T, Lee HM, Golub LM, Brown DL, Mäntylä P.
Matrix metalloproteinases: contribution to pathogenesis, diagnosis and treatment of
periodontal inflammation. Ann Med 2006; 38: 306-321.
Spellberg BJ, Ibrahim AS, Avanesian V, Fu Y, Myers C, Phan QT, Filler SG, Yeaman MR, Edwards
JE, Jr. Efficacy of the anti-Candida rAls3p-N or rAls1p-N vaccines against disseminated and
mucosal candidiasis. J Infect Dis 2006; 194: 256-260.
Squier CA, Kremer MJ, Wertz PW. Effect of ethanol on lipid metabolism and epithelial permeability
barrier of skin and oral mucosa in the rat. J Oral Pathol Med 2003; 32: 595-599.
Sreebny LM. Saliva in health and disease: an appraisal and update. Int Dent J 2000; 50: 140-161.
Staab JF, Bradway SD, Fidel PL, Sundstrom P. Adhesive and mammalian transglutaminase
substrate properties of Candida albicans Hwp1. Science 1999; 283: 1535-1538.
Steevens J, Schouten LJ, Goldbohm RA, van den Brandt PA. Alcohol consumption, cigarette
smoking and risk of subtypes of oesophageal and gastric cancer: a prospective cohort study.
Gut 2010; 59: 39-48.
Stichternoth C, Ernst JF. Hypoxic adaptation by Efg1 regulates biofilm formation by Candida
albicans. Appl Environ Microbiol 2009; 75: 3663-3672.
117
Stickel F, Schuppan D, Hahn EG, Seitz HK.
hepatocarcinogenesis. Gut 2002; 51: 132-139.
Cocarcinogenic
effects
of
alcohol
in
Stowell AR. An improved method for the determination of acetaldehyde in human blood with
minimal ethanol interference. Clin Chim Acta 1979; 98: 201-205.
Strijbis K, van Roermund CW, Hardy GP, van den Burg J, Bloem K, de Haan J, van Vlies N,
Wanders RJ, Vaz FM, Distel B. Identification and characterization of a complete carnitine
biosynthesis pathway in Candida albicans. FASEB J 2009; 23: 2349-2359.
Strijbis K, van Roermund CW, Visser WF, Mol EC, van den Burg J, MacCallum DM, Odds FC,
Paramonova E, Krom BP, Distel B. Carnitine-dependent transport of acetyl coenzyme A in
Candida albicans is essential for growth on nonfermentable carbon sources and contributes to
biofilm formation. Eukaryot Cell 2008; 7: 610-618.
Sudjana AN, Carson CF, Carson KC, Riley TV, Hammer KA. Candida albicans adhesion to human
epithelial cells and polystyrene and formation of biofilm is reduced by sub-inhibitory Melaleuca
alternifolia (tea tree) essential oil. Med Mycol 2012; 50: 863-870.
Sun JN, Solis NV, Phan QT, Bajwa JS, Kashleva H, Thompson A, Liu Y, Dongari-Bagtzoglou A,
Edgerton M, Filler SG. Host cell invasion and virulence mediated by Candida albicans Ssa1.
PLoS Pathog 2010; 6: e1001181.
Swoboda RK, Bertram G, Delbruck S, Ernst JF, Gow NA, Gooday GW, Brown AJ. Fluctuations in
glycolytic mRNA levels during morphogenesis in Candida albicans reflect underlying changes
in growth and are not a response to cellular dimorphism. Mol Microbiol 1994; 13: 663-672.
Synnott JM, Guida A, Mulhern-Haughey S, Higgins DG, Butler G. Regulation of the hypoxic
response in Candida albicans. Eukaryot Cell 2010; 9: 1734-1746.
Szabo G. Consequences of alcohol consumption on host defence. Alcohol Alcohol 1999; 34: 830841.
Szocs K, Lassegue B, Wenzel P, Wendt M, Daiber A, Oelze M, Meinertz T, Munzel T, Baldus S.
Increased superoxide production in nitrate tolerance is associated with NAD(P)H oxidase and
aldehyde dehydrogenase 2 downregulation. J Mol Cell Cardiol 2007; 42: 1111-1118.
Söderling EM, Hietala-Lenkkeri AM. Xylitol and erythritol decrease adherence of polysaccharideproducing oral streptococci. Curr Microbiol 2010; 60: 25-29.
Taff HT, Nett JE, Andes DR. Comparative analysis of Candida biofilm quantitation assays. Med
Mycol 2012; 50: 214-218.
Tanaka F, Yamamoto K, Suzuki S, Inoue H, Tsurumaru M, Kajiyama Y, Kato H, Igaki H, Furuta K,
Fujita H, Tanaka T, Tanaka Y, Kawashima Y, Natsugoe S, Setoyama T, Tokudome S, Mimori
K, Haraguchi N, Ishii H, Mori M. Strong interaction between the effects of alcohol consumption
and smoking on oesophageal squamous cell carcinoma among individuals with ADH1B and/or
ALDH2 risk alleles. Gut 2010; 59: 1457-1464.
Thanyasrisung P, Kesakomol P, Pipattanagovit P, Youngnak-Piboonratanakit P, Pitiphat W,
Matangkasombut O. Oral Candida carriage and immune status in Thai HIV-infected
individuals. J Med Microbiol 2014.
118
Theilade E. Factors controlling the microflora of the healthy mouth. In: Hill M J, Marsh P D. Human
microbial ecology. CRC Press, Inc. 1990 ;2–56.
Thomas DP, Viudes A, Monteagudo C, Lazzell AL, Saville SP, Lopez-Ribot JL. A proteomic-based
approach for the identification of Candida albicans protein components present in a subunit
vaccine that protects against disseminated candidiasis. Proteomics 2006; 6: 6033-6041.
Thomasson HR, Edenberg HJ, Crabb DW, Mai XL, Jerome RE, Li TK, Wang SP, Lin YT, Lu RB,
Yin SJ. Alcohol and aldehyde dehydrogenase genotypes and alcoholism in Chinese men. Am
J Hum Genet 1991; 48: 677-681.
Thompson DS, Carlisle PL, Kadosh D. Coevolution of morphology and virulence in Candida
species. Eukaryot Cell 2011; 10: 1173-1182.
Thompson GR, 3rd, Patel PK, Kirkpatrick WR, Westbrook SD, Berg D, Erlandsen J, Redding SW,
Patterson TF. Oropharyngeal candidiasis in the era of antiretroviral therapy. Oral Surg Oral
Med Oral Pathol Oral Radiol Endod 2010; 109: 488-495.
Thurlow LR, Hanke ML, Fritz T, Angle A, Aldrich A, Williams SH, Engebretsen IL, Bayles KW,
Horswill AR, Kielian T. Staphylococcus aureus Biofilms Prevent Macrophage Phagocytosis
and Attenuate Inflammation In vivo. Journal of Immunology 2011; 186: 6585-6596.
Tillonen J, Homann N, Rautio M, Jousimies-Somer H, Salaspuro M. Role of yeasts in the salivary
acetaldehyde production from ethanol among risk groups for ethanol-associated oral cavity
cancer. Alcohol Clin Exp Res 1999; 23: 1409-1415.
Tillonen J, Kaihovaara P, Jousimies-Somer H, Heine R, Salaspuro M. Role of catalase in in vitro
acetaldehyde formation by human colonic contents. Alcohol Clin Exp Res 1998; 22: 11131119.
Tortorano AM, Kibbler C, Peman J, Bernhardt H, Klingspor L, Grillot R. Candidaemia in Europe:
epidemiology and resistance. Int J Antimicrob Agents 2006; 27: 359-366.
Uittamo J, Siikala E, Kaihovaara P, Salaspuro M, Rautemaa R. Chronic candidosis and oral cancer
in APECED-patients: production of carcinogenic acetaldehyde from glucose and ethanol by
Candida albicans. Int J Cancer 2009; 124: 754-756.
Ullmann AJ, Lipton JH, Vesole DH, Chandrasekar P, Langston A, Tarantolo SR, Greinix H, Morais
de Azevedo W, Reddy V, Boparai N, Pedicone L, Patino H, Durrant S. Posaconazole or
fluconazole for prophylaxis in severe graft-versus-host disease. N Engl J Med 2007; 356: 335347.
Untergasser A, Nijveen H, Rao X, Bisseling T, Geurts R, Leunissen JA. Primer3Plus, an enhanced
web interface to Primer3. Nucleic Acids Res 2007; 35: W71-74.
Uppuluri P, Chaturvedi AK, Srinivasan A, Banerjee M, Ramasubramaniam AK, Kohler JR, Kadosh
D, Lopez-Ribot JL. Dispersion as an important step in the Candida albicans biofilm
developmental cycle. PLoS Pathog 2010a; 6: e1000828.
Uppuluri P, Pierce CG, Thomas DP, Bubeck SS, Saville SP, Lopez-Ribot JL. The transcriptional
regulator Nrg1p controls Candida albicans biofilm formation and dispersion. Eukaryot Cell
2010b; 9: 1531-1537.
119
Urban CF, Reichard U, Brinkmann V, Zychlinsky A. Neutrophil extracellular traps capture and kill
Candida albicans yeast and hyphal forms. Cell Microbiol 2006; 8: 668-676.
Utne HE, Winkler K. Hepatic and extrahepatic elimination of ethanol in cirrhosis. With estimates of
intrahepatic shunts and Km for ethanol elimination. Scand J Gastroenterol 1980; 15: 297-304.
Wachtler B, Citiulo F, Jablonowski N, Forster S, Dalle F, Schaller M, Wilson D, Hube B. Candida
albicans-epithelial interactions: dissecting the roles of active penetration, induced endocytosis
and host factors on the infection process. PLoS One 2012; 7: e36952.
Wagner RD, Pierson C, Warner T, Dohnalek M, Hilty M, Balish E. Probiotic effects of feeding heatkilled Lactobacillus acidophilus and Lactobacillus casei to Candida albicans-colonized
immunodeficient mice. J Food Prot 2000; 63: 638-644.
Wagner RD, Warner T, Roberts L, Farmer J, Balish E. Colonization of congenitally immunodeficient
mice with probiotic bacteria. Infect Immun 1997; 65: 3345-3351.
Wakhisi J, Patel K, Buziba N, Rotich J. Esophageal cancer in north rift valley of Western Kenya.
African Health Sci 2005; 5: 157-163.
Walraven CJ, Lee SA. Antifungal lock therapy. Antimicrob Agents Chemother 2013; 57: 1-8.
Walsh PS, Metzger DA, Higuchi R. Chelex-100 as a Medium for Simple Extraction of DNA for PcrBased Typing from Forensic Material. Biotechniques 1991; 10: 506-513.
van Enckevort FH, Netea MG, Hermus AR, Sweep CG, Meis JF, Van der Meer JW, Kullberg BJ.
Increased susceptibility to systemic candidiasis in interleukin-6 deficient mice. Med Mycol
1999; 37: 419-426.
van Leeuwenhoek A. An abstract of a Letter from Antonie van Leeuwenhoek, Sep. 12, 1683. About
Animals in the scurf of the Teeth. Philosophical Transactions of the Royal Society 1684; 14:
568–574.
van Monsjou HS, Wreesmann VB, van den Brekel MW, Balm AJ. Head and neck squamous cell
carcinoma in young patients. Oral Oncol 2013; 49: 1097-1102.
Van Urk H, Voll WS, Scheffers WA, Van Dijken JP. Transient-state analysis of metabolic fluxes in
crabtree-positive and crabtree-negative yeasts. Appl Environ Microbiol 1990; 56: 281-287.
Vardakas KZ, Michalopoulos A, Falagas ME. Fluconazole versus itraconazole for antifungal
prophylaxis in neutropenic patients with haematological malignancies: a meta-analysis of
randomised-controlled trials. Br J Haematol 2005; 131: 22-28.
Warnakulasuriya S. Global epidemiology of oral and oropharyngeal cancer. Oral Oncol 2009; 45:
309-316.
Warnakulasuriya S, Parkkila S, Nagao T, Preedy VR, Pasanen M, Koivisto H, Niemelä O.
Demonstration of ethanol-induced protein adducts in oral leukoplakia (pre-cancer) and cancer.
J Oral Pathol Med 2008a; 37: 157-165.
120
Warnakulasuriya S, Reibel J, Bouquot J, Dabelsteen E. Oral epithelial dysplasia classification
systems: predictive value, utility, weaknesses and scope for improvement. J Oral Pathol Med
2008b; 37: 127-133.
Warnock DW. Fungal infections in neutropenia: current problems and chemotherapeutic control. J
Antimicrob Chemother 1998; 41 Suppl D: 95-105.
Vasquez EM, Shin GP, Sifontis N, Benedetti E. Concomitant clotrimazole therapy more than
doubles the relative oral bioavailability of tacrolimus. Ther Drug Monit 2005; 27: 587-591.
Watts HJ, Very AA, Perera TH, Davies JM, Gow NA. Thigmotropism and stretch-activated
channels in the pathogenic fungus Candida albicans. Microbiology 1998; 144 ( Pt 3): 689-695.
Vediyappan G, Rossignol T, d'Enfert C. Interaction of Candida albicans biofilms with antifungals:
transcriptional response and binding of antifungals to beta-glucans. Antimicrob Agents
Chemother 2010; 54: 2096-2111.
Wei WQ, Abnet CC, Lu N, Roth MJ, Wang GQ, Dye BA, Dong ZW, Taylor PR, Albert P, Qiao YL,
Dawsey SM. Risk factors for oesophageal squamous dysplasia in adult inhabitants of a high
risk region of China. Gut 2005; 54: 759-763.
Velly AM, Franco EL, Schlecht N, Pintos J, Kowalski LP, Oliveira BV, Curado MP. Relationship
between dental factors and risk of upper aerodigestive tract cancer. Oral Oncol 1998; 34: 284291.
Wen CP, Tsai MK, Chung WS, Hsu HL, Chang YC, Chan HT, Chiang PH, Cheng TY, Tsai SP.
Cancer risks from betel quid chewing beyond oral cancer: a multiple-site carcinogen when
acting with smoking. Cancer Causes Control 2010; 21: 1427-1435.
Westermarck HW, Hietala P, Jaarma M, Sorsa T, Vaara M, inventors; Oy Extracta LTD, assignee.
Use of alpha-hydroxy acids in the manufacture of a medicament for the treatment of
inflammation. European Patent Office patent no. EP0871438B1. 1996; Published 27.2.2008.
Wheeler RT, Kombe D, Agarwala SD, Fink GR. Dynamic, morphotype-specific Candida albicans
beta-glucan exposure during infection and drug treatment. PLoS Pathog 2008; 4: e1000227.
Whitchurch CB, Tolker-Nielsen T, Ragas PC, Mattick JS. Extracellular DNA required for bacterial
biofilm formation. Science 2002; 295: 1487.
White TJ, T. Burns, S. Lee, and J. Taylor. Amplification and sequencing of fungal ribosomal RNA
genes for phylogenetics. In: M.A. Innis DHG, J. J. Sninsky, and T. J. White (ed.), ed. PCR
protocols. A guide to methods and applications. Academic Press Inc. 1990.
Wight AJ, Ogden GR. Possible mechanisms by which alcohol may influence the development of
oral cancer--a review. Oral Oncol 1998; 34: 441-447.
Wilcox CM, Straub RF, Clark WS. Prospective evaluation of oropharyngeal findings in human
immunodeficiency virus-infected patients with esophageal ulceration. Am J Gastroenterol
1995; 90: 1938-1941.
Villena J, Salva S, Aguero G, Alvarez S. Immunomodulatory and protective effect of probiotic
Lactobacillus casei against Candida albicans infection in malnourished mice. Microbiol
Immunol 2011; 55: 434-445.
121
Wilson D, Hebecker B, Moyes DL, Miramón P, Jablonowski N, Wisgott S, Allert S, Naglik JR, Hube
B. Clotrimazole dampens vaginal inflammation and neutrophil infiltration in response to
Candida albicans infection. Antimicrob Agents Chemother 2013; 57: 5178-5180.
Wilson K. Large scale CsCl preparation of genomic DNA from bacteria. . In: Ausubel FMR, Brent,
R.E., Kingston, D.D., Moore, J.G., Seidman, J.A., Smith, K. Struhl ed. Current Protocols in
Molecular Biology. New York: Greene Publishing Associates and Wiley Interscience; 1987.
Vonk AG, Netea MG, van Krieken JH, Iwakura Y, van der Meer JW, Kullberg BJ. Endogenous
interleukin (IL)-1 alpha and IL-1 beta are crucial for host defense against disseminated
candidiasis. J Infect Dis 2006; 193: 1419-1426.
Vora S, Purimetla N, Brummer E, Stevens DA. Activity of voriconazole, a new triazole, combined
with neutrophils or monocytes against Candida albicans: effect of granulocyte colonystimulating factor and granulocyte-macrophage colony-stimulating factor. Antimicrob Agents
Chemother 1998; 42: 907-910.
Woutersen RA, Appelman LM, Van Garderen-Hoetmer A, Feron VJ. Inhalation toxicity of
acetaldehyde in rats. III. Carcinogenicity study. Toxicology 1986; 41: 213-231.
Vylkova S, Nayyar N, Li W, Edgerton M. Human beta-defensins kill Candida albicans in an energydependent and salt-sensitive manner without causing membrane disruption. Antimicrob
Agents Chemother 2007; 51: 154-161.
Väkeväinen S, Mentula S, Nuutinen H, Salmela KS, Jousimies-Somer H, Färkkilä M, Salaspuro M.
Ethanol-derived microbial production of carcinogenic acetaldehyde in achlorhydric atrophic
gastritis. Scand J Gastroenterol 2002; 37: 648-655.
Väkeväinen S, Tillonen J, Agarwal DP, Srivastava N, Salaspuro M. High salivary acetaldehyde
after a moderate dose of alcohol in ALDH2-deficient subjects: strong evidence for the local
carcinogenic action of acetaldehyde. Alcohol Clin Exp Res 2000; 24: 873-877.
Väkeväinen S, Tillonen J, Blom M, Jousimies-Somer H, Salaspuro M. Acetaldehyde production and
other ADH-related characteristics of aerobic bacteria isolated from hypochlorhydric human
stomach. Alcohol Clin Exp Res 2001; 25: 421-426.
Xie Z, Thompson A, Sobue T, Kashleva H, Xu H, Vasilakos J, Dongari-Bagtzoglou A. Candida
albicans biofilms do not trigger reactive oxygen species and evade neutrophil killing. J Infect
Dis 2012; 206: 1936-1945.
Yang L, Lu X, Nossa CW, Francois F, Peek RM, Pei Z. Inflammation and intestinal metaplasia of
the distal esophagus are associated with alterations in the microbiome. Gastroenterology
2009; 137: 588-597.
Yeater KM, Chandra J, Cheng G, Mukherjee PK, Zhao X, Rodriguez-Zas SL, Kwast KE,
Ghannoum MA, Hoyer LL. Temporal analysis of Candida albicans gene expression during
biofilm development. Microbiology 2007; 153: 2373-2385.
Yin SJ, Chou FJ, Chao SF, Tsai SF, Liao CS, Wang SL, Wu CW, Lee SC. Alcohol and aldehyde
dehydrogenases in human esophagus: comparison with the stomach enzyme activities.
Alcohol Clin Exp Res 1993; 17: 376-381.
122
Yin SJ, Liao CS, Wu CW, Li TT, Chen LL, Lai CL, Tsao TY. Human stomach alcohol and aldehyde
dehydrogenases: comparison of expression pattern and activities in alimentary tract.
Gastroenterology 1997; 112: 766-775.
Yokoyama A, Tsutsumi E, Imazeki H, Suwa Y, Nakamura C, Mizukami T, Yokoyama T. Salivary
acetaldehyde concentration according to alcoholic beverage consumed and aldehyde
dehydrogenase-2 genotype. Alcohol Clin Exp Res 2008; 32: 1607-1614.
Yokoyama A, Tsutsumi E, Imazeki H, Suwa Y, Nakamura C, Yokoyama T. Contribution of the
alcohol dehydrogenase-1B genotype and oral microorganisms to high salivary acetaldehyde
concentrations in Japanese alcoholic men. Int J Cancer 2007; 121: 1047-1054.
Yoshida A, Rzhetsky A, Hsu LC, Chang C. Human aldehyde dehydrogenase gene family. Eur J
Biochem 1998; 251: 549-557.
Yu G, Gail MH, Shi J, Klepac-Ceraj V, Paster BJ, Dye BA, Wang GQ, Wei WQ, Fan JH, Qiao YL,
Dawsey SM, Freedman ND, Abnet CC. Association between Upper Digestive Tract Microbiota
and Cancer-Predisposing States in the Esophagus and Stomach. Cancer Epidemiol
Biomarkers Prev 2014.
Zhao YF, Xiong W, Wu XL. Mesenchymal stem cell-based developmental endothelial locus-1 gene
therapy for acute lung injury induced by lipopolysaccharide in mice. Mol Med Rep 2014; 9:
1583-1589.
Zijnge V, van Leeuwen MB, Degener JE, Abbas F, Thurnheer T, Gmur R, Harmsen HJ. Oral biofilm
architecture on natural teeth. PLoS One 2010; 5: e9321.
123