BookPDF Available

International Society for the Study of Vulvovaginal Disease Recommendations for the Diagnosis and Treatment of Vaginitis

Authors:

Abstract and Figures

Members and non-members of the International Society for the Study of Vulvovaginal Dis- ease (ISSVD), acknowledged as experts in the field of vulvovaginitis, from different countries and backgrounds, were invited to participate in this mission. Participants were involved in one or more working groups, according to their expertise and interest. Each group performed a systematic review and produced a draft based on that. The next step of the process consisted of discussion of the drafts, open to all participants involved in the development of this document. Finally, all drafts were reviewed by the editors and sent back for discussion in case of need.
Content may be subject to copyright.
International Society for the Study
of Vulvovaginal Disease
Recommendations for the Diagnosis
and Treatment of Vaginitis
Editors:
Pedro Vieira-Baptista, Colleen K. Stockdale, Jack Sobel
March 2023
Cite document as:
Vieira-Baptista P, Stockdale CK, Sobel J (eds). International Society for the Study of Vulvovaginal
Disease recommendations for the diagnosis and treatment of vaginitis. Lisbon: Admedic, 2023
ISBN: 978-989-53489-3-0
Legal deposit: 512780/23
DOI: https://doi.org/10.59153/adm.rdtv.001
Property
ISSVD
www.issvd.org
Edition & Pagination
Ad Médic, Lda.
www.admedic.pt
Published March 2023
This publication was inspired by the goals and mission of the International Society for the
Study of Vulvovaginal Disease (ISSVD) which are:
to promote international communication among gynecologists, pathologists, dermatol-
ogists, and other healthcare providers;
to establish international agreement on terminology and definitions of vulvovaginal diseases;
to promote clinical investigation, basic research and dissemination of knowledge in this field.
Visit www.issvd.org for more information.
INDEX
Authors 13
Preface 17
Note 18
CHAPTER 1 – The normal discharge 21
1.1 The vaginal microbiome and other components of the normal discharge 21
Host cell components 22
Soluble components and mucus 23
Microorganisms 23
Bacteria 23
Viruses 24
Yeasts 25
1.2 Normal vaginal discharge variations during the menstrual cycle 25
1.3 Normal vaginal discharge in physiological estrogen deficiency 26
(pre-menstrual girls, postpartum, and post-menopausal women)
1.4 Normal vaginal discharge during pregnancy 28
1.5 Factors contributing to variations in the composition of the vaginal discharge 28
Stress 28
Sexual activity 28
Douching 29
Smoking 29
Diet 29
1.6 Racial differences in the composition of the vaginal discharge and vaginal microbiota 29
1.7 Summary and conclusions 30
Recommendations 30
References 30
CHAPTER 2 – Diagnostic tools 37
2.1 Introduction 37
2.2 pH 37
2.3 Whiff test 39
2.4 Wet mount microscopy 39
2.5 Gram and other staining techniques 44
2.6 Cultures 46
2.7 Amsel criteria 46
2.8 Enzymatic tests 48
2.9 Molecular tests 49
Recommendations 53
References 55
CHAPTER 3 – Bacterial vaginosis 61
3.1 Introduction 61
3.2 Etiology and physiopathology 62
3.3 Prevalence and epidemiology 63
Global and regional estimates of bacterial vaginosis prevalence 63
Bacterial vaginosis prevalence among pregnant women 64
Bacterial vaginosis prevalence among other populations/sub-groups of women 64
3.4 Risk factors 65
3.5 Complications 66
Preterm birth 66
Endometritis/ postpartum fever 66
Post hysterectomy vaginal cuff cellulitis 67
Post abortal infection 67
Pelvic inflammatory disease 67
Other sexually transmitted infections 67
3.6 Signs and symptoms 67
3.7 Diagnosis 68
Clinical diagnosis 69
Gram-stain diagnosis 70
Cultures 71
Point-of-care tests (non-molecular) 71
Molecular diagnostics 72
Differential diagnosis 72
3.8 Treatment 73
The principles of treatment for non-pregnant women; drug selection, dosing, 74
adverse effects, and efficacy
Oral versus vaginal treatment 74
Metronidazole 74
Clindamycin 74
Overview of second-line and alternative treatments 75
Dequalinium chloride 75
Tinidazole 75
Secnidazole 76
Focused assessments of experimental/investigational treatments 76
Efficacy of probiotics for bacterial vaginosis treatment 76
Follow-up 76
Treatment regimens during pregnancy and lactation 76
Special considerations 77
Counseling-supportive management in infertile women attending fertility treatment 77
Screening and treatment of asymptomatic bacterial vaginosis in pregnancy 77
Approache to preoperative screening strategies for bacterial vaginosis 77
Management of sexual partners 77
Management of recurrent and refractory bacterial vaginosis 78
3.9 Special situations 80
Infancy 80
Postmenopausal women 80
Immunosuppression 81
Bacterial vaginosis in pregnancy 81
3.10 Future perspectives 82
Recommendations 83
References 84
CHAPTER 4 – Candidiasis 95
4.1 Introduction 95
4.2 Etiology and pathophysiology 95
4.3 Prevalence and epidemiology 96
4.4 Risk factors 97
4.5 Classification of vulvovaginal candidiasis 97
Severe infections 98
Recurrent infections 98
Non-albicans Candida infections 98
4.6 Signs and symptoms 99
4.7 Diagnosis 100
4.8 Treatment of vulvovaginal candidiasis 102
Asymptomatic colonization 102
Uncomplicated vulvovaginal candidiasis 102
Recurrent vulvovaginal candidiasis due to C. albicans 103
Non-albicans Candida vulvovaginitis 104
4.9 Special situations 104
Prepubertal children 104
Pregnancy 105
Postpartum and breastfeeding mothers 105
Menopause 105
Immunosuppression 106
4.10 Future perspectives 106
Recommendations 107
References 108
CHAPTER 5 – Trichomoniasis 113
5.1 Introduction 113
5.2 Etiology and pathophysiology 113
5.3 Prevalence and epidemiology 115
5.4 Risk factors 115
5.5 Complications 116
Adverse birth outcomes 116
Risk of HIV 117
Risk of other sexually transmitted infections 118
Pelvic inflammatory disease 118
Infertility 118
Risk of cervical cancer 118
5.6 Signs and symptoms 119
5.7 Diagnosis 119
5.8 Treatment and follow-up 122
5.9 Special situations 123
Infants 123
Pregnant and lactating women 123
5-nitroimidazole hypersensitivity 123
Persistent T. vaginalis infection 124
HIV-infected women 125
Partner management 125
5.10 Future perspectives 125
Recommendations 126
References 127
CHAPTER 6 – Cytolytic vaginosis, lactobacillosis and leptothrix 133
6.1 Introduction 133
6.2 Cytolytic vaginosis 133
Prevalence and epidemiology 134
Risk factors 134
Complications 134
Signs and symptoms 134
Diagnosis 134
Treatment 135
Special situations (infancy, pregnancy, postpartum/breastfeeding, menopause, 136
immunosuppression)
Future perspectives 136
6.3 Leptothrix 136
Prevalence and epidemiology 137
Risk factors 137
Complications 137
Signs and symptoms 138
Diagnosis 138
Treatment 139
Special situations (infancy, pregnancy, postpartum/breastfeeding, menopause, 140
immunosuppression)
6.4 Future perspectives 140
Recommendations 140
References 140
CHAPTER 7 –Aerobic vaginitis/desquamative inflammatory vaginitis 143
7.1 Introduction 143
7.2 Etiology and physiopathology 144
7.3 Prevalence and epidemiology 145
7.4 Risk factors 145
7.5 Complications 145
7.6 Signs and symptoms 146
7.7 Diagnosis 147
7.8 Treatment 148
7.9 Special situations (pregnancy, postpartum/breastfeeding) 150
7.10 Future perspectives 151
Recommendations 151
References 152
CHAPTER 8 – Vulvovaginal atrophy 157
8.1 Introduction 157
8.2 Etiology and physiopathology 157
8.3 Prevalence and epidemiology 158
8.4 Complications 159
8.5 Signs and symptoms 159
8.6 Diagnosis 160
8.7 Treatment 162
Vaginal lubricants and moisturizers 162
Estrogen (systemic and vaginal) and selective estrogen receptor modulator therapy 163
Vaginal androgen (testosterone) therapy 164
Vaginal dehydroepiandrosterone (prasterone) therapy 165
Vaginal LASER 165
Vaginal radiofrequency 166
Pelvic floor rehabilitation 166
8.8 Special situations (postpartum/breastfeeding, breast cancer) 166
8.9 Future perspectives 167
Recommendations 168
References 168
CHAPTER 9 – Vaginitis in children 173
9.1 Introduction 173
9.2 The vagina in the prepubertal child 173
The vestibule 173
The hymen 173
The vagina and vaginal discharge 174
9.3 How to conduct a vaginal examination in a child 175
9.4 Testing for infection 176
9.5 Vaginal discharge 176
Vaginal discharge due to infection 176
Bacterial vaginitis 176
Introduction 176
Prevalence 177
Etiology and pathophysiology 177
Risk factors 177
Signs and symptoms 177
Diagnosis 177
Treatment 178
Pinworms/threadworms 178
Introduction 178
Etiology and pathophysiology 178
Prevalence 179
Risk Factors 179
Signs and symptoms 179
Diagnosis 179
Treatment 179
Candidiasis 179
Bacterial vaginosis 180
Other infections producing vaginitis 180
Vaginal foreign bodies 181
Introduction 181
Etiology and pathophysiology 181
Prevalence 181
Signs and Symptoms 181
Diagnosis 181
Treatment 182
9.6 Dermatoses and dermatitis which involve the vagina 182
Recommendations 182
References 183
CHAPTER 10 – Probiotics, prebiotics and synbiotics for vaginitis 185
10.1 Introduction 185
10.2 Bacterial vaginosis 186
Probiotics 186
Probiotics for bacterial vaginosis in pregnancy 188
Prebiotics and synbiotics 188
Clinical recommendations 189
10.3 Vulvovaginal candidiasis 189
Probiotics 189
Prebiotics and synbiotics 190
Clinical recommendations 190
10.4 Aerobic vaginitis/desquamative inflammatory vaginitis 190
Probiotics 190
Prebiotics and synbiotics 191
Clinical recommendations 191
10.5 Trichomoniasis 191
Probiotics 191
Prebiotics and synbiotics 192
Clinical recommendations 192
10.6 Conclusion 192
Recommendations 192
References 193
INDEX OF FIGURES
FIGURE 1.1 Normal vaginal discharge. 21
A – wet mount microscopy (400x, phase contrast)
B – Gram stain (1000x, oil immersion)
FIGURE 1.2 Wet mount microscopy of a vaginal sample collected from 27
a breastfeeding woman (400x, phase contrast).
FIGURE 2.1 Vaginal pH measurement. 38
FIGURE 2.2 Wet mount microscopy (400x, phase contrast). 40
A– normal
B– bacterial vaginosis
C– candidiasis
D– trichomoniasis
E– cytolytic vaginosis
F– leptothrix
G– desquamative inflammatory vaginitis (severe aerobic vaginitis)
H– vaginal atrophy
FIGURE 2.3 Wet mount microscopy (400x); granular microbiota suggestive of bacterial vaginosis. 41
A– without phase contrast
B– with phase contrast
FIGURE 2.4 Sampling of vaginal discharge for wet mount microscopy. 41
FIGURE 2.5 Wet mount microscopy with phase contrast (400x) showing the presence 48
of a “mixed infection” (Candida spp. and bacterial vaginosis).
FIGURE 3.1 Typical discharge associated with bacterial vaginosis. 68
FIGURE 3.2 Wet mount microscopy (400x, phase contrast). 69
A-C– bacterial vaginosis: absence of lactobacilli, granular microbiota
and presence of clue cells (seen in A)
FIGURE 3.3 Gram stain (1000x, oil immersion). 70
A and B– Bacterial vaginosis
FIGURE 4.1 Acute vulvovaginal candidiasis. Exteriorization of white “cheesy” discharge, 99
vulvar erythema and edema.
FIGURE 4.2 Acute vulvovaginal candidiasis. Erythema and fissures of the intelabial sulci. 99
FIGURE 4.3 Acute vulvovaginal candidiasis. Adherent white discharge on the sidewalls 100
of the vagina and cervix.
FIGURE 4.4 Wet mount microscopy (400x, phase contrast). 100
A– Blastospores (culture positive for C. krusei)
B– Hyphae and blastospores (culture positive for C. albicans)
FIGURE 4.5 Gram stain (1000x, oil immersion). 101
A– Blastospores
B– Hyphae and blastospores
12
FIGURE 5.1 A and B – Trichomonads seen with Gram stain (1000x, oil immersion). 114
FIGURE 5.2 “Strawberry cervix”. 119
FIGURE 5.3 Wet mount microscopy. 119
A– Several trichomonads and inflammation (200x)
B– Trichomonad with its typical structures: flagella on the outside
and hydrogenosomes on the inside (400x)
C– Several trichomonads, inflammation and bacterial vaginosis (400x,
phase contrast)
FIGURE 6.1 Typical discharge associated with cytolytic vaginosis. 134
FIGURE 6.2 Cytolytic vaginosis. 135
A– wet mount microscopy (400x, phase contrast)
B– Gram stain (1000x, oil immersion); Pap smear (conventional) (400x)
FIGURE 6.3 Leptothrix seen in wet mount microscopy (400x, phase contrast). 138
A– leptothrix and normal background microbiota
B– leptothrix and Candida spp. blastospores
FIGURE 6.4 Leptothrix seen using Gram stain (1000x, oil immersion). 139
FIGURE 7.1 Severe aerobic vaginitis/desquamative inflammatory vaginitis. 146
A– vaginal and cervical petechiae
B– copious discharge
C– vestibular involvement
FIGURE 7.2 Aerobic vaginitis/desquamative inflammatory vaginitis in wet mount 147
microscopy (400x, phase contrast).
A– moderate aerobic vaginitis
B and C– severe aerobic vaginitis/desquamative inflammatory vaginitis
FIGURE 7.3 Aerobic vaginitis/desquamative inflammatory vaginitis aspects with Gram stain 147
(1000x, oil immersion). Chains of cocci seen in A and B.
FIGURE 8.1 A and B - Colposcopic aspect of the vagina of a postmenopausal woman. 159
Loss of vaginal rugae, petechiae, and easy bleeding.
FIGURE 8.2 Flowchart for the clinical assessment in suspected vaginal atrophy/atrophic vaginitis. 160
FIGURE 8.3 Wet mount microscopy (400x, phase contrast). 161
A– vaginal atrophy
B– atrophic vaginitis
FIGURE 8.4 Gram stain (1000x, oil immersion), vaginal atrophy. 161
A– vaginal atrophy
B– atrophic vaginitis
FIGURE 9.1 Wet mount microscopy (200x) from a prepubertal girls’ vagina. 174
A– Exclusive presence of parabasal cells; lactobacilli absent
B– Presence of inflammation in a case of bacterial vaginitis
FIGURE 9.2 Vulvar and perianal redness in a young girl with bacterial vaginitis (S. pyogenes). 177
13
Editors:
Pedro Vieira-Baptista (https://orcid.org/0000-0001-5335-6770)
Colleen K. Stockdale (https://orcid.org/0000-0003-0074-3261)
Jack Sobel (https://orcid.org/0000-0002-5589-4609)
Authors (alphabetical order):
Susana Aidé (https://orcid.org/0000-0002-4212-0022)
Maternal and Child Department, Faculty of Medicine, Universidade Federal Fluminense, Rio de
Janeiro, Brazil
Lower Genital Tract Disease, Hospital Universitário António Pedro, Niterói, Rio de Janeiro, Brazil
Jacob Bornstein (https://orcid.org/0000-0003-1932-5270)
Bar-Ilan University Faculty of Medicine
Fulvio Borella (https://orcid.org/0000-0001-8398-7557)
Gynecology and Obstetrics 1, Department of Surgical Sciences, City of Health and Science, Uni-
versity of Turin, Turin, Italy.
Catriona Bradshaw (https://orcid.org/0000-0002-6643-5678)
Melbourne Sexual Health Centre Monash University, Victoria, Australia
Libby Edwards (https://orcid.org/0009-0001-8902-5532)
Southeast Vulvar Clinic, NC, USA
Sophia Ehrström (https://orcid.org/0000-0001-8945-9050)
Considra Gyn, Nacka Hospital, Stockholm, Sweden
Karolinska Institutet, Danderyd Hospital, Stockholm, Sweden
Gayle Fischer (https://orcid.org/0000-0002-6382-2576)
The University of Sydney, NSW, Australia
Royal North Shore Hospital, St Leonard’s, NSW, Australia
Švitrigailė Grincevičienė (https://orcid.org/0000-0003-0370-0523)
Vilnius University, Life Science Centre, Institute of Biotechnology, Vilnius, Lithuania
Patricia J. Kissinger (https://orcid.org/0000-0002-2038-9151)
Tulane University School of Public Health and Tropical Medicine, LA, USA
Roni Kraut (https://orcid.org/0000-0003-3354-4955)
Department of Family Medicine, University of Alberta, Alberta, Canada
Iara Linhares (https://orcid.org/0000-0002-7846-6885)
Department of Obstetrics and Gynecology, Faculty of Medicine, University of São Paulo, Brazil
Colin MacNeill (https://orcid.org/0009-0000-7486-2701)
Vulvovaginal Service, Department of Obstetrics & Gynecology, Harvard Vanguard Medical Asso-
ciates, Atrius Health, MA, USA
José Martinez de Oliveira (https://orcid.org/0000-0001-8608-0248)
Heath Sciences Research Centre, University of Beira Interior, Covilhã, Portugal
14
Werner Mendling (https://orcid.org/0000-0002-6845-9411)
German Center for Infections in Gynecology and Obstetrics, at Helios University Hospital, Wup-
pertal, Germany
Caroline Mitchell (https://orcid.org/0000-0001-9924-2349)
Vincent Center for Reproductive Biology, Massachusetts General Hospital, MA, USA
Harvard Medical School, MA, USA
Christina A. Muzny (https://orcid.org/0000-0002-4005-3858)
Division of Infectious Diseases, University of Alabama at Birmingham, Birmingham, AL, USA
Paul Nyirjesy (https://orcid.org/0000-0001-5309-2971)
Sidney Kimmel Medical College at Thomas Jefferson University, Philadelphia, PA, USA
Caroline Oliveira (https://orcid.org/0000-0002-7207-7969)
Universidade Federal Fluminense, Rio de Janeiro, Brazil
Faustino R. Pérez-López (https://orcid.org/0000-0002-2801-416X)
Faculty of Medicine, University of Zaragoza, Zaragoza, Spain
Instituto Aragonés de Ciencias de la Salud, Zaragoza, Spain
Erica L. Plummer (https://orcid.org/0000-0001-5710-2671)
Central Clinical School, Monash University, Melbourne, Australia
Mario Preti (https://orcid.org/0000-0002-1573-3114)
Department of Surgical Sciences, University of Torino, Torino, Italy
Koray Görkem Saçıntı (https://orcid.org/0000-0002-8602-9714)
Ankara University School of Medicine, Department of Obstetrics and Gynecology, Ankara, Turkey
Francesco de Seta (https://orcid.org/0000-0003-1611-0813)
Department of Medical, Surgical and Health Sciences, University of Trieste, Trieste, Italy
Institute of Maternal and Child Health IRCCS B. Garofolo, Trieste, Italy
Ana Rita Silva (https://orcid.org/0009-0003-7587-3922)
Unilabs Portugal - Molecular Laboratory Diagnostics, Porto, Portugal
Henrique Soares (https://orcid.org/0000-0001-6864-9912)
Neonatology Department, Centro Hospitalar Universitário de São João, Porto, Portugal
Department of Gynecology-Obstetrics and Pediatrics, Faculdade de Medicina da Universidade do
Porto, Porto, Portugal
Jack Sobel (https://orcid.org/0000-0002-5589-4609)
Wayne State University School of Medicine, MI, USA
Ryan Sobel (https://orcid.org/0000-0002-1959-2910)
Jefferson Health, Philadelphia, PA, USA
Sidney Kimmel Medical College at Thomas Jefferson University, Philadelphia, PA, USA
Carlos Sousa (https://orcid.org/0000-0003-0708-6874)
Unilabs Portugal - Molecular Laboratory Diagnostics, Porto, Portugal
15
Colleen K. Stockdale (https://orcid.org/0000-0003-0074-3261)
Department of Obstetrics & Gynecology, University of Iowa City, IA, USA
Päivi Tommola (https://orcid.org/0009-0000-2634-474X)
University of Helsinki, Finland
Isabel do Val (https://orcid.org/0000-0002-9885-2724)
Maternal and Child Department, Faculty of Medicine, Universidade Federal Fluminense, Rio de
Janeiro, Brazil
Lower Genital Tract Disease, Hospital Universitário António Pedro, Niterói, Rio de Janeiro, Brazil
Gary Ventolini (https://orcid.org/0000-0001-8067-7342)
School of Medicine, Texas Tech University Health Sciences Center PB, TX, USA
Hans Verstraelen (https://orcid.org/0000-0002-4070-1587)
Department of Obstetrics & Gynecology, Ghent University Hospital, Ghent, Belgium
Department of Human Structure and Repair, Faculty of Medicine & Health Sciences, Ghent Uni-
versity, Ghent, Belgium
Pedro Vieira-Baptista (https://orcid.org/0000-0001-5335-6770)
Lower Genital Tract Unit, Centro Hospitalar de São João, Porto, Portugal
Hospital Lusíadas Porto, Porto, Portugal
Department of Gynecology-Obstetrics and Pediatrics, Faculdade de Medicina da Universidade do
Porto, Porto, Portugal
Conflicts of interest
Catriona Bradshaw
Research funding: National Health and Medical Research Funding and Australian Re-
searchCouncil
Werner Mendling
Advisory Boards: Bayer AG, Gedeon Richter Plc., Johnson & Johnson Inc., Meda Pharma
GmbH, PhagoMed Biopharma GmbH, Dr. August Wolff GmbH & Co.
Consultant fees/honoraria: Aristo Pharma GmbH, Bayer AG, Dr. August Wolff GmbH & Co.,
Dr. Kade Pharmazeutische Fabrik GmbH, KARO Pharma AB, Medinova AG, Pierre Fabre SA,
Sekisui Diagnostics Co.
Caroline Mitchell
Research funding: Scynexis Inc.
Consultant fees/honoraria: Ferring Pharmaceuticals Plc., Scynexis Inc., UpToDate Inc.
Christina A. Muzny
Research funding: NIH/NIAID, Abbott, Gilead Inc., Lupin Inc.
Consultant fees/honoraria: Scynexis Inc., BioNTech SE, Cepheid Inc., Abbott, Roche AG
Honoraria for educational speaking events and reviews: Visby Medical Inc., Scynexis Inc.,
Elsevier Ltd, UpToDate Inc.
16
Paul Nyirjesy
Consultant fees/honoraria: Mycovia Pharmaceuticals Inc., Scynexis Inc., Hologic Inc.
Francesco de Seta
Advisory Boards: Bayer AG, Alfasigma SpA, Gedeon-Richter Plc., Medinova AG
Consultant fees/honoraria: HIS Ltd, Uniderm Ltd, Depofarma SpA
Jack Sobel
Consultant fees/honoraria: Mycovia Pharmaceuticals Inc., Scynexis Inc.
Honoraria for educational speaking events and reviews: UpToDate Inc.
Pedro Vieira-Baptista
Research funding: Seegene Inc.
Honoraria for educational speaking events: Seegene Inc., Medinova AG, Merck & Co., Inc.
All other authors did not declare any conflicts of interest.
Sponsorship
Unrestricted educational grant from Scynexis, Inc.
All content and data reported are exclusively of the author’s responsibility.
17
Vulvovaginitis is among the most common gynecologic diagnoses in both primary care
and lower genital tract specialist care worldwide, with most women experiencing at least
one lifetime episode. Accordingly, the need for uniform, simplified and standardized man-
agement directives to both diagnose and treat vulvovaginal infection is great globally and
therefore on a national basis, medical professional societies have already undertaken to
publish guidelines to optimize therapy, but often with considerable differences given vari-
ation in availability of diagnostic tests, clinical expertise, drug availability and access. More-
over, the rapid progress in development and availability of new diagnostic tests and thera-
peutic agents dictates that guidelines be updated frequently. Unfortunately, timely updates
are frequently not forthcoming. So do practitioners need yet another version of instruction
and guidelines? The International Society for the Study of Vulvovaginal Disease (ISSVD) is
unique, with a membership that is worldwide reflecting the variable needs and standards of
different communities. So the international design of the “writing” teams afforded an oppor-
tunity to standardize guidelines to reflect the variable needs of women in societies with dif-
ferences in patient needs and practitioner availability. The new ISSVD recommendations are
designed to overcome cultural, social and financial differences in global societies utilizing
our team approach. Also unique to ISSVD recommendations was the inclusion of a strong
educational background for each clinical entity together with treatment rationale. Authors
recognized that there have been major advances in diagnostic tests reflecting the applica-
tion of advances in molecular technology in new superior diagnostic tests. Author opinion
emphasized that the “syndromic” approach is no longer acceptable and that empiricism in
treatment selection has to be avoided at all costs. Final guidelines followed extensive review
and discussion. The new ISSVD recommendations will be updated on a regular and frequent
basis and represent the views and experience of the society membership including highly
respected experts with global reputations.
Producing the 2023 recommendations not only represents a major contribution to women’s
health but serves as an act of altruism by all contributors.
J D Sobel MD
Distinguished Professor of Medicine
Dean Emeritus Wayne State University School of Medicine
PREFACE
18
19
Members and non-members of the International Society for the Study of Vulvovaginal Dis-
ease (ISSVD), acknowledged as experts in the field of vulvovaginitis, from different countries
and backgrounds, were invited to participate in this mission.
Participants were involved in one or more working groups, according to their expertise and
interest.
Each group performed a systematic review and produced a draft based on that. The next
step of the process consisted of discussion of the drafts, open to all participants involved in
the development of this document. Finally, all drafts were reviewed by the editors and sent
back for discussion in case of need.
The levels of evidence and grades of recommendation in the final tables of each chapter
were based on the “Oxford Centre for Evidence-Based Medicine: Levels of Evidence”.1
The final version of the document was accepted by all authors.
1Oxford Centre for Evidence-Based Medicine: Levels of Evidence.
https://www.cebm.ox.ac.uk/resources/levels-of-evidence/oxford-centre-for-evidence-based-medicine-lev-
els-of-evidence-march-2009
NOTE
20
21
1.1
The vaginal microbiome and other components
of the normal discharge
Vaginal discharge is described as the fluid excreted from the vagina. It may be pathological
or physiological.1 Normal discharge is usually clear or white, and without an offensive odor.
The consistency varies from thick and sticky to stretchy.2 The normal amount of vaginal dis-
charge is about 1-3 mL daily.3 Women may have different concepts concerning what is a nor-
mal discharge.4 Sometimes, women may note an increased discharge (as a symptom) and have
a “normal” discharge. Nevertheless, more knowledge and markers of normality are needed.5, 6
Fluids present in the vagina include those that originate in the vagina itself, but also from
the cervix and the upper genital tract and some not produced by the woman.7 (Figure 1.1)
Figure 1.1 Normal vaginal discharge.
A– Wet mount microscopy (400x, phase contrast) B– Gram stain (1000x, oil immersion).
1
THE NORMAL DISCHARGE
(alphabetical order)
Švitrigailė Grincevičienė
Iara Linhares
José Martinez de Oliveira
22
Consequently, the pH value of the vaginal fluid results from the mixture of those from the
cervix, vagina, and semen if the woman recently had unprotected intercourse.
Vaginal lubrication depends on the amount and quality of the transudate from the arterial
circulation. Its amount represents the predominant force of pressure from the vessels and its
counterpart, the epithelial pressure. The interstitial fluid passes to the cavity and, accordin-
gly to its rheological properties, spreads and covers the entire vagina.8
Among the substances contained in the vaginal fluid, there are some volatile ones that
cause its particular odor, like acetic acid or cresol.9 As lactic acid is the predominant product
of lactobacilli metabolism and one of the main acidifying agents of the vaginal fluid, it is
expected that lactic odor will be representative of normality. Nevertheless, the absence of
perception of odor is also considered normal.
The future will show whether among the more than a thousand proteins present in the vag-
inal fluid there are some that will be useful in differentiating normal from non-normal fluid.10
The components of the vaginal discharge can be categorized into: host cells, microorganisms,
and soluble components. All three create the color, odor, viscosity, and amount of the fluid.
Host cell components
Host cell components include epithelial cells and leukocytes. Multiple layers of stratified
squamous epithelium line the vagina. Epithelial cells are continually shedding into the vagi-
nal lumen.11 Healthy vaginal fluid predominantly contains cells of the vaginal superficial lay-
er and ectocervical epithelium, because they are not held together by tight junctions.11, 12
It takes approximately 96 hours for epithelial cells to transit from the basal layer to the api-
cal one. One cell layer is lost every 4 hours; however the rate of desquamation varies with
intercourse, vaginal product use, and hormonal status. Disintegration of the epithelial cells
is a major source of glycogen – the main substrate for lactobacilli. The junctions between
epithelial cells are weaker compared with skin ones and do not keratinize or form a lipid
envelope. The permeability is increased for all components, including for leucocytes.11
Leucocytes are also part of the cells present in a healthy vagina, with T-lymphocytes com-
prising the dominant type.13 Granulocytes, B-lymphocytes and macrophages are also de-
tectable, but are minor components.13, 14 The composition of leucocytes differs from blood
indicating that they are not a result of “passive” infiltration through the tissue.13 However,
natural killers in the vagina resemble those in the blood stream, contrary to those identified
in the upper genital tract, and play an important role in limiting viral infections.14 Cervical
ectropion (a normal developmental finding in which the squamocolumnar junction is lo-
cated in the ectocervix), when prominent, can cause discharge with leucocytes.15 Transient
presence of leucocytes from a partner’s sperm can also occur and be a source of disease
transmission. This must be considered when interpreting the presence of inflammatory cells
in wet mount microscopy, therefore making it important to know the time elapsed since
last intercourse.
23
Soluble components and mucus
Soluble components include secretions of glandular cells from the cervix and upper repro-
ductive tract, the remains of desquamated vaginal epithelial cells, microorganisms´ metab-
olites, as well as multiple products transduced to the vagina from the systemic circulation.
Cervical mucus coats the vaginal surface and forms a protective barrier. The composition of
vaginal mucus includes 2 to 5% mucin glycoproteins and 1% of other secreted agents such
as antibodies, antibacterial proteins, and peptides. Secreted mucins form a viscoelastic gel.
Carbohydrates in the fluid are responsible for more than 80% of the mucinal weight and
consist of N-acetyl-glucosamine, N-acetyl-galactosamine, galactose, fucose, and sialic acid.16
Estrogens and progesterone influence the vaginal pH, viscosity and protein content.16, 17
The subsequent release of glycogen from the shedded cells and its breakdown by vaginal
amylase provide a major source of nutrients that are utilized by lactobacilli.18 Data show that
vaginal amylase is produced by both the host and various bacteria (i.e. Lactobacillus crispatus, L.
iners, Bifidobacterium lacrimalis, and B. vaginale).19 It degrades glycogen to monosaccharides,
disaccharides, and trisaccharides, making it available for the lactobacilli’s metabolism.19, 20
Vaginal concentrations of neutrophil gelatinase-associated lipocalin, matrix metallopro-
teinase 8, and D- and L-lactic acid levels have been reported.18 Vaginal epithelial cells are a
component of the innate immune system and release antimicrobial compounds, as well as
cytokines that activate antigen-specific immunity, which are part of the soluble media of
vaginal discharge.21 The concentration of immune-active cells and compounds in the vagina
varies with the composition of the vaginal microbiota. Levels are typically lower when L.
crispatus is the dominant bacterium.22
Microorganisms
Bacteria, fungi, viruses, archaea, and protozoans are present in the vaginal fluids.23 The di-
verse saccharolytic population, mainly composed of lactobacilli, are often referred to as
Döderlein bacilli and are the most common acidifying organisms of the vaginal milieu.
The following is a general description of the vaginal microbiota and other components of the
vagina that are typically present in healthy reproductive age women. It must be acknowledged,
however, that due to variations in genetics, physiological factors and environmental exposures
it is difficult to define the “normal” vaginal environment that encompasses all healthy women.23
Bacteria
Microbiota release metabolites and degrade macronutrients. The lexicon describing differ-
ent aspects of the microbiome has been clarified by Verstraelen et al..23 In the majority of
women, one of four species of the genus Lactobacillus is numerically dominant in the va-
gina: L. crispatus, L. iners, L. jensenii or L. gasseri.24, 25 Lactobacilli produce lactic acid and reg-
ulate pH, modulate local immunity and release bacteriocins.26 The reason why usually only
one of these species of lactobacilli becomes predominant in a specific woman remains un-
determined. The most often cited classification of vaginal microbiomes is the one established by
Ravel et al. in 2011, which divides it into five community state types (CSTs).24 Four of the CSTs are
dominated by lactobacilli: CST-I, CST-II, CST-III, and CST-V, in which the predominant species are L.
24
crispatus, L. gasseri, L. iners, and L. jensenii, respectively.23, 24 CST-IV is characterized by high bacterial
diversity and is usually considered a “risky” community state in the scientific literature.23, 27
When lactobacilli are not numerically abundant, the most frequent alternatives are domi-
nance by Gardnerella spp. or a situation in which no bacterium constitutes over 50% of the
total bacterial species identified and, instead, there is a mixture of variable composition of
multiple species of anaerobic and facultative bacteria. Dominance by L. crispatus, L. jensenii
and L. gasseri has historically been associated with vaginal health, while the predominance
of L. iners or diversity of bacteria are associated with vaginal dysbiosis.28, 29 It must be noted
that the majority of women, in whom the latter microbes predominate, are asymptomatic. L.
iners is described both as having superior adaptation capacities due to resistance to hydrogen
peroxide and tolerance to environmental fluctuation (pH, menstrual bleeding, mucus concen-
tration, infection, hormones) as well as contributing to bacterial vaginosis (BV) through secre-
tion of inerolysin.29, 30 Long term health of women and their offspring, rather than only absence of
symptoms, should be considered when evaluating the “normality” of the vaginal microbiome.23
Non-lactobacilli species are also present in the vagina and sometimes prevalent. This prev-
alence may vary according to the different stages of life and ethnical/racial factors.23 For
instance, Prevotella spp. or Sneathia spp. may dominate in neonates while Gardnerella spp.
and Bifidobacterium spp. may be encountered more often in post-menopausal women.31, 32
Leptotrichia amnionii and Fannyhessea (Atopobium) vaginae are more common among Afri-
can Americans.33 Molecular methods allow the detection of a huge variety of bacteria, but
the pathogenicity of the majority remains unknown. Mycoplasma spp. and Ureaplasma spp.
are examples of such.34 Detection of bacteria (with the exclusion of cases such as Chlamydia
trachomatis or Neisseria gonorrhoeae) does not define normality or abnormality of vaginal
discharge. For instance, L. iners is present in both women with and without vaginal dys-
biosis.34 Moreover, detection of Gardnerella spp. is not evidence of dysbiosis.23 Abundance
and diversity of bacteria, fluctuation of CST during menstrual cycle and life span has been
described in the scientific literature, showing both instability of the microbiome and, in par-
ticular cases, stability of fluctuation patterns.3, 32, 35, 36
Viruses
Recent studies have added to the list of identified viruses in the vagina of healthy women.
Two broad group of viruses have been described: bacteriophages (viruses that infect bacte-
ria) and other eucaryotic viruses.37, 38
The dominant bacteriophages belong to the Caudovirales order, especially members of the
Myoviridae, Siphoviridae and Podoviridae families.37, 39 However, this dominance may be due
to reporting bias. Other bacteriophage families are Herelleviridae and Ackermannviridae,
Inoviridae, Microviridae, Lipothrixviridae, Tectiviridae and Plasmaviridae. Bacteriophages
play an important role in vaginal mucosa inflammation by inducing an inflammatory type-1
interferon response.37, 40 Da Costa et al. in 2021, evaluating samples from 107 pregnant wom-
en, described the prevalence of phage species as: Bacillus spp. phages in 43.6% of women,
Escherichia spp. phages in 40.9%, Staphylococcus spp. phages in 36.4%, Gokushovirus in
30.0% and Lactobacillus spp. phages in 26.4%.41
25
Among eucaryotic viruses, the Papillomaviridae predominate, followed by other double
stranded deoxyribonucleic acid (DNA) viruses including Polyomaviridae, Herpesviridae,
Genomoviridae, Adenoviridae, and Poxviridae and single-stranded DNA viruses such as
those of the Anelloviridae family.37, 42-44 DNA viruses such as Herpesviridae, Papillomaviridae,
Polyomaviridae, Poxviridae and Adenoviridae families are considered pathogenic as well
as ribonucleic acid (RNA) viruses such as human immunodeficiency virus (HIV) and Zika.37
While the presence of an apparently non-pathogenic herpesvirus has been shown to in-
crease immune sensitivity to endogenous vaginal bacteria in mice, the influence of DNA
viruses in the vagina on bacterial composition has not been reported.45
Yeasts
The colonization of the vagina of healthy women by Candida spp., especially C. albicans,
occurs frequently.46, 47 Immune system components present in the vagina of healthy women
are usually able to prevent the conversion of C. albicans from a benign colonization yeast
morphology to an invasive hyphal form and also to limit its capacity of replication.48, 49
In most healthy women, the presence of a low level of C. albicans has no apparent influ-
ence in the composition of the vaginal bacteria.50, 51 Even more, some researchers hypoth-
esized potential benefits, such as inhibition of E. coli in colonization cases.52 C. glabrata, a
non-hyphae forming yeast, is the second most common fungus that can be isolated from
the vagina.53, 54 In some women, especially if diabetic type I, this yeast may be responsible
for symptoms.53, 55 It seems that not the specific species, but rather the interaction of the
pathogen (i.e. candidalysin secretion), host (i.e. inflammatory cytokines) and environment
(i.e. microbiome, hormones, sexual activity) may lead to symptoms.56
1.2
Normal vaginal discharge variations during the menstrual cycle
The amount of cervical secretion decreases over the menstrual cycle, while the vaginal tran-
sudate increases. Approximately 1-3 mL of discharge are produced daily close to the men-
struation. Its consistency and distribution remain stable during the whole cycle.3
Cyclic fluctuations of estrogen and progesterone have impact in the genital mucosal immune
milieu.57, 58 Abundance of proteins differ in follicular, ovulatory and luteal phases. The luteal
and follicular phases are associated with higher activation of neutrophils/leukocytes and
cell migration pathways. During the ovulatory phase the antimicrobial and wound-healing
pathways are increased, while that of inflammatory cytokines is reduced. The microbiome
modulates luteal phase-dependent alterations to the vaginal mucosal proteome, leading to
a mucosal barrier function decrease in that phase.57
The proportion of leucocytes observed in wet mount microscopy slides tends to be stable during
the menstrual cycle, and not correlated with white blood cell count.3 The neutrophil, antimicrobi-
al, and tissue homeostasis pathways may be significantly changed during the menstrual phase.57
Vaginal microbiome of healthy woman can be stable in each menstrual cycle or fluctuate.36,
59, 60 Menstruation dramatically changes the microbiome composition. Two thirds of women
26
have high amounts of lactobacilli at the beginning of menses.3 However, the abundance of
L. crispatus decreases more than 100 fold, while the proportion of L. iners increases.61 Vaginal
microbiome diversity is much higher compared to that seen during the follicular or luteal
phase.59 Heavy growth of non-lactobacilli species is observed in the last days of the men-
strual cycle, namely Gardnerella spp., P. bivia, and F. vaginae.3, 61 The process is associated
with vaginal pH increase. After the menses, the abundance of group B streptococci, E. coli,
Gardnerella spp. and Prevotela spp. slightly decreases, while the amount of C. albicans, Bac-
teroides fragilis and Ureaplasma urealiticum increases.3
Menstruation patterns also influence the microbiome, with heavier flow associated with
higher abundance of Propionibacterium acnes in cervical samples. Regular periods are nega-
tively correlated with L. vaginalis, L. johnsonii, and Weissella spp., and lower levels of plasma
metabolites (androstenedione, testosterone, and serum low density lipoprotein).62
Contraception use may lead to microbiome changes in some women, while in others the
impact is minimal.63, 64 The use of combined oral contraceptives or levonorgestrel intrauter-
ine systems (LVN-IUS) does not seem to have a deleterious effect in the vaginal microbiome
composition or diversity.59, 65, 66 In fact, some data suggest that the use of sex hormones for
contraception promotes eubiosis; this effect is not clear for progestin-only contraceptives.64, 65
Shifts in the vaginal microbiome are mostly observed in women not using hormonal contracep-
tion and is less noticed for LVN-IUS, even after excluding women without menstrual bleeding.59
In conclusion, the vaginal microbiome is very sensitive to menstrual cycle and to circulating
hormones. However, it is still unknown why some women have a stable microbiome and in
others microbial diversity and abundance changes very rapidly.
1.3
Normal vaginal discharge in physiological estrogen deficiency
(pre-menstrual girls, postpartum, and post-menopausal women)
Normal vaginal discharge varies during the different stages of life, as the vaginal microbi-
ome is a dynamic system, depending on the host (inflammatory factors), the environment
and on the adaptation of vaginal bacteria (domination of species in an ecological niche and
their metabolites) to the environment (hormonal factors, sexual activity).23, 56
During a woman´s life cycle there are three physiologic hypotrophic vaginal periods: 1) during in-
fancy, 2) during the post-partum and early lactation period and 3) after established menopause.67
After birth, the newborn’s vagina is colonized by maternal lactobacilli.31 During the first
month of life, the vaginal mucosa is under the influence of maternal estrogens.68 Due to the
consequent high levels of vaginal glycogen, the amount of lactobacilli is high, the vaginal
pH is low and discharge can be perceived.69 Lactobacillus spp., Prevotella spp., or Sneathia
spp. can be detected amongst the vaginal microbiota of newborns and tend to be similar to
the mother’s vaginal or skin microbiome, in case of vaginal birth or cesarian section, respec-
tively.23, 31 After this short period, due to vaginal mucosa de-estrogenization, the vaginal pH
increases, as well as the diversity of the vaginal microbiota.
27
This hypoestrogenic environment, and increased pH, is maintained until puberty.23 The vaginal mi-
crobiota in prepubertal girls is abundant in non-lactobacilli species.68, 70 The discharge in girls is usu-
ally scarce. As the circulating levels of sex hormones gradually increase, the lactobacilli-deficient
microbiota gradually shifts toward lactobacilli dominance.23, 71, 72 Still, it remains unclear which
CST and how stable the microbiome will be for the adolescent girl. The correlation of the mi-
crobiome composition between mothers and their daughters remains unestablished.71, 73-75
The second period of hypoestrogenic state is the post-partum period. The number of lac-
tobacilli decreases dramatically.73 The process can be associated with a decreased estrogen
level after delivery and during breastfeeding. (Figure 1.2) Another theory claims that alka-
line lochial discharge impedes Lactobacillus spp. growth.76, 77 A reduced amount of lactoba-
cilli is followed by an increase in the proportion of Clostridia spp., Bacteroidia spp., Prevotella
spp., Finegodia magna, Streptococcus anginosus and other rare species.77 These communities
are similar to gut microbiome post-partum.73
Some authors state that an increased vaginal pH without symptoms of BV or other form of
dysbiosis is an indicator of menopause.78 The correlation between estradiol level and pH is
well established.79 Nevertheless, the process is more complex: during perimenopause, the
level of circulating hormones decrease, reducing lactobacilli dominance and increasing the
diversity of other species.23 After the menopause, the vaginal mucosa again turns into a
de-estrogenized state that leads to thinning of the epithelium. Due to the decrease in glyco-
gen, and consequent lactobacilli reduction and elevated pH, the diversity of species present
increases.68 Gliniewizc et al. described six ecological clusters for postmenopausal women in
accordance to dominant species: L. crispatus, L. iners, L. gasseri, Gardnerella spp., Bifidobacte-
rium spp. and co-dominance by several taxa.32 Previous studies that have characterized the
vaginal microbiota of postmenopausal
women have reported associations
between different taxonomic composi-
tions and vaginal symptoms. For exam-
ple, Brotman et al. found that a vaginal
microbiome dominated by Fannyhes-
sea (Atopobium) spp. (CST IVB) was
associated with mild or moderate atro-
phy, while dominance by Streptococcus
spp. and Prevotella spp. (CST IVA) was
associated with severe symptoms,80
and Shen et al. reported that post-
menopausal women with L. gasseri/L.
jensenii dominated communities had
less vaginal dryness compared to other
women.81
The importance of the microbiome in re-
lation to urinary incontinence and other
urinary symptoms remains unclear.23, 81
Figure 1.2 Wet mount microscopy of a vaginal sample
collected from a breastfeeding woman (400x, phase contrast).
Note the absence of lactobacilli and of intermediate and
superficial epithelial cells.
28
1.4
Normal vaginal discharge during pregnancy
The physiological state of pregnancy is led by hormonal changes associated with immune
modulation, behavioral changes, physio-chemical changes in the mucosa, and changes in
the genital tract. These factors modulate the vaginal microbiome, that is different from that
of non-pregnant women.73 The amount of vaginal discharge during pregnancy increases,
presumably due to increased transudation associated with vaginal congestion. The fluid is
usually white or yellowish and creamy.
Lactobacilli usually dominate during pregnancy.61, 77 The community becomes more stable
and less diverse with the progression of pregnancy, an effect probably mediated by the in-
crease in estrogen levels.23, 73, 82 Physiological changes increase deposition of glycogen that
will be broken down into lactic acid and consequently lead to a decrease of pH.35, 61, 73, 83, 84
Upregulation of pro-inflammatory processes and D-lactic acid induce autophagy of bacte-
ria.61, 85 Increased ratio of D- to L-lactic acid promotes the expression of vaginal extracellular
matrix metalloproteinase inducer, which in turn can activate matrix metalloproteinase-8
and subsequently alter the cervical integrity.77, 86 Lactic acid enhances the release of IL-1β
and IL-8 from vaginal epithelial cells, suggesting a synergistic relationship between inflam-
matory activation in the host and microbial composition.77 These mechanisms decrease the
probability of aerobic vaginitis during pregnancy, inhibit E. coli and stimulate L. crispatus domi-
nance.61, 87, 88 Studies show that, besides L. crispatus dominance among Caucasian and Asian preg-
nant women, L. jensenii and L. gasseri dominated communities are also common. African Amer-
ican women vaginal microbiota is more likely to be dominated by L. iners during pregnancy.77
1.5
Factors contributing to variations in the composition
of the vaginal discharge
Stress
Stress is associated with higher risk of BV.89, 90 Stress initiates the release of cortisol and norep-
inephrine from the adrenal cortex. Cortisol affects estrogen level and has an inhibitory effect
on the maturation of the vaginal epithelial cells. As a consequence, due to the decrease in
the amount of glycogen, the proportion of lactobacilli is reduced, and thus less lactic acid is
produced.91-93 Several studies have shown that even when co-administered with estrogens,
cortisol inhibits glycogen deposition.89, 93 This change decreases the anti-inflammatory prop-
erties of lactobacilli products and potentiates proinflammatory response, leading to abun-
dance of facultative anaerobes or worsening symptoms of vulvovaginitis.91 Increase of nor-
epinephrine potentiates pro-inflammatory response and affects stability of the microbiota.94
Sexual activity
A study on sexual workers has shown that recent sexual activity is related with increased mi-
crobiome diversity.95 Certain sexual behaviors increase the probability of BV.60, 96-99 Identified
risk factors include: frequency, increased number of sex partners, unprotected penile-vaginal sex
or women with a female partner. On the other hand, condom use is a protective factor.93, 100, 101
29
Douching
Douching has been associated with changes in vaginal discharge, namely BV, in some women.
The washing of the vagina mechanically reduces the abundance of bacteria, including lac-
tobacilli.93 However, inhibition of bacteria is not always the main reason for disturbance due
to douching; according to Hesham et al., some products (i.e. vinegar) do not inhibit bacterial
growth.102 Nevertheless, there was an increased induction of vaginal epithelial cell death,
triggered by a proinflammatory response with elevation of IL-6, IL-1β when vinegar or io-
dine were used.93, 102 Douching inhibited E. coli growth. If more lactobacilli are present, less
epithelial cell death is observed.102 Some researchers observed that products, even without
effect on the vaginal pH, increase the diversity of anaerobic bacteria and promote occur-
rence of symptomatic candidiasis.103, 104 However, some studies have shown that stopping
vaginal douching in women with BV is not enough to restore a lactobacilli dominated microbi-
ota.105 After careful evaluation of cultural background, douching should be discouraged.
Smoking
Cigarette smoking is associated with BV and higher prevalence of CST-IV.106 This may be due
to promotion of the growth of Gardnerella spp. and Mobiluncus spp., rather than lactobacil-
li depletion.107 Biogenic amines, such as agmatine, cadaverine, putrescine, tryptamine and
tyramine, affecting the virulence of infective pathogens and contributing to vaginal mal-
odor, were higher among cigarette users.108
Diet
A number of studies have investigated the relationship between diet and vaginal microbiota
composition. The association between the vaginal microbiota and sugar, fiber, protein, or
fat intake remains unclear, despite some studies hypothesizing that a starch rich diet in-
creases vaginal glycogen and thus may have a positive effect on lactobacilli.109, 110 However,
a high glycemic load has been related with progression of BV.111 Also, Saraf et al. noted that
a high fat diet contributes to estrogen increase and consequently to vaginal glycogen lev-
el.112 Alternatively, Naggers et al. found that high fat intake was associated with severe BV.112
Vegetarian diets can be related to higher vaginal microbial diversity.109
1.6
Racial differences in the composition of the vaginal discharge
and vaginal microbiota
It is well established that the numerically dominant bacterium in the vagina varies with race.
L. crispatus, L. jensenii and L. gasseri are typically more prevalent in White and Asian women
than in Black and Hispanic women.113 In African American women, the most common pro-
file is dominance by L. iners, followed by Gardnerella spp., BV-associated bacteria (BVAB) 1,
and L. crispatus.114 A similar pattern has been observed among Hispanic women as well.113
Vaginal community composition in young Black women was related to glycogen levels, not
estradiol and psychosocial stress.113 Accordingly, the mean vaginal pH is also higher in these
30
groups.24, 115 Reasons for these variations in vaginal microbiota composition remain speculative.
While differences in vaginal microbiota composition across different ethnic or racial groups
have been reported, the association between ethnicity and vaginal microbiota composition
is likely to be confounded by other factors known to influence it, including douching, sexual
networks, cultural practices and other factors discussed above.
1.7
Summary and conclusions
Vaginal discharge can be both a sign and a symptom. The characteristics of the discharge
that were discussed earlier included amount, color, consistency, odor, and pH value. There is
no uniform pattern of normality for the vaginal discharge, as its characteristics change dur-
ing menstruation, in accordance with lifestyle, over time and have racial differences. In fact
and taking particularly into account the relevant role of sexual hormones in the physiology
of the vagina, different normal states are defined for the newborn, the prepubertal girl, the
non-pregnant woman during the reproductive age and the post menopause involutional
stage. In between, transitional stages must be considered: puberty, pregnancy, puerperium
and perimenopause. For decades, it was assumed almost dogmatically that the “healthy va-
gina” was necessarily dominated by lactobacilli. However, more recently it has been shown
that a significant number of asymptomatic women harbor a non-lactobacilli dominated mi-
crobiota.24
Molecular methods that enable comprehensive characterization of the vaginal microbiome,
metabolome and proteome, will provide further insight into the composition and function
of the vaginal microbiome in health and disease.
Recommendations
Recommendation Quality of evidence Strength of
recommendation
A perceived increased discharge is not necessarily pathological
and should not automatically prompt treatment. 4 C
There is no recommendation to treat a non-ideal vaginal
microbiota in the absence of symptoms. 5 D
The use of hormonal contraceptives promotes eubiosis. 4 C
References
1. Mitchell, H., Vaginal discharge--causes, diagnosis, and treatment. Bmj 2004, 328, (7451), 1306-8.
2. Rao, V. L.; Mahmood, T., Vaginal discharge. Obstetrics, Gynaecology & Reproductive Medicine 2020, 30, (1), 11-18.
3. Eschenbach, D. A.; Thwin, S. S.; Patton, D. L.; Hooton, T. M.; Stapleton, A. E.; Agnew, K.; Winter, C.; Meier, A.; Stamm, W. E.,
Influence of the normal menstrual cycle on vaginal tissue, discharge, and microflora. Clin Infect Dis 2000, 30, (6), 901-7.
4. Karasz, A.; Anderson, M., The vaginitis monologues: women’s experiences of vaginal complaints in a primary care
setting. Soc Sci Med 2003, 56, (5), 1013-21.
31
5. Anderson, M.; Karasz, A.; Friedland, S., Are vaginal symptoms ever normal? a review of the literature. MedGenMed
2004, 6, (4), 49.
6. Chatur vedi, S. K.; Chandra, P. S.; Issac, M. K.; Sudarshan, C. Y., Somatization misattributed to non-pathological vaginal
discharge. J Psychosom Res 1993, 37, (6), 575-9.
7. Owen, D. H.; Katz, D. F., A vaginal fluid simulant. Contraception 1999, 59, (2), 91-5.
8. Dubinsk aya, A.; Guthrie, T.; Anger, J. T.; Eilber, K. S.; Berman, J. R., Local Genital Arousal: Mechanisms for Vaginal Lubri-
cation. Current Sexual Health Reports 2021, 13, (2), 45-53.
9. Huggins, G. R.; Preti, G., Vaginal odors and secretions. Clin Obstet Gynecol 1981, 24, (2), 355-77.
10. Kim, Y. E.; K im, K.; Oh, H. B.; Lee, S. K.; Kang, D., Quantitative proteomic profiling of Cervicovaginal fluid from pregnant
women with term and preterm birth. Proteome Sci 2021, 19, (1), 3.
11. Anderson, D. J.; Marathe, J.; Pudney, J., The structure of the human vaginal stratum corneum and its role in immune
defense. Am J Reprod Immunol 2014, 71, (6), 618-23.
12. Donders, G. G., Definition and classification of abnormal vaginal flora. Best Pract Res Clin Obstet Gynaecol 2007, 21,
(3), 355-73.
13. Givan, A. L.; White, H. D.; Stern, J. E.; Colby, E.; Gosselin, E. J.; Guyre, P. M.; Wira, C. R., Flow cytometric analysis of leuko-
cytes in the human female reproductive tract: comparison of fallopian tube, uterus, cervix, and vagina. Am J Reprod
Immunol 1997, 38, (5), 350-9.
14. Monin, L.; Whettlock, E. M.; Male, V., Immune responses in the human female reproductive tract. Immunology 2020,
160, (2), 106-115.
15. Mancuso, A. C.; Ryan, G. L., Normal Vulvovaginal Health in Adolescents. J Pediatr Adolesc Gynecol 2015, 28, (3), 132-5.
16. Moncla, B. J.; Chappell, C. A.; Debo, B. M.; Meyn, L. A., The Effects of Hormones and Vaginal Microflora on the Glycome
of the Female Genital Tract: Cervical-Vaginal Fluid. PLoS One 2016, 11, (7), e0158687.
17. Chappell, C. A.; Rohan, L. C.; Moncla, B. J.; Wang, L.; Meyn, L. A.; Bunge, K.; Hillier, S. L., The effects of reproductive
hormones on the physical properties of cervicovaginal fluid. Am J Obstet Gynecol 2014, 211, (3), 226.e1-7.
18. Nasioudis, D.; Beghini, J.; Bongiovanni, A. M.; Giraldo, P. C.; Linhares, I. M.; Witkin, S. S., α-Amylase in Vaginal Fluid:
Association With Conditions Favorable to Dominance of Lactobacillus. Reprod Sci 2015, 22, (11), 1393-8.
19. Nunn, K. L.; Clair, G. C.; Adkins, J. N.; Engbrecht, K.; Fillmore, T.; Forney, L. J., Amylases in the Human Vagina. mSphere
2020, 5, (6).
20. Spear, G. T.; French, A. L.; Gilbert, D.; Zariffard, M. R.; Mirmonsef, P.; Sullivan, T. H.; Spear, W. W.; Landay, A.; Micci, S.; Lee,
B. H.; Hamaker, B. R., Human α-amylase present in lower-genital-tract mucosal fluid processes glycogen to support
vaginal colonization by Lactobacillus. J Infect Dis 2014, 210, (7), 1019-28.
21. Linhares, I. M.; Sisti, G.; Minis, E.; de Freitas, G. B.; Moron, A. F.; Witkin, S. S., Contribution of Epithelial Cells to Defense
Mechanisms in the Human Vagina. Curr Infect Dis Rep 2019, 21, (9), 30.
22. Dabee, S.; Barnabas, S. L.; Lennard, K. S.; Jaumdally, S. Z.; Gamieldien, H.; Balle, C.; Happel, A. U.; Murugan, B. D.; Wil-
liamson, A. L.; Mkhize, N.; Dietrich, J.; Lewis, D. A.; Chiodi, F.; Hope, T. J.; Shattock, R.; Gray, G.; Bekker, L. G.; Jaspan, H.
B.; Passmore, J. S., Defining characteristics of genital health in South African adolescent girls and young women at
high risk for HIV infection. PLoS One 2019, 14, (4), e0213975.
23. Verstraelen, H.; Vieira-Baptista, P.; De Seta, F.; Ventolini, G.; Lonnee-Hoffmann, R.; Lev-Sagie, A., The Vaginal Micro-
biome: I. Research Development, Lexicon, Defining “Normal” and the Dynamics Throughout Women’s Lives. J Low
Genit Tract Dis 2022, 26, (1), 73-78.
24. Ravel, J.; Gajer, P.; Abdo, Z.; Schneider, G. M.; Koenig, S. S.; McCulle, S. L.; Karlebach, S.; Gorle, R.; Russell, J.; Tacket, C.
O.; Brotman, R. M.; Davis, C. C.; Ault, K.; Peralta, L.; Forney, L. J., Vaginal microbiome of reproductive-age women. Proc
Natl Acad Sci U S A 2011, 108 Suppl 1, (Suppl 1), 4680-7.
25. Lamont, R. F.; Sobel, J. D.; Akins, R. A.; Hassan, S. S.; Chaiworapongsa, T.; Kusanovic, J. P.; Romero, R., The vaginal mi-
crobiome: new information about genital tract flora using molecular based techniques. Bjog 2011, 118, (5), 533-49.
26. Kovachev, S., Defence factors of vaginal lactobacilli. Crit Rev Microbiol 2018, 44, (1), 31-39.
27. McKinnon, L. R.; Achilles, S. L.; Bradshaw, C. S.; Burgener, A.; Crucitti, T.; Fredricks, D. N.; Jaspan, H. B.; Kaul, R.; Kaushic,
C.; Klatt, N.; Kwon, D. S.; Marrazzo, J. M.; Masson, L.; McClelland, R. S.; Ravel, J.; van de Wijgert, J.; Vodstrcil, L. A.; Tached-
jian, G., The Evolving Facets of Bacterial Vaginosis: Implications for HIV Transmission. AIDS Res Hum Retroviruses 2019,
35, (3), 219-228.
28. Han, Y.; Liu, Z.; Chen, T., Role of Vaginal Microbiota Dysbiosis in Gynecological Diseases and the Potential Interven-
tions. Front Microbiol 2021, 12, 643422.
29. Petrova, M. I.; Reid, G.; Vaneechoutte, M.; Lebeer, S., Lactobacillus iners: Friend or Foe? Trends Microbiol 2017, 25, (3), 182-191.
32
30. Macklaim, J. M.; Gloor, G. B.; Anukam, K. C.; Cribby, S.; Reid, G., At the crossroads of vaginal health and disease, the
genome sequence of Lactobacillus iners AB-1. Proc Natl Acad Sci U S A 2011, 108 Suppl 1, (Suppl 1), 4688-95.
31. Dominguez-Bello, M. G.; Costello, E. K.; Contreras, M.; Magris, M.; Hidalgo, G.; Fierer, N.; Knight, R., Delivery mode
shapes the acquisition and structure of the initial microbiota across multiple body habitats in newborns. Proc Natl
Acad Sci U S A 2010, 107, (26), 11971-5.
32. Gliniewicz, K.; Schneider, G. M.; Ridenhour, B. J.; Williams, C. J.; Song, Y.; Farage, M. A.; Miller, K.; Forney, L. J., Com-
parison of the Vaginal Microbiomes of Premenopausal and Postmenopausal Women. Front Microbiol 2019, 10, 193.
33. Srinivasan, S.; Hoffman, N. G.; Morgan, M. T.; Matsen, F. A.; Fiedler, T. L.; Hall, R. W.; Ross, F. J.; McCoy, C. O.; Bumgarner,
R.; Marrazzo, J. M.; Fredricks, D. N., Bacterial communities in women with bacterial vaginosis: high resolution phy-
logenetic analyses reveal relationships of microbiota to clinical criteria. PLoS One 2012, 7, (6), e37818.
34. Lev-Sagie, A.; De Seta, F.; Verstraelen, H.; Ventolini, G.; Lonnee-Hoffmann, R.; Vieira-Baptista, P., The Vaginal Microbi-
ome: II. Vaginal Dysbiotic Conditions. J Low Genit Tract Dis 2022, 26, (1), 79-84.
35. Romero, R.; Hassan, S. S.; Gajer, P.; Tarca, A. L.; Fadrosh, D. W.; Nikita, L.; Galuppi, M.; Lamont, R. F.; Chaemsaithong, P.;
Miranda, J.; Chaiworapongsa, T.; Ravel, J., The composition and stability of the vaginal microbiota of normal preg-
nant women is different from that of non-pregnant women. Microbiome 2014, 2, (1), 4.
36. Chaban, B.; Links, M. G.; Jayaprakash, T. P.; Wagner, E. C.; Bourque, D. K.; Lohn, Z.; Albert, A. Y.; van Schalkwyk, J.; Reid,
G.; Hemmingsen, S. M.; Hill, J. E.; Money, D. M., Characterization of the vaginal microbiota of healthy Canadian wom-
en through the menstrual cycle. Microbiome 2014, 2, 23.
37. Madere, F. S.; Monaco, C. L., The female reproductive tract virome: understanding the dynamic role of viruses in
gynecological health and disease. Curr Opin Virol 2022, 52, 15-23.
38. Happel, A. U.; Balle, C.; Maust, B. S.; Konstantinus, I. N.; Gill, K.; Bekker, L. G.; Froissart, R.; Passmore, J. A.; Karaoz, U.;
Varsani, A.; Jaspan, H., Presence and Persistence of Putative Lytic and Temperate Bacteriophages in Vaginal Metage-
nomes from South African Adolescents. Viruses 2021, 13, (12).
39. Jakobsen, R. R.; Haahr, T.; Humaidan, P.; Jensen, J. S.; Kot, W. P.; Castro-Mejia, J. L.; Deng, L.; Leser, T. D.; Nielsen, D. S.,
Characterization of the Vaginal DNA Virome in Health and Dysbiosis. Viruses 2020, 12, (10).
40. Van Belleghem, J. D.; Dąbrowska, K.; Vaneechoutte, M.; Barr, J. J.; Bollyky, P. L., Interactions between Bacteriophage,
Bacteria, and the Mammalian Immune System. Viruses 2018, 11, (1).
41. Da Costa, A. C.; Moron, A. F.; Forney, L. J.; Linhares, I. M.; Sabino, E.; Costa, S. F.; Mendes-Correa, M. C.; Witkin, S. S.,
Identification of bacteriophages in the vagina of pregnant women: a descriptive study. Bjog 2021, 128, (6), 976-982.
42. Wylie, K. M.; Wylie, T. N.; Cahill, A. G.; Macones, G. A.; Tuuli, M. G.; Stout, M. J., The vaginal eukaryotic DNA virome and
preterm birth. Am J Obstet Gynecol 2018, 219, (2), 189.e1-189.e12.
43. Wylie, K. M.; Mihindukulasuriya, K. A.; Zhou, Y.; Sodergren, E.; Storch, G. A.; Weinstock, G. M., Metagenomic analysis of
double-stranded DNA viruses in healthy adults. BMC Biol 2014, 12, 71.
44. Eskew, A. M.; Stout, M. J.; Bedrick, B. S.; Riley, J. K.; Omurtag, K. R.; Jimenez, P. T.; Odem, R. R.; Ratts, V. S.; Keller, S. L.;
Jungheim, E. S.; Wylie, K. M., Association of the eukaryotic vaginal virome with prophylactic antibiotic exposure and
reproductive outcomes in a subfertile population undergoing in vitro fertilisation: a prospective exploratory study.
Bjog 2020, 127, (2), 208-216.
45. Cardenas, I.; Mor, G.; Aldo, P.; Lang, S. M.; Stabach, P.; Sharp, A.; Romero, R.; Mazaki-Tovi, S.; Gervasi, M.; Means, R.
E., Placental viral infection sensitizes to endotoxin-induced pre-term labor: a double hit hypothesis. Am J Reprod
Immunol 2011, 65, (2), 110-7.
46. Sobel, J. D., Recurrent vulvovaginal candidiasis. Am J Obstet Gynecol 2016, 214, (1), 15-21.
47. Sobel, J. D., Vulvovaginal candidosis. Lancet 2007, 369, (9577), 1961-71.
48. Verma, A.; Gaffen, S. L.; Swidergall, M., Innate Immunity to Mucosal Candida Infections. J Fungi (Basel) 2017, 3, (4).
49. Rosati, D.; Bruno, M.; Jaeger, M.; Ten Oever, J.; Netea, M. G., Recurrent Vulvovaginal Candidiasis: An Immunological
Perspective. Microorganisms 2020, 8, (2).
50. Brown, S. E.; Schwartz, J. A.; Robinson, C. K.; O Hanlon, D. E.; Bradford, L. L.; He, X.; Mark, K. S.; Bruno, V. M.; Ravel, J.;
Brotman, R. M., The Vaginal Microbiota and Behavioral Factors Associated With Genital Candida albicans Detection
in Reproductive-Age Women. Sex Transm Dis 2019, 46, (11), 753-758.
51. D’Enfert, C.; Kaune, A. K.; Alaban, L. R.; Chakraborty, S.; Cole, N.; Delavy, M.; Kosmala, D.; Marsaux, B.; Fróis-Martins, R.;
Morelli, M.; Rosati, D.; Valentine, M.; Xie, Z.; Emritloll, Y.; Warn, P. A.; Bequet, F.; Bougnoux, M. E.; Bornes, S.; Gresnigt,
M. S.; Hube, B.; Jacobsen, I. D.; Legrand, M.; Leibundgut-Landmann, S.; Manichanh, C.; Munro, C. A.; Netea, M. G.;
Queiroz, K.; Roget, K.; Thomas, V.; Thoral, C.; Van den Abbeele, P.; Walker, A. W.; Brown, A. J. P., The impact of the Fun-
gus-Host-Microbiota interplay upon Candida albicans infections: current knowledge and new perspectives. FEMS
Microbiol Rev 2021, 45, (3).
33
52. De Seta, F.; Lonnee-Hoffmann, R.; Campisciano, G.; Comar, M.; Verstraelen, H.; Vieira-Baptista, P.; Ventolini, G.;
Lev-Sagie, A., The Vaginal Microbiome: III. The Vaginal Microbiome in Various Urogenital Disorders. J Low Genit Tract
Dis 2022, 26, (1), 85-92.
53. Kennedy, M. A.; Sobel, J. D., Vulvovaginal Candidiasis Caused by Non-albicans Candida Species: New Insights. Curr
Infect Dis Rep 2010, 12, (6), 465-70.
54. Powell, A. M.; Gracely, E.; Nyirjesy, P., Non-albicans Candida Vulvovaginitis: Treatment Experience at a Tertiary Care
Vaginitis Center. J Low Genit Tract Dis 2016, 20, (1), 85-9.
55. Ray, D.; Goswami, R.; Banerjee, U.; Dadhwal, V.; Goswami, D.; Mandal, P.; Sreenivas, V.; Kochupillai, N., Prevalence
of Candida glabrata and its response to boric acid vaginal suppositories in comparison with oral fluconazole in
patients with diabetes and vulvovaginal candidiasis. Diabetes Care 2007, 30, (2), 312-7.
56. Willems, H. M. E.; Ahmed, S. S.; Liu, J.; Xu, Z.; Peters, B. M., Vulvovaginal Candidiasis: A Current Understanding and
Burning Questions. J Fungi (Basel) 2020, 6, (1).
57. Bradley, F.; Birse, K.; Hasselrot, K.; Noël-Romas, L.; Introini, A.; Wefer, H.; Seifert, M.; Engstrand, L.; Tjernlund, A.; Brolid-
en, K.; Burgener, A. D., The vaginal microbiome amplifies sex hormone-associated cyclic changes in cervicovaginal
inflammation and epithelial barrier disruption. Am J Reprod Immunol 2018, 80, (1), e12863.
58. Wira, C. R.; Rodriguez-Garcia, M.; Patel, M. V., The role of sex hormones in immune protection of the female reproduc-
tive tract. Nat Rev Immunol 2015, 15, (4), 217-30.
59. Krog, M. C.; Hugerth, L. W.; Fransson, E.; Bashir, Z.; Nyboe Andersen, A.; Edfeldt, G.; Engstrand, L.; Schuppe-Koistinen,
I.; Nielsen, H. S., The healthy female microbiome across body sites: effect of hormonal contraceptives and the men-
strual cycle. Hum Reprod 2022, 37, (7), 1525-1543.
60. Gajer, P.; Brotman, R. M.; Bai, G.; Sakamoto, J.; Schütte, U. M.; Zhong, X.; Koenig, S. S.; Fu, L.; Ma, Z. S.; Zhou, X.; Abdo,
Z.; Forney, L. J.; Ravel, J., Temporal dynamics of the human vaginal microbiota. Sci Transl Med 2012, 4, (132), 132ra52.
61. Amabebe, E.; Anumba, D. O. C., The Vaginal Microenvironment: The Physiologic Role of Lactobacilli. Front Med (Lau-
sanne) 2018, 5, 181.
62. Jie, Z.; Chen, C.; Hao, L.; Li, F.; Song, L.; Zhang, X.; Zhu, J.; Tian, L.; Tong, X.; Cai, K.; Zhang, Z.; Ju, Y.; Yu, X.; Li, Y.; Zhou, H.;
Lu, H.; Qiu, X.; Li, Q.; Liao, Y.; Zhou, D.; Lian, H.; Zuo, Y.; Chen, X.; Rao, W.; Ren, Y.; Wang, Y.; Zi, J.; Wang, R.; Liu, N.; Wu, J.;
Zhang, W.; Liu, X.; Zong, Y.; Liu, W.; Xiao, L.; Hou, Y.; Xu, X.; Yang, H.; Wang, J.; Kristiansen, K.; Jia, H., Life History Record-
ed in the Vagino-cervical Microbiome Along with Multi-omics. Genomics Proteomics Bioinformatics 2021.
63. Balle, C.; Konstantinus, I. N.; Jaumdally, S. Z.; Havyarimana, E.; Lennard, K.; Esra, R.; Barnabas, S. L.; Happel, A. U.;
Moodie, Z.; Gill, K.; Pidwell, T.; Karaoz, U.; Brodie, E.; Maseko, V.; Gamieldien, H.; Bosinger, S. E.; Myer, L.; Bekker, L.
G.; Passmore, J. S.; Jaspan, H. B., Hormonal contraception alters vaginal microbiota and cytokines in South African
adolescents in a randomized trial. Nat Commun 2020, 11, (1), 5578.
64. Bastianelli, C.; Farris, M.; Bianchi, P.; Benagiano, G., The effect of different contraceptive methods on the vaginal
microbiome. Expert Rev Clin Pharmacol 2021, 14, (7), 821-836.
65. Ratten, L. K.; Plummer, E. L.; Bradshaw, C. S.; Fairley, C. K.; Murray, G. L.; Garland, S. M.; Bateson, D.; Tachedjian, G.;
Masson, L.; Vodstrcil, L. A., The Effect of Exogenous Sex Steroids on the Vaginal Microbiota: A Systematic Review.
Front Cell Infect Microbiol 2021, 11, 732423.
66. Bassis, C. M.; Allsworth, J. E.; Wahl, H. N.; Sack, D. E.; Young, V. B.; Bell, J. D., Effects of intrauterine contraception on the
vaginal microbiota. Contraception 2017, 96, (3), 189-195.
67. Pérez-López, F. R.; Vieira-Baptista, P.; Phillips, N.; Cohen-Sacher, B.; Fialho, S.; Stockdale, C. K., Clinical manifestations
and evaluation of postmenopausal vulvovaginal atrophy. Gynecol Endocrinol 2021, 37, (8), 740-745.
68. Godha, K.; Tucker, K. M.; Biehl, C.; Archer, D. F.; Mirkin, S., Human vaginal pH and microbiota: an update. Gynecol
Endocrinol 2018, 34, (6), 451-455.
69. Farage, M.; Maibach, H., Lifetime changes in the vulva and vagina. Arch Gynecol Obstet 2006, 273, (4), 195-202.
70. Zuckerman, A.; Romano, M., Clinical Recommendation: Vulvovaginitis. J Pediatr Adolesc Gynecol 2016, 29, (6), 673-679.
71. Hickey, R. J.; Zhou, X.; Settles, M. L.; Erb, J.; Malone, K.; Hansmann, M. A.; Shew, M. L.; Van Der Pol, B.; Fortenberry, J.
D.; Forney, L. J., Vaginal microbiota of adolescent girls prior to the onset of menarche resemble those of reproduc-
tive-age women. mBio 2015, 6, (2).
72. Biro, F. M.; Pinney, S. M.; Huang, B.; Baker, E. R.; Walt Chandler, D.; Dorn, L. D., Hormone changes in peripubertal girls.
J Clin Endocrinol Metab 2014, 99, (10), 3829-35.
73. Gupta, P.; Singh, M. P.; Goyal, K., Diversity of Vaginal Microbiome in Pregnancy: Deciphering the Obscurity. Front
Public Health 2020, 8, 326.
74. Rasmussen, M. A.; Thorsen, J.; Dominguez-Bello, M. G.; Blaser, M. J.; Mortensen, M. S.; Brejnrod, A. D.; Shah, S. A.;
Hjelmsø, M. H.; Lehtimäki, J.; Tr ivedi, U.; Bisgaard, H.; Sørensen, S. J.; Stokholm, J., Ecological succession in the vaginal
microbiota during pregnancy and birth. The ISME Journal 2020, 14, (9), 2325-2335.
34
75. France, M. T.; Brown, S. E.; Rompalo, A. M.; Brotman, R. M.; Ravel, J., Identification of shared bacterial strains in the
vaginal microbiota of related and unrelated reproductive-age mothers and daughters using genome-resolved
metagenomics. PLOS ONE 2022, 17, (10), e0275908.
76. Odogwu, N. M.; Onebunne, C. A.; Chen, J.; Ayeni, F. A.; Walther-Antonio, M. R. S.; Olayemi, O. O.; Chia, N.; Omigbodun, A.
O., Lactobacillus crispatus thrives in pregnancy hormonal milieu in a Nigerian patient cohort. Sci Rep 2021, 11, (1), 18152.
77. MacIntyre, D. A.; Chandiramani, M.; Lee, Y. S.; Kindinger, L.; Smith, A.; Angelopoulos, N.; Lehne, B.; Arulkumaran, S.;
Brown, R.; Teoh, T. G.; Holmes, E.; Nicoholson, J. K.; Marchesi, J. R.; Bennett, P. R., The vaginal microbiome during
pregnancy and the postpartum period in a European population. Sci Rep 2015, 5, 8988.
78. Lehtoranta, L.; Ala-Jaakkola, R.; Laitila, A.; Maukonen, J., Healthy Vaginal Microbiota and Influence of Probiotics Across
the Female Life Span. Front Microbiol 2022, 13, 819958.
79. Caillouette, J. C.; Sharp, C. F., Jr.; Zimmerman, G. J.; Roy, S., Vaginal pH as a marker for bacterial pathogens and men-
opausal status. Am J Obstet Gynecol 1997, 176, (6), 1270-5; discussion 1275-7.
80. Brotman, R. M.; Shardell, M. D.; Gajer, P.; Fadrosh, D.; Chang, K.; Silver, M. I.; Viscidi, R. P.; Burke, A. E.; Ravel, J.; Gravitt, P. E., Associa-
tion between the vaginal microbiota, menopause status, and signs of vulvovaginal atrophy. Menopause 2014, 21, (5), 450-8.
81. Shen, J.; Song, N.; Williams, C. J.; Brown, C. J.; Yan, Z.; Xu, C.; Forney, L. J., Effects of low dose estrogen therapy on the
vaginal microbiomes of women with atrophic vaginitis. Sci Rep 2016, 6, 24380.
82. Walther-António, M. R.; Jeraldo, P.; Berg Miller, M. E.; Yeoman, C. J.; Nelson, K. E.; Wilson, B. A.; White, B. A.; Chia, N.;
Creedon, D. J., Pregnancy’s stronghold on the vaginal microbiome. PLoS One 2014, 9, (6), e98514.
83. Smith, S. B.; Ravel, J., The vaginal microbiota, host defence and reproductive physiology. J Physiol 2017, 595, (2), 451-463.
84. Hay, P., Vaginal discharge. Medicine 2018, 46, (6), 319-324.
85. Ramos Bde, A.; Kanninen, T. T.; Sisti, G.; Witkin, S. S., Microorganisms in the female genital tract during pregnancy:
tolerance versus pathogenesis. Am J Reprod Immunol 2015, 73, (5), 383-9.
86. Witkin, S. S.; Mendes-Soares, H.; Linhares, I. M.; Jayaram, A.; Ledger, W. J.; Forney, L. J., Influence of vaginal bacteria
and D- and L-lactic acid isomers on vaginal extracellular matrix metalloproteinase inducer: implications for protec-
tion against upper genital tract infections. mBio 2013, 4, (4).
87. Donders, G.; Bellen, G.; Rezeberga, D., Aerobic vaginitis in pregnancy. Bjog 2011, 118, (10), 1163-70.
88. Vieira-Baptista, P.; Lima-Silva, J.; Pinto, C.; Saldanha, C.; Beires, J.; Martinez-de-Oliveira, J.; Donders, G., Bacterial vaginosis, aer-
obic vaginitis, vaginal inflammation and major Pap smear abnormalities. Eur J Clin Microbiol Infect Dis 2016, 35, (4), 657-64.
89. Nansel, T. R.; Riggs, M. A.; Yu, K. F.; Andrews, W. W.; Schwebke, J. R.; Klebanoff, M. A., The association of psychosocial
stress and bacterial vaginosis in a longitudinal cohort. Am J Obstet Gynecol 2006, 194, (2), 381-6.
90. Turpin, R.; Slopen, N.; Borgogna, J. C.; Yeoman, C. J.; He, X.; Miller, R. S.; Klebanoff, M. A.; Ravel, J.; Brotman, R. M.,
Perceived Stress and Molecular Bacterial Vaginosis in the National Institutes of Health Longitudinal Study of Vaginal
Flora. Am J Epidemiol 2021, 190, (11), 2374-2383.
91. Amabebe, E.; Anumba, D. O. C., Psychosocial Stress, Cortisol Levels, and Maintenance of Vaginal Health. Front Endo-
crinol (Lausanne) 2018, 9, 568.
92. Padgett, D. A.; Glaser, R., How stress influences the immune response. Trends Immunol 2003, 24, (8), 444-8.
93. Kwon, M. S.; Lee, H. K., Host and Microbiome Interplay Shapes the Vaginal Microenvironment. Front Immunol 2022,
13, 919728.
94. Witkin, S. S.; Linhares, I. M., Why do lactobacilli dominate the human vaginal microbiota? Bjog 2017, 124, (4), 606-611.
95. Sivro, A.; Mwatelah, R.; Kambaran, C.; Gebrebrhan, H.; Becker, M. G.; Ma, H.; Klatt, N. R.; Zevin, A. S.; King’ola, N.; Wambua,
S.; Gichangi, P.; Cheuk, E.; McLaren, P. J.; Mishra, S.; Becker, M.; McKinnon, L. R., Sex Work Is Associated With Increased
Vaginal Microbiome Diversity in Young Women From Mombasa, Kenya. J Acquir Immune Defic Syndr 2020, 85, (1), 79-87.
96. Ratten, L. K.; Plummer, E. L.; Murray, G. L.; Danielewski, J.; Fairley, C. K.; Garland, S. M.; Hocking, J. S.; Tachedjian, G.;
Chow, E.; Bradshaw, C. S.; Vodstrcil, L. A., Sex is associated with the persistence of non-optimal vaginal microbiota
following treatment for bacterial vaginosis: a prospective cohort study. Bjog 2021, 128, (4), 756-767.
97. Plummer, E. L.; Vodstrcil, L. A.; Fairley, C. K.; Tabrizi, S. N.; Garland, S. M.; Law, M. G.; Hocking, J. S.; Fethers, K. A.; Bulach,
D. M.; Murray, G. L.; Bradshaw, C. S., Sexual practices have a significant impact on the vaginal microbiota of women
who have sex with women. Sci Rep 2019, 9, (1), 19749.
98. Wessels, J. M.; Lajoie, J.; Vitali, D.; Omollo, K.; Kimani, J.; Oyugi, J.; Cheruiyot, J.; Kimani, M.; Mungai, J. N.; Akolo, M.;
Stearns, J. C.; Surette, M. G.; Fowke, K. R.; Kaushic, C., Association of high-risk sexual behaviour with diversity of the
vaginal microbiota and abundance of Lactobacillus. PLoS One 2017, 12, (11), e0187612.
99. Schwebke, J. R.; Richey, C. M.; Weiss, H. L., Correlation of behaviors with microbiological changes in vaginal flora. J
Infect Dis 1999, 180, (5), 1632-6.
35
100. Fethers, K. A.; Fairley, C. K.; Hocking, J. S.; Gurrin, L. C.; Bradshaw, C. S., Sexual risk factors and bacterial vaginosis: a
systematic review and meta-analysis. Clin Infect Dis 2008, 47, (11), 1426-35.
101. Vodstrcil, L. A.; Muzny, C. A.; Plummer, E. L.; Sobel, J. D.; Bradshaw, C. S., Bacterial vaginosis: drivers of recurrence and
challenges and opportunities in partner treatment. BMC Med 2021, 19, (1), 194.
102. Hesham, H.; Mitchell, A. J.; Bergerat, A.; Hung, K.; Mitchell, C. M., Impact of vaginal douching products on vaginal
Lactobacillus, Escherichia coli and epithelial immune responses. Sci Rep 2021, 11, (1), 23069.
103. Van der Veer, C.; Bruisten, S. M.; van Houdt, R.; Matser, A. A.; Tachedjian, G.; van de Wijgert, J.; de Vries, H. J. C.; van
der Helm, J. J., Effects of an over-the-counter lactic-acid containing intra-vaginal douching product on the vaginal
microbiota. BMC Microbiol 2019, 19, (1), 168.
104. Gondwe, T.; Ness, R.; Totten, P. A.; Astete, S.; Tang, G.; Gold, M. A.; Martin, D.; Haggerty, C. L., Novel bacterial vagino-
sis-associated organisms mediate the relationship between vaginal douching and pelvic inflammatory disease. Sex
Transm Infect 2020, 96, (6), 439-444.
105. Brown, S. E.; He, X.; Shardell, M. D.; Ravel, J.; Ghanem, K. G.; Zenilman, J. M.; Brotman, R. M., Douching cessation and
molecular bacterial vaginosis: a reanalysis of archived specimens. Sex Transm Infect 2022.
106. Brotman, R. M.; He, X.; Gajer, P.; Fadrosh, D.; Sharma, E.; Mongodin, E. F.; Ravel, J.; Glover, E. D.; Rath, J. M., Association
between cigarette smoking and the vaginal microbiota: a pilot study. BMC Infect Dis 2014, 14, 471.
107. Tužil, J.; Filková, B.; Malina, J.; Kerestes, J.; Doležal, T., Smoking in women with chronic vaginal discomfort is not asso-
ciated with decreased abundance of Lactobacillus spp. but promotes Mobiluncus and Gardnerella spp. overgrowth
- secondary analysis of trial data including microbio-me analysis. Ceska Gynekol 2021, 86, (1), 22-29.
108. Nelson, T. M.; B orgogna, J. C.; Michalek, R. D.; Roberts, D. W.; Rath, J. M.; Glover, E. D.; Ravel, J.; Shardell, M. D.; Yeoman, C. J.;
Brotman, R. M., Cigarette smoking is associated with an altered vaginal tract metabolomic profile. Sci Rep 2018, 8, (1), 852.
109. Song, S. D.; Acharya, K. D.; Zhu, J. E.; Deveney, C. M.; Walther-Antonio, M. R. S.; Tetel, M. J.; Chia, N., Daily Vaginal Micro-
biota Fluctuations Associated with Natural Hormonal Cycle, Contraceptives, Diet, and Exercise. mSphere 2020, 5, (4).
110. Miller, E. A.; Beasley, D. E.; Dunn, R. R.; Archie, E. A., Lactobacilli Dominance and Vaginal pH: Why Is the Human Vaginal
Microbiome Unique? Front Microbiol 2016, 7, 1936.
111. Thoma, M. E.; Klebanoff, M. A.; Rovner, A. J.; Nansel, T. R.; Neggers, Y.; Andrews, W. W.; Schwebke, J. R., Bacterial vagi-
nosis is associated with variation in dietary indices. J Nutr 2011, 141, (9), 1698-704.
112. Saraf, V. S.; Sheikh, S. A.; Ahmad, A.; Gillevet, P. M.; Bokhari, H.; Javed, S., Vaginal microbiome: normalcy vs dysbiosis.
Arch Microbiol 2021, 203, (7), 3793-3802.
113. Nunn, K. L.; Ridenhour, B. J.; Chester, E. M.; Vitzthum, V. J.; Fortenberry, J. D.; Forney, L. J., Vaginal Glycogen, Not Es-
tradiol, Is Associated With Vaginal Bacterial Community Composition in Black Adolescent Women. J Adolesc Health
2019, 65, (1), 130-138.
114. Fettweis, J. M.; Brooks, J. P.; Serrano, M. G.; Sheth, N. U.; Girerd, P. H.; Edwards, D. J.; Strauss, J. F.; The Vaginal Microbi-
ome, C.; Jefferson, K. K.; Buck, G. A., Differences in vaginal microbiome in African American women versus women
of European ancestry. Microbiology (Reading) 2014, 160, (Pt 10), 2272-2282.
115. Sun, S.; Serrano, M. G.; Fettweis, J. M.; Basta, P.; Rosen, E.; Ludwig, K.; Sorgen, A. A.; Blakley, I. C.; Wu, M. C.; Dole, N.;
Thorp, J. M.; Siega-Riz, A. M.; Buck, G. A.; Fodor, A. A.; Engel, S. M., Race, the Vaginal Microbiome, and Spontaneous
Preterm Birth. mSystems 2022, 7, (3), e0001722.
36
37
2.1
Introduction
The diagnosis of vaginitis is often done empirically, based on symptoms and mere gyneco-
logical observation, leading frequently to misdiagnosis and mistreatment. Before initiating
any therapy, confirmation of the diagnosis must be obtained, thus minimizing those risks.1, 2
Nevertheless, diagnostic tests often are not used. In a study performed in 2021 in the US, in
women with symptoms of vaginitis, it was shown that pH measurement, whiff test and wet
mount microscopy (WMM) were performed in only 15%, 21% and 17% of cases, respectively.3
2.2
pH
Despite significant ethnic and geographical variation, as well as across the womans life cy-
cle, the mean pH ranges between 3.8-5.0.4, 5 Lactobacilli, under the influence of estrogens,
are responsible for maintaining the low pH, thus, dysbiotic conditions associated with de-
creased lactobacilli are typically associated with an increased pH.6 Other genera, such as
Fannyhessea (Atopobium)7, Megasphaera, and Leptotrichia are also lactic acid–producing
bacteria and can contribute to reduce the vaginal pH. The relation is bidirectional, as not
only the bacteria modulate the pH, but the opposite is also true.8
Nevertheless, a “normal” pH cannot be assumed as equivalent to absence of “vaginitis”. For
instance, Candida spp. which itself does not affect the vaginal pH, can be found across the
whole spectrum of the vaginal pH and for other conditions, such as bacterial vaginosis (BV),
there is an overlap between normal and abnormal pH.9
Evaluation of the vaginal pH is not useful in the presence of blood, recent exposure to semen
or to vaginal medication. In postmenopausal women not using menopause hormone treat-
ment, the pH is typically increased – making it a good predictor of hypoestrogenism, but of
limited value for the diagnosis of vaginitis.
2
DIAGNOSTIC TOOLS
(alphabetical order)
Ana Rita Silva
Carlos Sousa
Pedro Vieira-Baptista
38
In the presence of an elevated pH, the diagnosis of BV, aerobic vaginitis (AV)/desquamative
inflammatory vaginitis (DIV), trichomoniasis, atrophic vaginitis or of cervicitis (Chlamydia
trachomatis or Neisseria gonorrhoeae) must be considered.10 While candidiasis can occur at
any pH, it is more common and more symptomatic at pH ≤4.5.8 Within the range of normal
to low pH, the diagnosis of cytolytic vaginosis (CV) must also be considered.11
The pH is one of the four Amsel criteria, which are used for the diagnosis of BV (see section
2.7). While the sensitivity of the isolated pH measurement (using a cut-off of 4.5) is relatively
good for the diagnosis of BV (79.0% [95% CI 72.08–84.95%]), the specificity and positive pre-
dictive values are low (56.4% [95% CI 52.22–60.43] and 34.2% [31.53–36.97], respectively).12
This comes as no surprise, given the overlap of situations that course with a high pH. Better
sensitivities can be achieved by reducing the cut-off, but that leads to an unacceptable drop
in the specificity and positive predictive value.12 Of note, the 4.5 cut-off value is still within
the normal range of vaginal pH, especially when considering Black and Hispanic populations.5
For measurement, a pH strip with an adequate range and intervals can be placed directly, for
a few seconds, in contact with the vaginal wall or with the discharge collected in the blade of
the (unmoistened) speculum, avoid-
ing cervical secretions. (Figure 2.1)
Alternatively, the discharge can be
collected with an unmoistened spat-
ula or swab and put in contact with
the pH strip. Another possible strate-
gy may be applying the sample into
the microscopy slide and checking
the pH on the slide, before the appli-
cation of saline.13 The reading should
be made within 5 minutes after the
collection. One study suggests that if
the pH is measured after the adding
of saline to the sample, it may be rea-
sonable to subtract 0.5 from the pH
reading, but we do not recommend
this approach.14
In research, a broad range pH
scale may be desirable (i.e. Mache-
rey-Nagel 3.6–6.1 [Macherey–Nagel GmbH & Co. KG, Duren, Germany]).15, 16 For clinical prac-
tice, a narrower interval may be more practical (i.e. 4.0-5.5).
One study showed that self-evaluation of the pH, by collecting a sample with a gloved finger
and spreading it into a slide to which a pH strip was attached, allowed effective identifica-
tion of abnormal vaginal microbiota and was considered “easy” by more than 90% of the
participating women.17, 18 Other studies showed that self-sampling, using a swab, was also
feasible in both adolescents and adult women and had a moderate (Cohen’s κ 0.53) and al-
most perfect agreement (Cohen’s κ 0.90), respectively, with the clinicians interpretation.19, 20
Figure 2.1 Vaginal pH measurement.
39
Several tests are commercially available, under the recommendation to be used in case of
symptoms of vaginitis.21 Their usefulness is relative and based on a dichotomic and insuffi-
cient concept (candidiasis vs. BV). Some studies show a very good performance for these tests
– surprisingly superior to that of pH evaluation by clinicians in the setting of clinical studies.22
There are commercial tests, such as the FemExam® (Cooper Surgical, CT, USA) and the
QuickVue Advance pH and Amines® (Quidel Corporation, CA, USA) that evaluate the pH
along with the presence of specific enzymes (see section 2.8).23, 24
The role of vaginal pH is limited in settings in which a microscope, cultures and molecular
tests are available. Nevertheless, in settings in which these are not available and when iden-
tifying dysbiosis is relevant (i.e. during pregnancy), it can be useful.8 Also, it has been sug-
gested as possibly being useful for “screening of BV and/or trichomoniasis during pregnancy.25, 26
2.3
Whiff test
The whiff test is another of the four components of the Amsel criteria. A whiff-amine test is
considered positive when a fishy odor occurs following the addition of a drop of 10% potas-
sium hydroxide (KOH) to a sample of vaginal discharge. In cases in which the fishy odor is
readily noticed, there is no need to add the KOH. The alkalinization of the discharge leads to
the release of volatile amines, which are perceived as a rotten or fishy odor. A positive whiff
test is highly suggestive of BV and/or trichomoniasis.
It must be kept in mind that KOH is caustic and thus must be handled carefully. For that
reason, the slide should not be placed too close to the face or directly under the nose. Also,
it can damage the microscope objectives.27
Different studies showed the whiff test to have an excellent specificity (>90%), but a very
low sensitivity (around 40%).12, 28
The Cohen’s kappa agreement rate between users is substantial (0.68). Possible causes for disa-
greement include different KOH concentrations, delay in the performance of the test, insufficient
discharge sample, interference of the collecting device (i.e. cotton swab), and different abilities
to smell.29 There are evidence showing promising results with the use of “electronic noses”.30, 31
2.4
Wet mount microscopy
WMM is currently the most important tool in the diagnosis of women with symptoms of vul-
vovaginitis, despite being performed only by a limited number of practitioners.3 While it is not
a perfect test, it is the gold standard for the diagnosis of AV/DIV and CV and performs very well
in the diagnosis of BV (sensitivity 82-100% and specificity 93-97%). The performance of WMM
for candidiasis seems to be more discrepant across studies (sensitivity 44-78% and specificity
75-89%). In general, it is considered insufficient for the diagnosis of trichomoniasis (sensitiv-
ity 25-82% and specificity 98-100%). WMM can be a very useful resource to evaluate “mixed
infections”, the presence of inflammation and the maturation status of the vaginal mucosa.27
40
Apart from the investment in
a microscope, this an inexpen-
sive technique, that often al-
lows an immediate diagnosis
and appropriate treatment. It
has been shown that a training
of merely 10 hours is enough
to acquire the skills to perform
and accurately interpret WMM.32
The ISSVD recommends that all
providers diagnosing and treat-
ing women with vulvovaginal
symptoms should have training
on this technique.27
Specific details on the WMM diag-
nosis of vaginitis will be provided
in the diagnosis section of each
of the addressed entities and are
summarized in table 2.1; examples
can be seen in figure 2.2.
Further details on the perfor-
mance and interpretation of
WMM can be found in “The In-
ternational Society for the Study
of Vulvovaginal Disease Vaginal
Wet Mount Microscopy Guide-
lines: How to Perform, Applica-
tions, and Interpretation”, pub-
lished in 2021.27
We recommend the use of phase
contrast, despite the initial
higher investment it represents.
Since the slides are usually eval-
uated under a 400x magnifi-
cation, only the 40x objective
lens needs to be phase contrast,
in order to reduce the invest-
ment.27 The use of phase con-
trast allows better identification
of fungal structures, as well as of
cells and background microbiota.
(Figure 2.3)
Figure 2.2 Wet mount microscopy (400x, phase constrast).
A– Normal
B– Bacterial vaginosis
C– Candidiasis
D– Trichomoniasis
E– Cytolytic vaginosis
F– Leptothrix
G– Desquamative inflammatory vaginitis (severe aerobic vaginitis)
H– Vaginal atrophy
41
Nevertheless, it does not solve
the lack of agreement between
observers, concerning the
evaluation of inflammation.33
The use of KOH is not neces-
sary if phase contrast is used.27
The sample should ideally be
collected when the woman
is symptomatic, and without
having used any vaginal prod-
ucts, nor sexual intercourse or
bleeding in the previous 48-
72 hours.34 For women with
suspected BV, the follicular
phase may be the best moment
for evaluation, while the luteal
phase may be better if candidi-
asis or CV are suspected.35
There is no consensus on the
best place from where to col-
lect the sample. However, the
posterior fornix should be
avoided, as it tends to be more
exposed to the cervical secre-
tions and thus, have a higher
pH and more inflammatory
cells. For the same reason,
touching the cervix should be
avoided. (Figure 2.4)
One study showed a higher
sensitivity for Candida spp.
and BV if the sampling is per-
formed from the lower third
of the vagina; for CV, the best
results were achieved when
sampling from the anterior
fornix. However, the study
was performed on mostly asymptomatic women and these differences may be less relevant
in symptomatic women.36
There are no studies validating the use of self-sampling. Except for the diagnosis of trichomo-
niasis, there seems to be no advantage in the use of a speculum for sampling.
Figure 2.3 Wet mount microscopy (400x); granular microbiota
suggestive of bacterial vaginosis.
A– Without phase contrast B– With phase contrast
Figure 2.4 Sampling of vaginal discharge for wet mount microscopy
42
TABLE 2.1 Wet mount microscopy characteristics of different conditions.
AV – aerobic vaginitis, BV – bacterial vaginosis, DIV – desquamative inflammatory vaginitis,
NAAT – nucleic acid amplification tests
Lactobacilli Background
microbiota Inflammation Epithelial cells Other/
comments
Candidiasis
Usually nor-
mal, but may
coexist with
any type of
background
microbiota
Usually lactoba-
cilli dominance
Absent or mild/
moderate
According
to expected
for hormonal
status
Presence of
blastospores
Sometimes
mycelium pres-
ent (suggestive of
C. albicans)
Bacterial
vaginosis
Absent or
scarce
Granular
microbiota
Curved motile rods
may be present
(Mobiluncus spp.)
Absent
A significant
portion covered
by bacteria
(clue cells)
“Partial BV” is
possible
Trichomoniasis
Usually
absent or
scarce
Granular
microbiota (BV)
often present
Usually present
(moderate/
severe)
“Toxic” (swollen)
leukocytes often
present
Parabasal and
basal cells
often present
Presence of
motile proto-
zoans (absence
does not exclude
the hypothesis)
Erythrocytes
sometimes
present
Cytolytic
vaginosis Abundant Only lactobacilli Absent
Variable de-
grees of cellular
lysis (bare nuclei
and cytoplasm
debris)
Evaluate
preferably in the
2nd phase of the
cycle
Lactobacillosis
/ leptothtrix
Presence of
elongated
lactobacilli
(8-15x the
normal size)
May coexist
with normal
lactobacilli
Variable Variable Variable
Evaluation
preferably in the
2nd phase of the
cycle
Can coexist with
any other type
of microbiota
Lactobacillosis
used sometimes
to describe
increased
lactobacilli,
without cytolysis
Aerobic
vaginitis/
desquamative
inflammatory
vaginitis
Absent or
scarce
Dominance
of cocci
(sometimes in
chains) or small
rods
Moderate/severe
inflammation
“Toxic”
leukocytes often
present
Parabasal and
basal cells
often present
Trichomoniasis
should be
excluded in
severe forms
43
Thus, not using a speculum may be a good option in women with vulvodynia or other causes
of vulvar pain (i.e. herpes), for those who refuse speculum examination, or girls. If a speculum
is used, it should be unmoistened.37, 38
The sample should be collected avoiding touching the cervix or scraping the vaginal epi-
thelium. Several devices can be used, including an endocervical brush, a plastic spatula, a
Dacron swab or even a talc-free gloved finger.39 Cotton swabs and wooden spatulas are not
ideal, as they can leave fibers or absorb water, respectively.27
The preparation of the slide can be done either by applying a tiny drop of saline and adding
a small amount of discharge to it or by spreading the sample on the slide and then adding
the drop of saline (the first method may be preferable if the discharge is profuse, as it will
help to dilute it and allow a better visualization of its components, with less overlap). After-
wards, a coverslip is placed and pressed to avoid the formation of air bubbles, and the exces-
sive amount of fluid is removed using absorbent paper. Some authors opt to prepare another
slide, using KOH. As mentioned before, it seems less relevant if using phase contrast. If used, ex-
tra care should be taken to clean the excess amount of fluid, as it can damage the objective lens.
The sample should be read immediately, in order to increase the sensitivity for T. vaginalis
(warming the slide can increase its motility and allow a better identification).40 Nevertheless,
deferred reading is also an option. In that case, the sample is spread onto the slide, allowed
to air dry, and later rehydrated.
Vaginal
atrophy
Absent or
scarce
Absent or
presence of cocci
or small rods
Absent
Mostly parabas-
al and basal
cells (usually
low quantity)
Erythrocytes
often present
Atrophic
vaginitis
Absent or
scarce
Absent or
presence of cocci
or small rods
Moderate/severe
inflammation
“Toxic” leukocytes
often present
Parabasal and
basal cells
Differential
diagnosis with
AV/DIV,
trichomoniasis
and lichen planus
Cervicitis Normal or
decreased Normal/mixed
Moderate/severe
Sometimes
with mucous
filaments
According to
expected
hormonal
status
NAAT test
(C. trachomatis
and N. gonor-
rhoea) needed
if cervicitis is
suspected
Lichen planus Often
decreased
Normal/mixed/
absent Moderate/severe
Parabasal cells,
sometimes
with cellular
wall defects
Vulvar
involvement
common
44
2.5
Gram and other staining techniques
Gram stain has been used for the diagnosis of BV for almost 60 years.41 The Nugent score,
applied on Gram-stained slides, is considered the gold standard for the diagnosis of BV.42
It allows an accurate evaluation of the background bacterial morphotypes, despite some de-
bate on whether or not it is superior to WMM, as the former is associated with some degree
of wash out of the background bacteria during the fixation and staining process.43 Self-sam-
pling can be used for Nugent score, with an excellent agreement with physician collected
samples.44, 45 For the best results, the smear should be spread evenly and be very thin, to
reduce the overlap of cells.
For the calculation of the Nugent score, the amount of the different bacterial morphotypes
is semi-quantitively evaluated directly in the Gram-stained smears of vaginal discharge and
the presence of epithelial cells covered with bacteria (clue cells) is also evaluated (Table 2.2).
That quantification is subjective: the scoring intervals are relatively narrow, only a limited num-
ber BV associated bacteria are evaluated and it relies on the identification of bacterial morpho-
types rather than on that of species. The evaluated bacterial genera are 1) Lactobacillus spp., 2)
Gardnerella spp./Bacteroides spp., and 3) Mobiluncus spp.46 This may be improved in the future
by the use of multiplex peptide nucleic acid fluorescence in situ hybridization, targeting the
different evaluated species of bacteria, rather than relying on morphological identification.47
The bacterial morphotypes are counted using an oil immersion objective (1000x magnifica-
tion). The score achieved for each of the evaluated bacterial morphotypes is added togeth-
er, with a total score of 0-3 considered normal (healthy microbiota) and a score of 7 or worse
is consistent with BV. A score of 4-6 is classified as intermediate48, which constitutes another
disadvantage of the Nugent score because it does not represent a “partial” or “light” form of
BV and there is no specific management for it. (Table 2.2)
TABLE 2.2 Calculation of the Nugent score.
The reading is performed under a high-power magnification.
Score 0-3 normal, 4-6 intermediate, >6 bacterial vaginosis
Score Lactobacilli
(Gram positive rods)
Gardnerella spp./Bacteroides spp.
(tiny, Gram-variable coccobacilli and round-
ed, pleomorphic Gram-negative rods)
Mobiluncus spp.
(curved, Gram-negative rods)
0 >30 0 0
1 5-30 <1 1-5
2 1-4 1-4 >5
3 <1 5-30 --
4 0 >30 --
The Nugent score is typically used in research, but as the result is not readily available, it is
time consuming, more expensive than other alternatives (i.e. WMM, Amsel criteria), and re-
quires experienced technicians and facilities, it is often not the first choice in clinical practice.
45
In the future, clinical criteria for the diagnosis of AV/DIV and CV may be further developed
and allow its accurate diagnosis using Gram stain samples. Also, artificial intelligence may
be applied to Gram stain and increase the role of this technique in the evaluation of women
with vulvovaginal symptoms.49
An interesting alternative to the Nugent score is the Ison and Hay criteria, in which a quantitative
estimation of the bacterial morphotypes is not performed. Instead, an evaluation of the relation-
ship between the amounts of bacteria (lactobacilli, mixed bacteria, and cocci morphotypes) is
performed. As the evaluation is relative, nor the field size of the microscope nor the “concentra-
tion” of the samples influence the impression.46, 50 Additionally, it provides a more comprehensive
evaluation of the vaginal microbiome, as it includes a pattern with absent bacteria (grade 0) and
one dominated by cocci (grade IV), probably corresponding partially to AV/DIV. (Table 2.3)
TABLE 2.3 Ison and Hay criteria (adapted from Ison et al.)46
Grade 0 No lactobacilli or other microbiota present
Grade I Lactobacillus spp. morphotypes only
Grade II Reduced Lactobacillus spp. morphotype with mixed bacterial morphotypes
Grade III Mixed bacterial morphotype with few or absent Lactobacillus spp. morphotypes
Grade IV Gram-positive cocci only
This classification system can be used on slides with different staining methods and also on
non-stained smears, but further validation is needed in this field. This simplified assessment
of Gram stained smears can be used as an alternative to Nugent score; a good agreement
between both has been shown.51
While Gram stain is currently not the first recommendation for the diagnosis of cervicitis, if
an endocervical sample is collected and significant inflammation is present (>30 polymor-
phonuclear leukocytes/high power field), and specially in a suggestive clinical context, it
must be considered. In cases of gonococcal cervicitis, the presence of Gram-negative diplo-
cocci can be noted. Nevertheless, a negative microscopy does not rule out the infection and,
if indicated, a molecular test should be performed.
While the Pap test can provide useful information regarding the presence of microorgan-
isms and/or inflammation, it is not a recommended test for the study of a woman with
suspected vaginitis, due to low sensitivity.52 The specificity for the diagnosis of candidiasis,
trichomoniasis and BV is usually high, but the sensitivities may be as low as 25%, 61.4%
and 55%, respectively. One possible explanation for this performance may be the fact that
the collected sample is cervical rather than vaginal. The performances may be different for
conventional and liquid based cytology (increased in the latter for the presence of Candida
spp. and T. vaginalis and decreased for BV).53, 54 Nevertheless, it can point to less common
diagnostics, such as presence of C. trachomatis, herpes simplex virus, cytomegalovirus, En-
terobius vermicularis or Schistosoma spp.53
46
2.6
Cultures
Cultures have a role in the diagnosis of vaginitis, especially in the case of candidiasis, for
which it is considered the gold standard. Even though, in most cases of acute candidiasis,
there may be no advantage in performing routine cultures, as long as WMM is performed
and confirms the diagnosis. Cultures for Candida are mandatory in cases of recurrent candid-
iasis, suspicion of non-albicans candidiasis, negative microscopy and symptoms suggestive
of candidiasis and in cases of therapeutic failure.54, 55
Antifungal sensitivity tests are not routinely used, in part because they are not easily avail-
able, but should be considered when a resistant strain is suspected. It must be taken into
consideration that the pH at which these are routinely performed (usually 7.0) is higher than
that of the vagina and does not reflect the real sensitivities in clinical practice. It has been
shown that the minimal inhibitory concentration (MIC) is increased at lower pH for micona-
zole, clotrimazole, fluconazole, and nystatin.56, 57
The samples for mycological cultures should preferably be transported using a transport
means (i.e. Amies or Stuart) and should be cultivated in no more than six hours.58
There seems to be no disparities between the different culture means (Sabouraud agar,
Nickerson medium, or Microstix-Candida medium).54
We do not recommend routine cultures for bacteria for the study of vaginitis, group A
streptococci infection suspicion being the exception.59 In general, a positive result does not
distinguish colonization from infection. Nevertheless, the presence of N. gonorrhea should
always prompt treatment, as it is always pathogenic. On the other hand, a positive culture for
Gardnerella spp., sometimes wrongly considered a surrogate for BV, is not useful, due to its very
low specificity: it can be cultivated from more than half of healthy, asymptomatic women.60
The usefulness of cultures in women with suspected AV/DIV is also of limited interest, as the
diagnosis is established by microscopy.61
Cultures are an option for the diagnosis of trichomoniasis. The media used can be the Modified
Diamond medium or the InPouch® TV (BioMed Diagnostics Inc., USA), which have a similar per-
formance.62 The reported sensitivities and specificities range between 44-96% and up to 100%,
respectively. 63-65 Cultures are especially useful in cases of suspected resistance but are not readily
available commercially. Cultures using Modified Diamond medium can take up to seven days; for
the InPouch® TV, the majority of positive results can be identified in the first three days.63
While the molecular approach is the recommended one for the diagnosis of gonorrhea,
a sample of endocervical exudate should be cultured to evaluate the antibiotic susceptibility
profiles, in case of treatment failure. A selective medium for N. gonorrhoeae (Thayer-Martin
or Martin-Lewis) should be used, along with a general medium, such as blood agar and
chocolate agar, as some strains of N. gonorrhoeae can be inhibited in selective media.66
47
2.7
Amsel criteria
In 1983, Amsel R et al. evaluated 397 consecutive women presenting to a gynecology office
and investigated different criteria for the diagnosis of “non-specific vaginitis”.67 Later, these
would be known as the “Amsel criteria” and the “non-specific vaginitis” as BV.
These include:
1. A thin grey or white, homogeneous vaginal discharge coating the vaginal walls;
2. Vaginal pH>4.5;
3. A fishy/rotten smell (before or after the addition of KOH [whiff test]);
4. Presence of clue cells on wet mount microscopy.
To be considered positive, a minimum of three out of the four criteria must be present.
The reported ranges of sensitivity across studies are very broad (37-70%), while the specific-
ity is systematically very high.42
These criteria have the advantage of allowing an immediate diagnosis, being of low com-
plexity and cheap. While it may be useful in clinical practice, in the absence of expertise or
availability of a microscope or other tests (i.e. enzymatic or molecular tests), there are significant
limitations to its use. For instance, the evaluation of the characteristics of the vaginal discharge
are subjective and may be affected by the use of previous vaginal medication or douches. Tradi-
tionally, it has been considered that the “normal” pH is lower than 4.5, a concept that is now under
challenge; a study published in 2021 showed that using this cut-off the sensitivity of the pH for the
diagnosis of BV is 79% but the specificity is only 56%.5, 12 Also, this criterion is of very limited value
in postmenopausal women not using hormone treatments for menopause/atrophy. The reading
of the pH in paper strips is also subjective, as previously mentioned. The same is also true for
the evaluation of odor. Also, given the microbiological heterogeneity of BV, there is a chance
that bacteria producing volatile amines may not be significantly present.68 The evaluation of
clue cells implies the existence of a microscope and expertise in its use – limiting the Amsel
use in most settings to the other three criteria. Under these circumstances, the sensitivity
of the Amsel criteria may drop to as low as 22.8%.12 On the other hand, if there is expertise
in the use of WMM, it is an excellent tool for the diagnosis of BV, by evaluating not only the
presence of clue cells, but also the background microbiota.27
The Amsel criteria are still recommended by some scientific societies, but its role is currently
questioned and considered unacceptable by some for a definite diagnosis.42, 69-71 If used, it must
be kept in mind that there is a risk of overlooking “mixed infections”, specially of BV and trichomo-
niasis or candidiasis (coexisting with BV in 60-80% and 20-30% of cases, respectively). (Figure 2.5)
48
Figure 2.5 72 Wet mount microscopy with phase contrast (400x) showing the presence of a “mixed infection”
(Candida spp. and bacterial vaginosis)
In a possible (future) context of screening (i.e. pregnancy), the use of these criteria may
prove to be more challenging, as its performance may be lower in these circumstances (lack
of significant discharge, no odor). These limitations, in part, translate the complexity of BV,
in which, up to now no specific isolated bacterial marker has been proved to be universally
present – and different bacteria and its interaction may lead to different symptoms/signs.68
2.8
Enzymatic tests
The OSOM® BVBlue® (Sekisui Diagnostics, MA, USA) is a commercially available chromogenic
enzymatic test for the diagnosis of BV, with a reported sensitivity of 88-94% and a specificity
of around 96%.73, 74 It evaluates the levels of the enzyme sialidase in a sample of vaginal
discharge. The results are available in 10 minutes. Self-sampling for sialidase tests has been
shown to be specific (90%) but to have a much lower sensitivity (40%).19
The FemExam® (Cooper Surgical, CT, USA) is a diagnostic kit, composed of two plastic cards,
with the approximate size of a credit card, designed for a fast and inexpensive diagnosis of BV.
The first card is used to evaluate the presence of an elevated pH and of trimethylamine; the sec-
ond card measures the proline iminopeptidase (PIP) activity of Gardnerella spp. (discussed ahead).
The vaginal discharge sample is rubbed in the indicated place in the cards and will show a
positive sign for pH if it is >4.7 and for trimethylamine if it is detected (acting as a surrogate
for a positive whiff test). One study showed, that a positive result for pH and trimethylamine
had a sensitivity of 71.4% (61.7–79.8) and a specificity of 72.8% (63.7–80.7).23 Card 2 of the
FemExam® evaluates the PIP activity of Gardnerella spp.. Its sensitivity and specificity for the
49
diagnosis of BV are 70.0% (55.4–82.1) and 80.9% (69.1–71.6), respectively. If both cards are
used and the test is considered positive when at least two out of the three evaluated varia-
bles are present, the sensitivity and specificity are 91.0% (83.1–96.0) and 61.5% (50.7–71.6),
respectively.23 Samples can be self-collected and results are available in five minutes; the use
of this test is not recommended when blood is present.23, 75
QuickVue Advance pH and Amines® (Quidel Corp, San Diego, CA), has a reported sensitivity
of 53% and a specificity of 97%.24 In this test, the pH cut-off is 4.5 and the amine evaluated
is trimethylamine.
These tests can be an option in settings where microscopy is not available and perform well
for the diagnosis of BV. It must be kept in mind that they do not provide information concern-
ing other conditions, such as candidiasis or trichomoniasis, which frequently coexist with BV.
2.9
Molecular tests
Given the limitation of microscopy and point-of-care tests, and in line with what is occurring in
other fields of medicine, there is a trend towards the use of molecular tests for diagnostic purpos-
es. These allow the diagnosis in a reasonable time (hours), are precise and reproducible, capable
of high throughput, and allow the identification of fastidious microorganisms. Also, the demand
for self-sampling is growing in women’s health and molecular tests will likely fulfill this require-
ment; there are already data suggesting that this approach is not inferior to the use of clinician
collected samples.76 Currently, in most settings, price is still a limitation to its widespread use.
The “first generation” of molecular tests for the diagnosis of vaginitis were direct probe as-
says. These are nucleic acid probes that bind to sequences (usually ribosomal RNA) specific
to the targeted microorganisms in a vaginal discharge sample.
The BD Affirm™ VPIII (BD Diagnostic Systems, Sparks, Maryland) is a test based on DNA
probes, that detects the presence of Candida spp. (including C. albicans, C. glabrata, C. kefyr,
C. krusei, C. parapsilosis, and C. tropicalis), T. vaginalis and Gardnerella spp., in vaginal sam-
ples, providing results in 45 minutes.77 The sampling is usually from the vaginal walls, but it
has been suggested that if done from the speculum blades leftover the performance is similar.78
The reported sensitivity and specificity for Candida spp., are 69.4 to 82.76% and 98.80 to 100%,
respectively.77, 79-81 The sensitivity for BV has been reported to range between 75.9 and 96.2%
and the specificity from 60.6 to >95%.27, 77, 81-83 While the specificity for trichomoniasis is also
high, the sensitivity is lower (46.3-100%).77, 79, 84 One study comparing the Affirm™ VPIII with
the BD Max™ vaginal panel (Table 2.4) showed the latter to be more accurate: higher spec-
ificity for BV and more sensitive to Candida spp., without a decrease in specificity.77 It can
be used in symptomatic women, along with pH and whiff test, for a better performance.42
In the Affirm™ VPIII, the diagnosis of BV is based on the detection of Gardnerella spp. (being positive
if the load is higher than 5 x 105 colony forming units [CFUs] per milliliter). While targeting a single
agent for the diagnosis of BV seems a strategy of limited interest, the results are reasonable.41, 85
While one study showed that the performance of the Affirm™ VPIII for the diagnosis of BV is
50
unchanged in pregnant women86, another one, using self-collected vaginal swabs concluded
that it had insufficient sensitivity for BV, candidiasis and trichomoniasis during pregnancy.87
Besides the provided transport means, the use of the ESwab™ (Copan Diagnostics, Murrieta,
CA) has also been validated as an option.88
Among a survey of members of the American College of Obstetricians and Gynecologists,
performed in 2018, 25% noted using this test in their clinical practice.89
More recently, nucleic acid amplification tests (NAATs), including polymerase chain reac-
tion (PCR) entered the market and are very promising approaches for the diagnosis and
management of vaginitis. (Table 2.4) These have been acknowledged as the gold standard
for the diagnosis of trichomoniasis for some years.42 Adding the possibility of using a molecular
approach to test at once not only trichomoniasis, but also candidiasis and BV is of great interest.
For T. vaginalis and Candida spp. the challenge was not that high (it was requested to iden-
tify one species or genus), in striking contrast with BV (a complex multibacterial condition,
of which there is no specific marker).6 While correctly identifying bacteria or other microor-
ganisms is an uncomplicated task for NAATs, defining BV through this technology is much
more complex: it requires identifying and quantifying diverse BV associated bacteria, as well
as lactobacilli, to evaluate their relative proportions.90
There are several validated and commercially available tests for the isolated diagnosis of
trichomoniasis (i.e. Xpert® TV, Aptima TV91, 92) and BV (i.e. Allplex™ Bacterial Vaginosis assay,
MDL OneSwab® BV Panel AmpliSens®).93-95
More relevant for the clinical practice are the tests that allow the concomitant diagnosis of
BV, candidiasis and trichomoniasis. We found clinical validation data for:
Allplex™ Vaginitis (Seegene, Seoul, Korea);
BD Max™ Vaginal Panel (Becton Dickinson, USA);
Hologic Aptima® BV (Hologic, USA);
NuSwab® VG (LabCorp, USA).
Another available test, to which we could not identify published validation data is the Quest
Diagnostics SureSwab® BV (Quest Diagnostics, USA).96
In general, the available tests have shown great performance, for all three conditions. One
study showed a lower sensitivity for the BD Max™ vaginal panel, but it must be taken into
account that the comparison was made against amplicon sequencing of the 16S ribosomal
RNA, which is currently not standardized.97
Van der Pol et al. showed a very high agreement in the detection of T. vaginalis between the
BD Max™ Vaginal Panel and the BD Max™ CT/GC/TV assay.98
These tests have the advantage of being unaffected by the presence of multiple infections or by
prevalence variation of the conditions.16, 76, 79 Real world data shows that the use of NAATs can im-
prove value-based care and even lead to a 12% risk reduction of preterm delivery in symptomat-
ic pregnant women.99 A molecular approach leads to more accurate diagnosis, decreasing both
nondiagnosis and overdiagnosis and, consequently, allowing timely and correct treatment.100
51
TABLE 2.4 Nucleic acid amplification tests for the diagnosis of vaginitis.
*Calculated from the data in the paper.; § clinician collected sample; + patient collected sample.
BV – bacterial vaginosis, BVAB – bacterial vaginosis associated bacteria, NA not applicable; NAC – non-albicans candidiasis
Test Study Bacterial vaginosis Candida Trichomoniasis Notes
Sensitivity Specificity Sensitivity Specificity Sensitivity Specificity
Allplex™
Vaginitis
(Seegene®,
Seoul,
Korea)
Vieira-Baptista
et al16
91.7%
(86.49-95.40%)
86.6%
(83.57-89.24%)
All species 91.1%
(82.23–96.08%)
C. albicans 88.1%
(77.82–94.70%)
NAC 100%
(85.18–100.00%)
All species
95.6% (93.65–
97.12%)
C. albicans 98.2%
(96.83–99.11%)
NAC 97.5%
(95.73–98.38%)
94.4%
(72.71–99.86%)
99.9%
(99.25–100%)
- One swab from the upper third of
the vagina
- NAC includes: C. glabrata, C.
tropicalis, C. parapsilosis, C. krusei, C.
lusitaniae and C. dubliniensis
- BV diagnosis based on the evalua-
tion of: Fannyhessea (Atopobium) va-
ginae, Gardnerella spp., Lactobacillus
spp., Mobiluncus spp
Aptima®
CV/TV
+
Aptima®
BV
(Hologic,
San Diego,
USA)
Richter S
et al27
90.0%
(76.4–96.6%)
92.3%
(81.3–97.5%) NA NA NA NA
- Candida group includes: C.
albicans, C. parapsilosis, C. tropicalis, C.
dubliniensis
- Separably identifies: C. glabrata
- BV diagnosis based on evaluatiom
of Lactobacillus spp. Gardnerella. spp
and F. vaginae
Schwebke JR
et al76
95.0%
(93.1–96.4%)§
97.3%
(95.8–98.2%)+
89.6%
(87.1–91.6%)§
85.8%
(83.1–88.2%)+
Candida
group: 91.7%
(88.7–94.0%)§
Candida group:
92.9 (905.0–
95.0%)+
C. glabrata: 84.7%
(73.5–91.8%)§
C. glabrata: 86.2%
(75.1–92.8%)+
Candida
group: 99.1%
(98.4–99.5%)§
Candida
group: 98.7%
(98.0–99.2%)+
C. glabra-
ta: 95.1%
(93.8–96.2%)§
C. glabra-
ta: 98.9%
(98.2–99.4%)+
96.5%
(92.0–98.5%)§
97.1%
(92.9–98.9%)+
95.1% (93.8–
96.2%)§
98.9% (98.2–
99.4%)+
52
using NAAT. These systems differ in their amplification methods and in the sequences used as tar-
get and, in general, they all offer a high sensitivity (>90 %) and a very high specificity (>99 %). Due
to its high sensitivity, it is important to adjust the working conditions to prevent contamination.
The samples can be cervical/vaginal swabs, either physician or self-collectedor or first void urine.42
While promising in research contexts, next generation sequencing (NGS) is not likely, at the
moment, to be cost-effective or practical for the diagnosis of vaginitis.97 When considering
BV and AV/DIV, they fall within Ravels community state type (CST) IV – the diversity group,
in which there is no dominance of lactobacilli.5, 101
NAATs will probably play a central role in the management of vaginitis in the near future. Sev-
eral characteristics contribute to it: reasonable turn-around times, accurate diagnosis, observer
independent results, possibility of large scale screening of populations (i.e. pregnant women),
accurate microbiological characterization of the situation (i.e. composition of BV, identification
of specific clades of bacteria with higher associated risk), identification of antibiotic resistance
genes, and possibility of coupling it with other relevant tests (i.e. cervicitis, HPV).41, 102 Neverthe-
less, the costs of molecular tests are still a significant barrier to access. While technologically
challenging, this approach may, paradoxically, be the response for less developed countries, in
which there is lack of trained personal – and where the highest toll due to vaginitis is paid.102
BD Max™
Vaginal
Panel
(Becton
Dickinson,
USA)
Sherrard J103 94.4% (86.2-
98.4%)
79.0% (70.8-
85.8%)
All species 86.4%
(75.0-94.0%)
All species: 86%
(79.1- 91.4%)
77.8% (40.0-
97.2%)
98.9% (96.2-
99.9%)
- One vaginal swab
- Candida group includes: C. albicans,
C. tropicalis, C. parapsilosis, and C.
dubliniensis)
- Separably identifies: C. glabrata,
and C. krusei
- BV diagnosis based on evaluatiom
of L. crispatus and L. jensenii and BV
associated bacteria (Gardnerella
spp., F. vaginae, Megasphaera-1, and
BVAB-2);
Aguirre
-Quiñonero
et al.104
89.8%
(85.0–93.1%)
96.5%
(95.1–97.6%)
All species: 97.2%
(94.8–98.5%)
C. glabrata: 100%
(85.7–100%)
C. krusei: 83.3%
(43.6–97.0%)
All species: 96.8%
(95.3–97.9%)
C. glabrata: 100%
(99.6–100%)
C. krusei: 100%
(99.6–100%)
100%
(79.6–100%)
100%
(99.6–100%)
Van den
Munckhof
et al 97
63.9% (47.58-
77.52%)
98.7% (93.17-
99.78%)
Not
evaluated
Not
evaluated Not evaluated Not
evaluated
Thompson
A et al77
96.2%
(89.3–99.2%)
96.1% (
89.8–98.7%)
All species: 98.4%
(91.3–99.6%)
All species:
95.4%
(90.3–98.3%)
87.5% (47.4-
99.7%)*
100% (98-
100%)*
NuSwab®
(LabCorp,
Burling-
ton,NC)
Danby C
et al105
78.7%
(67.7–87.3%)
97.6%
(91.6–99.7%)
92.4% (83.2–
97.5%)
100%
(96.1–100%)
100%
(79.4–100%)
100%
(97.4–100%) - Identifies C. albicans and C.
glabrata
- BV diagnosis based on evalua-
tion of A. vaginae, BVAB 2, and
Megasphaera spp.
Cartwright C
et al79
96.9%
(94.5-99.3%)
92.6%
(87.7-97.5%)
97.7%
(93.3-100%)
93.2%
(86.8-99.6%)
98.1%
(94.5-100%)
Not
calculated
53
Recommendations
Recommendation Quality of
evidence
Strength of
recommendation
Empirical diagnosis of vaginitis is not recommended. 1b A
A normal pH is insufficient to exclude the presence of vaginitis. 2b B
Evaluation of the vaginal pH is not useful in the presence of blood, recent
exposure to semen or to vaginal medication, or in postmenopausal
women not using hormone therapy.
4 C
In clinical practice, a narrow pH interval scale may be used (i.e. 4.0-5.5). 5 D
Self-sampling for vaginal pH measurement can be used. 3b C
In cases in which the fishy odor is readily noticed, there is no need to add KOH. 5 D
A positive whiff test is highly suggestive (specific) of bacterial vaginosis
and/or trichomoniasis, but the sensitivity is low. 2a B
The use of wet mount microscopy is recommended as the initial step in
the diagnosis of vaginitis. 2a B
All providers diagnosing and treating women with vulvovaginal symp-
toms should have training on wet mount microscopy. 5 D
The use of phase contrast is considered preferrable. 3b C
Wet mount microscopy is the gold standard for the diagnosis of cytolytic
vaginosis and aerobic vaginitis/desquamative inflammatory vaginitis. 5 D
The sample for microscopy should be collected when the woman is
symptomatic, and without having used any vaginal products, nor sexual
intercourse or bleeding in the previous 48-72 hours.
5 D
The posterior fornix and touching the cervix should be avoided when
sampling the vagina for microscopy. 4 C
The use of a speculum is not mandatory for sampling for microscopy. 4 C
The molecular approach is not free of pitfalls, including the need for clinical validation of the results.
For instance, similar to what happens with cultures, it does allow distinguishing colonization vs. in-
fection in cases in which Candida spp. is identified. While the identification of T. vaginalis will of ne-
cessity lead to treatment, the same may not be true in cases of BV. The s ymptom s of differe nt vul -
vovaginal conditions tend to be similar and overlap and that may be problematic, especially
with the trend to move to self-collection of samples: significant pathology may be missed,
and the symptoms wrongly attributed to whatever returns positive in the molecular test.
Also, the landscape of vaginitis is wider than BV, trichomoniasis and candidiasis – these plat-
forms will have to adapt in the future, as knowledge increases in the field. Currently, there
are no validated molecular approaches for the diagnosis of CV. The molecular diagnosis of
AV/DIV seems feasible but is not yet validated nor commercially available.95
NAATs should be used for the diagnosis of the agents involved in cervicitis. There are different plat-
forms on the market for the joint detection of C. trachomatis and N. gonorrhoeae in the same sample
54
The use of cotton swabs and wooden spatulas should be avoided for
sampling for microscopy. 5 D
The sample for microscopy should be read immediately, but deferred
reading is possible. 5D
The Nugent score, applied on Gram-stained slides, is considered the gold
standard for the diagnosis of bacterial vaginosis. 1a A
The Ison-Hay criteria are an alternative for the diagnosis of bacterial
vaginosis. 3b C
Self-sampling can be used for Nugent score. 4 C
Wet mount microscopy is a reliable tool for the diagnosis of bacterial
vaginosis in clinical practice. 1b A
The Pap test should not be used for the diagnosis of vaginitis. 4 C
Cultures are the gold standard for the diagnosis of candidiasis. 1a A
Cultures may be omitted in acute candidiasis, but should be performed
in chronic/recurrent cases, treatment failure, if non-albicans species is
suspected or if symptoms present and microscopy negative.
5 D
Sensitivity tests are only recommended is case of suspected resistance
(Candida). 5 D
The samples for mycological cultures should preferably be transported
using a transport means (i.e. Amies or Stuart) and should be cultivated in
no more than six hours.
5 D
There seem to be no disparities between the different culture means
(Sabouraud agar, Nickerson medium, or Microstix-Candida medium) for
Candida.
2b B
Routine cultures for bacteria for the study of vaginitis are not recom-
mended. 5 D
Cultures are not recommended for the diagnosis of aerobic vaginitis/
desquamative inflammatory vaginitis. 4 C
Cultures of trichomonas can be considered when resistance is suspected. 5 D
The Amsel criteria may be used in clinical practice, in the absence of
expertise or availability of a microscope or other tests. 1b A
Enzymatic tests for the diagnosis of bacterial vaginosis can be used in
settings in which there is no expertise in microscopy. 4 C
Validated nucleic acid amplification tests can be used for the diagnosis of
candidiasis, trichomoniasis and bacterial vaginosis. 1b A
Validated nucleic acid amplification tests are recommended for the
diagnosis of cervicitis. 1a A
Next generation sequencing is currently not recommended for the
diagnosis of vaginitis. 5 D
55
References
1. Landers, D. V.; Wiesenfeld, H. C.; Heine, R. P.; Krohn, M. A.; Hillier, S. L., Predictive value of the clinical diagnosis of lower
genital tract infection in women. Am J Obstet Gynecol 2004, 190, (4), 1004-10.
2. Cerca, N., Addressing the challenges with bacterial vaginosis pharmacotherapy. Expert Opin Pharmacother 2022, 1-3.
3. Hillier, S. L.; Austin, M.; Macio, I.; Meyn, L. A.; Badway, D.; Beigi, R., Diagnosis and Treatment of Vaginal Discharge
Syndromes in Community Practice Settings. Clin Infect Dis 2021, 72, (9), 1538-1543.
4. Carr, P. L.; Felsenstein, D.; Friedman, R. H., Evaluation and management of vaginitis. J Gen Intern Med 1998, 13, (5), 335-46.
5. Ravel, J.; Gajer, P.; Abdo, Z.; Schneider, G. M.; Koenig, S. S.; McCulle, S. L.; Karlebach, S.; Gorle, R.; Russell, J.; Tacket, C.
O.; Brotman, R. M.; Davis, C. C.; Ault, K.; Peralta, L.; Forney, L. J., Vaginal microbiome of reproductive-age women. Proc
Natl Acad Sci U S A 2011, 108 Suppl 1, (Suppl 1), 4680-7.
6. Lev-Sagie, A.; De Seta, F.; Verstraelen, H.; Ventolini, G.; Lonnee-Hoffmann, R.; Vieira-Baptista, P., The Vaginal Microbi-
ome: II. Vaginal Dysbiotic Conditions. J Low Genit Tract Dis 2022, 26, (1), 79-84.
7. Konschuh, S.; Jayaprakash, T.; Dolatabadi, A.; Dayo, E.; Ramay, H.; Sycuro, L., O02.3 Reclassification of Atopobium
vaginae as three novel Fannyhessea species: implications for understanding their role in bacterial vaginosis. Sexually
Transmitted Infections 2021, 97, (Suppl 1), A18-A18.
8. Linhares, I. M.; Summers, P. R.; Larsen, B.; Giraldo, P. C.; Witkin, S. S., Contemporary perspectives on vaginal pH and
lactobacilli. Am J Obstet Gynecol 2011, 204, (2), 120.e1-5.
9. Benyas, D.; Sobel, J. D., Mixed Vaginitis Due to Bacterial Vaginosis and Candidiasis. J Low Genit Tract Dis 2022, 26, (1), 68-70.
10. Pavletic, A. J.; Hawes, S. E.; Geske, J. A.; Bringe, K.; Polack, S. H., Experience with routine vaginal pH testing in a family
practice setting. Infect Dis Obstet Gynecol 2004, 12, (2), 63-8.
11. Cibley, L. J.; Cibley, L. J., Cytolytic vaginosis. Am J Obstet Gynecol 1991, 165, (4 Pt 2), 1245-9.
12. Vieira-Baptista, P.; Silva, A. R.; Costa, M.; Figueiredo, R.; Saldanha, C.; Sousa, C., Diagnosis of bacterial vaginosis: Clinical
or microscopic? A cross-sectional study. Int J Gynaecol Obstet 2022, 156, (3), 552-559.
13. Donders, G.; Slabbaert, K.; Vancalsteren, K.; Pelckmans, S.; Bellen, G., Can vaginal pH be measured from the wet
mount slide? Eur J Obstet Gynecol Reprod Biol 2009, 146, (1), 100-3.
14. Bakir, S.; Elas, D.; Stockdale, C. K.; Zimmerman, M. B.; Hardy-Fairbanks, A., Accuracy of Vaginal pH Testing Before and
After Addition of Sterile Saline. J Low Genit Tract Dis 2021, 25, (2), 181-185.
15. Donders, G. G.; Caeyers, T.; Tydhof, P.; Riphagen, I.; van den Bosch, T.; Bellen, G., Comparison of two types of dipsticks
to measure vaginal pH in clinical practice. Eur J Obstet Gynecol Reprod Biol 2007, 134, (2), 220-4.
16. Vieira-Baptista, P.; Silva, A. R.; Costa, M.; Aguiar, T.; Saldanha, C.; Sousa, C., Clinical validation of a new molecular test
(Seegene Allplex™ Vaginitis) for the diagnosis of vaginitis: a cross-sectional study. Bjog 2021, 128, (8), 1344-1352.
17. Donders, G. G.; Gonzaga, A.; Marconi, C.; Donders, F.; Michiels, T.; Eggermont, N.; Bellen, G.; Lule, J.; Byamughisa, J.,
Increased vaginal pH in Ugandan women: what does it indicate? Eur J Clin Microbiol Infect Dis 2016, 35, (8), 1297-303.
18. Donders, G. G.; Andabati, G.; Donders, F.; Michiels, T.; Eggermont, N.; Bellen, G.; Lulé, J., Acceptance of self-testing for
increased vaginal pH in different subsets of Ugandan women. Int J STD AIDS 2012, 23, (1), 30-5.
19. Huppert, J. S.; Hesse, E. A.; Bernard, M. C.; Bates, J. R.; Gaydos, C. A.; Kahn, J. A., Accuracy and trust of self-testing for
bacterial vaginosis. J Adolesc Health 2012, 51, (4), 400-5.
20. Roy, S.; Caillouette, J. C.; Faden, J. S.; Roy, T.; Ramos, D. E., Improving appropriate use of antifungal medications: the
role of an over-the-counter vaginal pH self-test device. Infect Dis Obstet Gynecol 2003, 11, (4), 209-16.
21. Lin, Y. P.; Chen, W. C.; Cheng, C. M.; Shen, C. J., Vaginal pH Value for Clinical Diagnosis and Treatment of Common
Vaginitis. Diagnostics (Basel) 2021, 11, (11).
22. Shen, C. J.; Yang, C. Y.; Chen, H. Y.; Chen, W. C.; Chang, T. C.; Cheng, C. M., Clinical Evaluation of a Self-Testing Kit for
Vaginal Infection Diagnosis. J Healthc Eng 2021, 2021, 4948954.
23. West, B.; Morison, L.; Schim van der Loeff, M.; Gooding, E.; Awasana, A. A.; Demba, E.; Mayaud, P., Evaluation of a new
rapid diagnostic kit (FemExam) for bacterial vaginosis in patients with vaginal discharge syndrome in The Gambia.
Sex Transm Dis 2003, 30, (6), 483-9.
24. Charonis, G.; Larsson, P. G., Use of pH/whiff test or QuickVue Advanced pH and Amines test for the diagnosis of bacterial
vaginosis and prevention of postabortion pelvic inflammatory disease. Acta Obstet Gynecol Scand 2006, 85, (7), 837-43.
25. Gjerdingen, D.; Fontaine, P.; Bixby, M.; Santilli, J.; Welsh, J., The impact of regular vaginal pH screening on the diagno-
sis of bacterial vaginosis in pregnancy. J Fam Pract 2000, 49, (1), 39-43.
26. Hosny, A.; El-Khayat, W.; Kashef, M. T.; Fakhry, M. N., Association between preterm labor and genitourinary tract
infections caused by Trichomonas vaginalis, Mycoplasma hominis, Gram-negative bacilli, and coryneforms. J Chin
Med Assoc 2017, 80, (9), 575-581.
56
27. Vieira-Baptista, P.; Grincevičienė, Š.; Oliveira, C.; Fonseca-Moutinho, J.; Cherey, F.; Stockdale, C. K., The International So-
ciety for the Study of Vulvovaginal Disease Vaginal Wet Mount Microscopy Guidelines: How to Perform, Applications,
and Interpretation. J Low Genit Tract Dis 2021, 25, (2), 172-180.
28. Modak, T.; Arora, P.; Agnes, C.; Ray, R.; Goswami, S.; Ghosh, P.; Das, N. K., Diagnosis of bacterial vaginosis in cases of ab-
normal vaginal discharge: comparison of clinical and microbiological criteria. J Infect Dev Ctries 2011, 5, (5), 353-60.
29. Cohrssen, A.; Anderson, M.; Merrill, A.; McKee, D., Reliability of the whiff test in clinical practice. J Am Board Fam Pract
2005, 18, (6), 561-2.
30. Hay, P.; Tummon, A.; Ogunfile, M.; Adebiyi, A.; Adefowora, A., Evaluation of a novel diagnostic test for bacterial vagi-
nosis: ‘the electronic nose. Int J STD AIDS 2003, 14, (2), 114-8.
31. Chandiok, S.; Crawley, B. A.; Oppenheim, B. A.; Chadwick, P. R.; Higgins, S.; Persaud, K. C., Screening for bacterial
vaginosis: a novel application of artificial nose technology. J Clin Pathol 1997, 50, (9), 790-1.
32. Donders, G. G.; Marconi, C.; Bellen, G.; Donders, F.; Michiels, T., Effect of short training on vaginal fluid microscopy
(wet mount) learning. J Low Genit Tract Dis 2015, 19, (2), 165-9.
33. Donders, G. G.; Larsson, P. G.; Platz-Christensen, J. J.; Hallén, A.; van der Meijden, W.; Wölner-Hanssen, P., Variability in
diagnosis of clue cells, lactobacillary grading and white blood cells in vaginal wet smears with conventional bright
light and phase contrast microscopy. Eur J Obstet Gynecol Reprod Biol 2009, 145, (1), 109-12.
34. Santiago, G. L.; Cools, P.; Verstraelen, H.; Trog, M.; Missine, G.; El Aila, N.; Verhelst, R.; Tency, I.; Claeys, G.; Temmerman,
M.; Vaneechoutte, M., Longitudinal study of the dynamics of vaginal microflora during two consecutive menstrual
cycles. PLoS One 2011, 6, (11), e28180.
35. Morison, L.; Ekpo, G.; West, B.; Demba, E.; Mayaud, P.; Coleman, R.; Bailey, R.; Walraven, G., Bacterial vaginosis in rela-
tion to menstrual cycle, menstrual protection method, and sexual intercourse in rural Gambian women. Sex Transm
Infect 2005, 81, (3), 242-7.
36. Azevedo, S.; Lima-Silva, J.; Vieira-Baptista, P., Impact of the Sampling Site in the Result of Wet Mount Microscopy. J
Low Genit Tract Dis 2019, 23, (2), 176-181.
37. Frobenius, W.; Bogdan, C., Diagnostic Value of Vaginal Discharge, Wet Mount and Vaginal pH - An Update on the
Basics of Gynecologic Infectiology. Geburtshilfe Frauenheilkd 2015, 75, (4), 355-366.
38. Audisio, T.; Penacino, M.; Cannistraci, R.; Bertolotto, P., Detection of bacterial vaginosis, Trichomonas vaginalis infec-
tion, and vaginal Candida infection: a comparative study of methods of extracting exudates, with and without a
speculum, during pregnancy. J Low Genit Tract Dis 2005, 9, (4), 213-5.
39. Hemalatha, R.; Ramalaxmi, B. A.; Swetha, E.; Balakrishna, N.; Mastromarino, P., Evaluation of vaginal pH for detection
of bacterial vaginosis. Indian J Med Res 2013, 138, (3), 354-9.
40. Kissinger, P. J.; Dumestre, J.; Clark, R. A.; Wenthold, L.; Mohammed, H.; Hagensee, M. E.; Martin, D. H., Vaginal swabs
versus lavage for detection of Trichomonas vaginalis and bacterial vaginosis among HIV-positive women. Sex
Transm Dis 2005, 32, (4), 227-30.
41. Coleman, J. S.; Gaydos, C. A., Molecular Diagnosis of Bacterial Vaginosis: an Update. J Clin Microbiol 2018, 56, (9).
42. Workowski, K. A.; Bachmann, L. H.; Chan, P. A.; Johnston, C. M.; Muzny, C. A.; Park, I.; Reno, H.; Zenilman, J. M.; Bolan, G.
A., Sexually Transmitted Infections Treatment Guidelines, 2021. MMWR Recomm Rep 2021, 70, (4), 1-187.
43. Donders, G. G.; Vereecken, A.; Dekeersmaecker, A.; Van Bulck, B.; Spitz, B., Wet mount microscopy reflects functional
vaginal lactobacillary flora better than Gram stain. J Clin Pathol 2000, 53, (4), 308-13.
44. Camus, C.; Penaranda, G.; Khiri, H.; Camiade, S.; Molet, L.; Lebsir, M.; Plauzolles, A.; Chiche, L.; Blanc, B.; Quarello, E.; Hal-
fon, P., Acceptability and efficacy of vaginal self-sampling for genital infection and bacterial vaginosis: A cross-sec-
tional study. PLoS One 2021, 16, (11), e0260021.
45. Kashyap, B.; Singh, R.; Bhalla, P.; Arora, R.; Aggarwal, A., Reliability of self-collected versus provider-collected vaginal
swabs for the diagnosis of bacterial vaginosis. Int J STD AIDS 2008, 19, (8), 510-3.
46. Ison, C. A.; Hay, P. E., Validation of a simplified grading of Gram stained vaginal smears for use in genitourinary med-
icine clinics. Sex Transm Infect 2002, 78, (6), 413-5.
47. Machado, A.; Cerca, N., Multiplex Peptide Nucleic Acid Fluorescence In Situ Hybridization (PNA-FISH) for Diagnosis
of Bacterial Vaginosis. Methods Mol Biol 2017, 1616, 209-219.
48. Joesoef, M. R.; Hillier, S. L.; Josodiwondo, S.; Linnan, M., Reproducibility of a scoring system for gram stain diagnosis
of bacterial vaginosis. J Clin Microbiol 1991, 29, (8), 1730-1.
49. Dong, M.; Wang, C.; Li, H.; Yan, Y.; Ma, X.; Li, H.; Li, X.; Wang, H.; Zhang, Y.; Qi, W.; Meng, K.; Tian, W.; Wang, Y.; Fan, A.;
Han, C.; Donders, G. G. G.; Xue, F., Aerobic Vaginitis Diagnosis Criteria Combining Gram Stain with Clinical Features:
An Establishment and Prospective Validation Study. Diagnostics (Basel) 2022, 12, (1).
57
50. Larsson, P. G.; Carlsson, B.; Fåhraeus, L.; Jakobsson, T.; Forsum, U., Diagnosis of bacterial vaginosis: need for validation
of microscopic image area used for scoring bacterial morphotypes. Sex Transm Infect 2004, 80, (1), 63-7.
51. Forsum, U.; Jakobsson, T.; Larsson, P. G.; Schmidt, H.; Beverly, A.; Bjørnerem, A.; Carlsson, B.; Csango, P.; Donders, G.; Hay,
P.; Ison, C.; Keane, F.; McDonald, H.; Moi, H.; Platz-Christensen, J. J.; Schwebke, J., An international study of the interob-
server variation between interpretations of vaginal smear criteria of bacterial vaginosis. Apmis 2002, 110, (11), 811-8.
52. Levi, A. W.; Harigopal, M.; Hui, P.; Schofield, K.; Chhieng, D. C., Comparison of Affirm VPIII and Papanicolaou tests in
the detection of infectious vaginitis. Am J Clin Pathol 2011, 135, (3), 442-7.
53. Fitzhugh, V. A.; Heller, D. S., Significance of a diagnosis of microorganisms on pap smear. J Low Genit Tract Dis 2008,
12, (1), 40-51.
54. Sobel, J. D., Vulvovaginal candidosis. Lancet 2007, 369, (9577), 1961-71.
55. Farr, A.; Effendy, I.; Tirri, B. F.; Hof, H.; Mayser, P.; Petricevic, L.; Ruhnke, M.; Schaller, M.; Schäfer, A. P. A.; Willinger, B.;
Mendling, W., Vulvovaginal Candidosis (Excluding Mucocutaneous Candidosis): Guideline of the German (DGGG),
Austrian (OEGGG) and Swiss (SGGG) Society of Gynecology and Obstetrics (S2k-Level, AWMF Registry Number
015/072, September 2020). Geburtshilfe Frauenheilkd 2021, 81, (4), 398-421.
56. Liu, W.; Zhang, X.; Liu, Z.; Luo, X., Impact of pH on the antifungal susceptibility of vaginal Candida albicans. Int J
Gynaecol Obstet 2011, 114, (3), 278-80.
57. Danby, C. S.; Boikov, D.; Rautemaa-Richardson, R.; Sobel, J. D., Effect of pH on in vitro susceptibility of Candida gla-
brata and Candida albicans to 11 antifungal agents and implications for clinical use. Antimicrob Agents Chemother
2012, 56, (3), 1403-6.
58. Donders, G. G.; Bellen, G.; Mendling, W., Management of recurrent vulvo-vaginal candidosis as a chronic illness.
Gynecol Obstet Invest 2010, 70, (4), 306-21.
59. Donders, G.; Greenhouse, P.; Donders, F.; Engel, U.; Paavonen, J.; Mendling, W., Genital Tract GAS Infection ISIDOG
Guidelines. J Clin Med 2021, 10, (9).
60. Stockdale, C. K., A Positive Culture Result for Gardnerella Is Not Diagnostic of Bacterial Vaginosis. J Low Genit Tract
Dis 2016, 20, (4), 281-2.
61. Donders, G. G. G.; Bellen, G.; Grinceviciene, S.; Ruban, K.; Vieira-Baptista, P., Aerobic vaginitis: no longer a stranger. Res
Microbiol 2017, 168, (9-10), 845-858.
62. Levi, M. H.; Torres, J.; Piña, C.; Klein, R. S., Comparison of the InPouch TV culture system and Diamond’s modified
medium for detection of Trichomonas vaginalis. J Clin Microbiol 1997, 35, (12), 3308-10.
63. Vieira-Baptista, P.; Bornstein, J., Candidiasis, Bacterial Vaginosis, Trichomoniasis and Other Vaginal Conditions Affecting the
Vulva. In Vulvar Disease: Breaking the Myths, Bornstein, J., Ed. Springer International Publishing: Cham, 2019; pp 167-205.
64. Ohlemeyer, C. L.; Hornberger, L. L.; Lynch, D. A.; Swierkosz, E. M., Diagnosis of Trichomonas vaginalis in adolescent
females: InPouch TV culture versus wet-mount microscopy. The Journal of adolescent health: official publication of the
Society for Adolescent Medicine 1998, 22, (3), 205-8.
65. Hobbs, M. M.; Sena, A. C., Modern diagnosis of Trichomonas vaginalis infection. Sex Transm Infect 2013, 89, (6), 434-8.
66. Ng, L. K.; Martin, I. E., The laboratory diagnosis of Neisseria gonorrhoeae. Can J Infect Dis Med Microbiol 2005, 16, (1), 15-25.
67. Amsel, R.; Totten, P. A.; Spiegel, C. A.; Chen, K. C.; Eschenbach, D.; Holmes, K. K., Nonspecific vaginitis. Diagnostic
criteria and microbial and epidemiologic associations. Am J Med 1983, 74, (1), 14-22.
68. Srinivasan, S.; Hoffman, N. G.; Morgan, M. T.; Matsen, F. A.; Fiedler, T. L.; Hall, R. W.; Ross, F. J.; McCoy, C. O.; Bumgarner,
R.; Marrazzo, J. M.; Fredricks, D. N., Bacterial communities in women with bacterial vaginosis: high resolution phy-
logenetic analyses reveal relationships of microbiota to clinical criteria. PLoS One 2012, 7, (6), e37818.
69. Van Schalkwyk, J.; Yudin, M. H., Vulvovaginitis: screening for and management of trichomoniasis, vulvovaginal can-
didiasis, and bacterial vaginosis. J Obstet Gynaecol Can 2015, 37, (3), 266-274.
70. Keane, F. E.; Maw, R.; Pritchard, C.; Ison, C. A., Methods employed by genitourinary medicine clinics in the United
Kingdom to diagnose bacterial vaginosis. Sex Transm Infect 2005, 81, (2), 155-7.
71. Brusselaers, N.; Shrestha, S.; van de Wijgert, J.; Verstraelen, H., Vaginal dysbiosis and the risk of human papillomavirus
and cervical cancer: systematic review and meta-analysis. Am J Obstet Gynecol 2019, 221, (1), 9-18.e8.
72. Sobel, J. D.; Subramanian, C.; Foxman, B.; Fairfax, M.; Gygax, S. E., Mixed vaginitis-more than coinfection and with
therapeutic implications. Curr Infect Dis Rep 2013, 15, (2), 104-8.
73. Bradshaw, C. S.; Morton, A. N.; Garland, S. M.; Horvath, L. B.; Kuzevska, I.; Fairley, C. K., Evaluation of a point-of-care test,
BVBlue, and clinical and laboratory criteria for diagnosis of bacterial vaginosis. J Clin Microbiol 2005, 43, (3), 1304-8.
74. Sumeksri, P.; Kopraser t, C.; Panichkul, S., BVBLUE test for diagnosis of bacterial vaginosis in pregnant women attend-
ing antenatal care at Phramongkutklao Hospital. J Med Assoc Thai 2005, 88 Suppl 3, S7-13.
58
75. Miller, L., Can Fem Exam Card Use Facilitate Bacterial Vaginosis Diagnosis on Day of Abortion to Prevent Postabortion
Endometritis? Obstetrics & Gynecology 2001, 97, (4), 58S-59S.
76. Schwebke, J. R.; Taylor, S. N.; Ackerman, R.; Schlaberg, R.; Quigley, N. B.; Gaydos, C. A.; Chavoustie, S. E.; Nyirjesy, P.; Remil-
lard, C. V.; Estes, P.; McKinney, B.; Getman, D. K.; Clark, C., Clinical Validation of the Aptima Bacterial Vaginosis and Aptima
Candida/Trichomonas Vaginitis Assays: Results from a Prospective Multicenter Clinical Study. J Clin Microbiol 2020, 58, (2).
77. Thompson, A.; Timm, K.; Borders, N.; Montoya, L.; Culbreath, K., Diagnostic performance of two molecular assays for
the detection of vaginitis in symptomatic women. Eur J Clin Microbiol Infect Dis 2020, 39, (1), 39-44.
78. Mulhem, E.; Boyanton, B. L., Jr.; Robinson-Dunn, B.; Ebert, C.; Dzebo, R., Performance of the Affirm VP-III using residual
vaginal discharge collected from the speculum to characterize vaginitis in symptomatic women. J Low Genit Tract
Dis 2014, 18, (4), 344-6.
79. Cartwright, C. P.; Lembke, B. D.; Ramachandran, K.; Body, B. A.; Nye, M. B.; Rivers, C. A.; Schwebke, J. R., Comparison
of nucleic acid amplification assays with BD affirm VPIII for diagnosis of vaginitis in symptomatic women. J Clin
Microbiol 2013, 51, (11), 3694-9.
80. Brown, H. L.; Fuller, D. D.; Jasper, L. T.; Davis, T. E.; Wright, J. D., Clinical evaluation of affirm VPIII in the detection and
identification of Trichomonas vaginalis, Gardnerella vaginalis, and Candida species in vaginitis/vaginosis. Infect Dis
Obstet Gynecol 2004, 12, (1), 17-21.
81. Byun, S. W.; Park, Y. J.; Hur, S. Y., Affirm VPIII microbial identification test can be used to detect gardnerella vaginalis, Can-
dida albicans and trichomonas vaginalis microbial infections in Korean women. J Obstet Gynaecol Res 2016, 42, (4), 422-6.
82. Sheiness, D.; Dix, K.; Watanabe, S.; Hillier, S. L., High levels of Gardnerella vaginalis detected with an oligonucleotide
probe combined with elevated pH as a diagnostic indicator of bacterial vaginosis. J Clin Microbiol 1992, 30, (3), 642-8.
83. Gazi, H.; Degerli, K.; Kurt, O.; Teker, A.; Uyar, Y.; Caglar, H.; Kurutepe, S.; Surucuoglu, S., Use of DNA hybridization test
for diagnosing bacterial vaginosis in women with symptoms suggestive of infection. Apmis 2006, 114, (11), 784-7.
84. Andrea, S. B.; Chapin, K. C., Comparison of Aptima Trichomonas vaginalis transcription-mediated amplification assay
and BD affirm VPIII for detection of T. vaginalis in symptomatic women: performance parameters and epidemiolog-
ical implications. J Clin Microbiol 2011, 49, (3), 866-9.
85. Richter, S. S.; Otiso, J.; Goje, O. J.; Vogel, S.; Aebly, J.; Keller, G.; Van Heule, H.; Wehn, D.; Stephens, A. L.; Zanotti, S.; Johnson, T.;
Leal, S. M.; Procop, G. W., Prospective Evaluation of Molecular Assays for Diagnosis of Vaginitis. J Clin Microbiol 2019, 58, (1).
86. Witt, A.; Petricevic, L.; Kaufmann, U.; Gregor, H.; Kiss, H., DNA hybridization test: rapid diagnostic tool for excluding
bacterial vaginosis in pregnant women with symptoms suggestive of infection. J Clin Microbiol 2002, 40, (8), 3057-9.
87. Dessai, F.; Nyirenda, M.; Sebitloane, M.; Abbai, N., Diagnostic evaluation of the BD Affirm VPIII assay as a point-of-care
test for the diagnosis of bacterial vaginosis, trichomoniasis and candidiasis. Int J STD AIDS 2020, 31, (4), 303-311.
88. Rivers, C. A.; Lee, J. Y.; Sharples, N.; Ledeboer, N. A.; Schwebke, J. R., ESwab as an optional collection device for use
with the Affirm VPIII microbial test system. J Clin Microbiol 2014, 52, (5), 1698-700.
89. Rompalo, A. M.; Castleberry, N.; Widdice, L.; Schulkin, J.; Gaydos, C. A., Patterns of point-of-care test use among
obstetricians and gynaecologists in the US. Sex Health 2018, 15, (4), 318-324.
90. Brotman, R. M.; Ravel, J., Ready or not: the molecular diagnosis of bacterial vaginosis. Clin Infect Dis 2008, 47, (1), 44-6.
91. Schwebke, J. R.; Gaydos, C. A.; Davis, T.; Marrazzo, J.; Furgerson, D.; Taylor, S. N.; Smith, B.; Bachmann, L. H.; Ackerman,
R.; Spurrell, T.; Ferris, D.; Burnham, C. A.; Reno, H.; Lebed, J.; Eisenberg, D.; Kerndt, P.; Philip, S.; Jordan, J.; Quigley, N.,
Clinical Evaluation of the Cepheid Xpert TV Assay for Detection of Trichomonas vaginalis with Prospectively Collect-
ed Specimens from Men and Women. J Clin Microbiol 2018, 56, (2).
92. De Salazar, A.; Espadafor, B.; Fuentes-López, A.; Barrientos-Durán, A.; Salvador, L.; Álvarez, M.; García, F., Comparison
between Aptima Assays (Hologic) and the Allplex STI Essential Assay (Seegene) for the diagnosis of Sexually trans-
mitted infections. PLoS One 2019, 14, (9), e0222439.
93. Drew, R. J.; Murphy, T.; Broderick, D.; O’Gorman, J.; Eogan, M., An interpretation algorithm for molecular diagnosis of
bacterial vaginosis in a maternity hospital using machine learning: proof-of-concept study. Diagn Microbiol Infect
Dis 2020, 96, (2), 114950.
94. Hilbert, D. W.; Smith, W. L.; Chadwick, S. G.; Toner, G.; Mordechai, E.; Adelson, M. E.; Aguin, T. J.; Sobel, J. D.; Gygax,
S. E., Development and Validation of a Highly Accurate Quantitative Real-Time PCR Assay for Diagnosis of Bacterial
Vaginosis. J Clin Microbiol 2016, 54, (4), 1017-24.
95. Rumyantseva, T. A.; Bellen, G.; Savochkina, Y. A.; Guschin, A. E.; Donders, G. G., Diagnosis of aerobic vaginitis by quan-
titative real-time PCR. Arch Gynecol Obstet 2016, 294, (1), 109-14.
96. Muzny, C. A.; Balkus, J.; Mitchell, C.; Sobel, J. D.; Workowski, K.; Marrazzo, J.; Schwebke, J. R., Diagnosis and Manage-
ment of Bacterial Vaginosis: Summary of Evidence Reviewed for the 2021 Centers for Disease Control and Preven-
tion Sexually Transmitted Infections Treatment Guidelines. Clin Infect Dis 2022, 74, (Suppl_2), S144-s151.
59
97. Van den Munckhof, E. H. A.; van Sitter, R. L.; Boers, K. E.; Lamont, R. F.; Te Witt, R.; le Cessie, S.; Knetsch, C. W.; van Doorn,
L. J.; Quint, W. G. V.; Molijn, A.; Leverstein-van Hall, M. A., Comparison of Amsel criteria, Nugent score, culture and two
CE-IVDmarked quantitative real-time PCRs with microbiota analysis for the diagnosis of bacterial vaginosis. Eur J Clin
Microbiol Infect Dis 2019, 38, (5), 959-966.
98. Van Der Pol, B.; Daniel, G.; Kodsi, S.; Paradis, S.; Cooper, C. K., Molecular-based Testing for Sexually Transmitted Infec-
tions Using Samples Previously Collected for Vaginitis Diagnosis. Clin Infect Dis 2019, 68, (3), 375-381.
99. Ackerman, S. J.; Knight, T.; Wahl, P. M.; Cartwright, C. P., Health care utilization and costs following amplified versus
non-amplified molecular probe testing for symptomatic patients with suspected vulvovaginitis: a US commercial
payer population. Clinicoecon Outcomes Res 2019, 11, 179-189.
100. Broache, M.; Cammarata, C. L.; Stonebraker, E.; Eckert, K.; Van Der Pol, B.; Taylor, S. N., Performance of a Vaginal Panel
Assay Compared With the Clinical Diagnosis of Vaginitis. Obstet Gynecol 2021, 138, (6), 853-859.
101. Lynch, T.; Peirano, G.; Lloyd, T.; Read, R.; Carter, J.; Chu, A.; Shaman, J. A.; Jarvis, J. P.; Diamond, E.; Ijaz, U. Z.; Church,
D., Molecular Diagnosis of Vaginitis: Comparing Quantitative PCR and Microbiome Profiling Approaches to Current
Microscopy Scoring. J Clin Microbiol 2019, 57, (9).
102. Vieira-Baptista, P.; Eleutério Jr., J., Diagnosis of vaginitis: time to improve and move on. DST - J bras Doenças Sex
Transm 2020, 32, (e203214), 1-3.
103. Sherrard, J., Evaluation of the BD MAX™ Vaginal Panel for the detection of vaginal infections in a sexual health service
in the UK. Int J STD AIDS 2019, 30, (4), 411-414.
104. Aguirre-Quiñonero, A.; Castillo-Sedano, I. S.; Calvo-Muro, F.; Canut-Blasco, A., Accuracy of the BD MAX™ vaginal
panel in the diagnosis of infectious vaginitis. Eur J Clin Microbiol Infect Dis 2019, 38, (5), 877-882.
105. Danby, C. S.; Althouse, A. D.; Hillier, S. L.; Wiesenfeld, H. C., Nucleic Acid Amplification Testing Compared With Cul-
tures, Gram Stain, and Microscopy in the Diagnosis of Vaginitis. J Low Genit Tract Dis 2021, 25, (1), 76-80.
60
61
3.1
Introduction
The female genital tract consists of a hormonally driven tissue structure with estrogenic hormones
driving trophic and maturing effects on the epithelium. Even before the vagina was thought
of as an ecosystem with its resident microbial communities, microscopic and culture-based ob-
servations indicated lactobacilli colonization appeared abundant after an infant became colo-
nized, declined until puberty, became prominent through the reproductive years, and declined
in menopause. Modern molecular methods have opened a range of new possibilities for the
characterization of the vaginal microbiota (VMB), allowing us to not only establish which
microorganisms are present, but also to begin to understand their functional properties.1, 2
Growing evidence suggests that low diversity, Lactobacillus spp. dominated VMB is associat-
ed with lower inflammation, and that this is protective. On the other hand, non-Lactobacil-
lus spp.-dominated, higher diversity VMB (sometimes termed “dysbiotic”) is associated with
risk of infections (including human immunodeficiency virus [HIV]) and, possibly, obstetric
complications.3-5 Vaginal dysbiosis may be physiological for some women, or pathological,
depending on the interplay of metabolic and microbial factors.6
The clinical syndrome that is currently known as bacterial vaginosis (BV), involving multiple
bacterial species which vary from woman to woman, has been extensively studied for the
last six decades. BV is a polymicrobial disorder of the vaginal microbiome that is character-
ized by the absence of protective lactobacilli. The most frequently detected bacterial taxa in-
clude: Gardnerella spp., Mycoplasma hominis, Fannyhessea (Atopobium) vaginae, Bacteroides,
Clostridiale, Fusobacterium spp., Mobiluncus spp., Peptostreptococcus spp., Porphyromonas
spp., Prevotella spp., and others that have been described and may cause dysbiosis. Dysbi-
osis may be generated when conditions disrupt, modify, reduce, block, fluctuate or deplete
the dominant lactobacilli.7
3
BACTERIAL VAGINOSIS
(alphabetical order)
Jacob Bornstein
Catriona Bradshaw
Erica Plummer
Koray Gorkem Sacinti
Francesco de Seta
Ryan Sobel
Colleen K. Stockdale
Gary Ventolini
Hans Verstraelen
Pedro Vieira-Baptista
62
The prevalence of these communities, with a paucity of Lactobacillus spp. varies among
women, and epidemiological studies have associated them with an increased risk of adverse
health outcomes: preterm labor and birth (PTB), premature rupture of membranes (PROM),
chorio-amnionitis, funisitis, post abortion infections, and increased risk of acquiring sexually
transmitted infections (STIs).4, 5 The mechanisms that drive these associations are yet to be
described in detail, with few studies establishing causative relationships. Despite advances
in our understanding of BV, it remains an enigmatic condition. While the associated clinical
symptoms are relatively uncomplicated and easily measured, the fact that not all symptoms
occur in all women diagnosed with BV remains problematic. This is not surprising given the
complexity of the vaginal microbiome, host immunity, and the variability in individual re-
sponses to potential inflammatory mediators produced by an array of microorganisms.8
The link between non-Lactobacillus spp.-dominated high diversity VMB/BV and a variety of adverse
outcomes for women’s sexual and reproductive health has been well-established. This suggests
that treatment or prevention of vaginal dysbiosis/BV may improve women’s health outcomes.
3.2
Etiology and physiopathology
BV is considered a polymicrobial dysbiosis of the VMB. The optimal VMB of reproductive
aged women is typically dominated by lactic acid producing lactobacilli which maintain
a vaginal pH<4.7. In contrast, BV is characterized by an increase in the load of facultative
and strict anaerobic bacteria, a reduction in beneficial lactobacilli, and a corresponding in-
crease in vaginal pH.9-14 Although the exact etiological agent(s) responsible for BV is(are)
not known, cultivation studies and more recently, molecular studies, have identified a
large number of bacteria that are associated with BV, collectively referred to as BV-associ-
ated bacteria (BVAB). These organisms include Gardnerella spp., Prevotella spp., F. vaginae,
Mobiluncus spp., Megasphaera spp., Sneathia spp., among others.11, 12, 15-17 Of note, specific
Gardnerella spp. are thought to play a key role in BV pathogenesis, potentially as a founder
or initiating organism.18-21 Until recently, Gardnerella vaginalis was the only species present
in the Gardnerella genus. In 2019, thirteen Gardnerella species were proposed;22 three new
species (G. swidsinskii, G. piotii and G. leopoldii) were given official taxonomic standing, and
the description of G. vaginalis was narrowed and amended. To reflect the current taxonomy,
Gardnerella spp. will be used below in place of G. vaginalis. Also, recently, Atopobium vaginae
was reclassified into the new genus Fannyhessea.23
Gardnerella spp. have been shown to adhere to vaginal epithelial cells and initiate biofilm
formation, which is thought to be a key factor in BV pathogenesis.24-26 Gardnerella spp. are of-
ten the predominant species present in BV-biofilms, and have demonstrated ability to form
a biofilm in acidic environments,27 which further supports its role as an integral organism
involved in the initiation of BV. However, synergistic relationships between Gardnerella spp.
and other BVAB exist.18 It is hypothesized that Gardnerella spp. act to lower the oxidation-reduc-
tion potential of the vaginal environment which promotes growth of strict anaerobic bacteria
including Prevotella spp. and F. vaginae.18, 20 Production of amino acids by Gardnerella spp. fur-
63
ther enhances the growth of Prevotella spp., which in turn produces ammonia, which enhances
growth of Gardnerella spp..28 Additionally, F. vaginae and Prevotella spp. are also present in
BV-biofilms alongside Gardnerella spp..29 Production of virulence factors (e.g. sialidase) by
Gardnerella spp. and Prevotella spp. degrade the protective cervicovaginal mucosa, further
enhancing biofilm formation and facilitate attachment of other BVAB,29-31 which contribute
to BV symptoms and sequelae.32 Exfoliation of vaginal epithelial cells produces the clue cells
that are characteristic of BV,33 and the increased load of anaerobic bacteria is associated with
production of amines, which contributes to the malodorous discharge observed in BV.34-36
Importantly, the event that triggers the vaginal dysbiosis observed in BV is not well understood.
It is not known if BV results from acquisition of a single organism (e.g. specific Gardnerella spp.)
or a polymicrobial consortium, or is a consequence of overgrowth of BVAB in response to spe-
cific host or behavioral factors.37 Importantly, both epidemiological and molecular data indi-
cate that sexual transmission is involved in both the acquisition and recurrence of BV.38-45 The
epidemiological profile of BV is similar to that of STIs, with a meta-analysis of 43 studies finding
that BV was associated with inconsistent condom use, as well as new and increased number of
sexual partners.46 BV is associated with early age of sexual debut,47, 48 and is rare among women
without a history of coital or noncoital sexual contact.48 Furthermore, risk factors for BV acqui-
sition among women with female partners include having a sexual partner with a history of
BV, BV symptoms or microbiologically confirmed BV.39, 45 BV has also been associated with
other behavioral practices including smoking49-52 and intravaginal douching.53-56
3.3
Prevalence and epidemiology
Global and regional estimates of bacterial vaginosis prevalence
The prevalence of BV varies widely across countries and between different population
groups and is influenced by differences in diagnostic and sampling methodology. In a re-
cent systematic review and meta-analysis (122 publications, up to 2017), the global preva-
lence of BV among reproductive-aged women in the general population was high, ranging
from 23 to 29%, with marked racial disparities.57 Pooled estimates by geographical region
included: Europe and Central Asia, 23%; East Asia and Pacific, 24%; Latin America and Car-
ibbean, 24%; Middle East and North Africa, 25%; sub-Saharan Africa, 25%; North America,
27%; South Asia, 29%).57 Within sub-Saharan Africa, BV prevalence was lower in Western and
Central Africa (20.6%; 95% confidence interval [CI], 6.1–40.6%) than in Southern and Eastern
Africa (33.3%; 95% CI 17.4–51.5%), although this was not statistically significant. The review
found BV prevalence varied with ethnicity within specific geographical regions. For exam-
ple, within North America overall BV prevalence was 27% (95% CI 24- 31%), but prevalence
estimates were higher in Black and Hispanic women (33% and 31%, respectively) compared
to White and Asian women (23% and 11%, respectively).57 Overall, there was an approxi-
mately 2-fold higher BV prevalence among majority Black populations when compared to
majority non-Black populations in this meta-analysis (46.5%; 95% CI 37.5–55.6% vs. 21.3%;
95% CI 16.7–26.3%; p<0.001).57
64
A meta-analysis among women participating in HIV prevention studies (n=18) across
three primary region and population groups in sub-Saharan Africa (South Africa commu-
nity-based, Southern/Eastern Africa community-based, and Eastern Africa higher-risk pop-
ulations), reported summary estimates for BV prevalence in excess of 30%.58 Among 15 to
24-year-olds in South African community-based populations, the summary estimate for BV
prevalence was 42.1% (95% CI 35.6-49.0%), in Southern and Eastern African clinic and com-
munity based populations it was 35.2% (95% CI 27.7-43.6%), and in the higher-risk popula-
tions in Eastern Africa, BV prevalence was 49.5% (95% CI 42.2-56.8%). Prevalence was similar
among women aged 25 to 49 years, with high heterogeneity across studies.58
Bacterial vaginosis prevalence among pregnant women
There have been three large meta-analyses examining BV prevalence in pregnant wom-
en. In a systematic review and meta-analysis of malaria, STIs and BV prevalence among
pregnant women attending antenatal care facilities in sub-Saharan Africa from 1990-2011
(340,904 women), the burden of BV was higher than that of any STI.59 The pooled BV prev-
alence estimate in East and Southern Africa was 50.8% (43.3-58.4%; n=4280) and in West
and Central Africa was 37.6% (18.0-57.2%; n=1208).59 A more recent meta-analysis of BV
prevalence during pregnancy in sub-Saharan Africa, which included publications dating
from 2015 to 2020 (48 studies, n=5042 women), yielded a pooled estimate of BV prevalence
of 36.6% (27.1–46.6%).60 Pooled BV prevalence estimates ranged from 28.5% (24.5–32.8%,
n=1030) in Eastern Africa to 52.4% (33.5–70.9%, n=2305) in Southern Africa.60 A subgroup
analysis of pregnant women in the large global meta-analysis by Peebles et al. also reported
data outside Africa, with pooled BV prevalence estimates in pregnant women ranging from
11.7% in South Asia (95% CI 9.0–14.7%) to 33.2% in Latin America and the Caribbean (95%
CI 14.8–54.7%). Within the United States, they confirmed the racial and ethnic disparities ob-
served in non-pregnant women. Prevalence of BV was highest among Black (49.0%; 95% CI
40.2–57.8%) and Hispanic pregnant women (42.7%; 95% CI 36.4–49.1%) and lowest among
Asian (20.3%; 95% CI 5.4–41.2%) and White pregnant women (19.9%; 95% CI 8.0–35.5%).61
Overall, there are limited pooled BV prevalence estimates from the Middle East for pregnant
women, but a systematic review and meta-analysis of BV prevalence in Iran, that included
studies up to 2017, reported the prevalence of BV in pregnant women to be 16.5% (95% CI
12.5–21.6%), compared to 28% (95% CI 15.1–45.9%) in non-pregnant women.62
Bacterial vaginosis prevalence among other populations/sub-groups
of women
BV prevalence estimates are generally reported to be high in women with female partners.
In a systematic review of BV prevalence among lesbian women, BV was the most frequent
genital infection reported, and prevalence ranged from 25.7 to 42.8%.63
Peebles et al. undertook a sub-group analysis in their global meta-analysis, and found that
BV prevalence was approximately 20% higher (33.5%; 95% CI 30.5-40.7%) in women with
female partners than in women of the general population (p=0.007).61 Peebles et al. also un-
65
dertook a sub-group analysis examining BV prevalence among women living with HIV (two
studies from Southeast Asia and four from sub-Saharan Africa). Relative to women of the
general population, BV prevalence was also higher among women living with HIV (35.6%;
95% CI 25.7–46.2% vs. 25.6; 95% CI 22.6–28.7%; p= 0.054).61
A recent systematic review and meta-analysis of women undergoing in vitro fertilization
(IVF) found the overall prevalence of vaginal dysbiosis (BV by microscopy or dysbiosis by
molecular methods including qPCR, 16S rRNA gene sequencing and interspace profiling)
to be 18% (95% CI 17–19%) (644/3543) with considerable heterogeneity across studies
(prevalence varying from 4 to 44%).64 Importantly, there was no significant difference in the
BV prevalence ratio between microscopy and molecular methods (0.87; 95% CI 0.74–1.02).
Studies based on microscopy (n=13) found an overall prevalence of 17% (517/3091), and
studies using molecular methods found a prevalence of 19% (171/889).64
While current diagnostic methods for BV are impacted by the menopausal status, a meta-analysis
of all published studies reporting BV prevalence in post-menopausal women up to 2020, found
that BV prevalence ranged from 2.0 to 57.1%, with a summary estimate of 16.93% (95% CI 8.5–
27.4%). There was significant heterogeneity between studies and quality varied considerably.65
3.4
Risk factors
Table 3.1 lists common risk factors for BV. BV is more common among African American
women.66 It is sexually associated, including sexual activity with male and female partners.
Fethers et al. identified sexual contact with new and multiple male and female partners to
be associated with an increased risk of BV in a systematic review and meta-analysis.46 Fur-
ther studies have noted BV is highly prevalent (25 to 50 percent) in females who have sex
with females and is associated with having a female partner with symptomatic BV, shared
used of sex toys, and increasing numbers of female sexual partners.39, 45, 67-70
Presence of STIs appears to be associated with an increased prevalence of BV and presence
of BV may also be a risk factor for HIV and other STIs.71-78
Other risk factors identified include inconsistent condom use, cigarette smoking, douching,
and obesity.50, 53, 55, 79-83 Past or current tobacco use has been reported to modify the vaginal
milieu increasing bacterial virulence, as well as promoting an antiestrogen environment
with additional vaginal amines.84 Vaginal douching has been connected with modifications
of the vaginal milieu and optimal VMB therefore, favoring an increased risk of BV.85-87 However,
cessation of douching does not seem to promote the return to a lactobacilli dominated VMB.88
The use of copper intrauterine devices has been associated with an elevated risk of BV.85
66
TABLE 3.1 Risk factors for bacterial vaginosis.
HIV – human immunodeficiency virus; BV – bacterial vaginosis; STI – sexually transmitted infection
African American race
Sexual activity with male and female partners
Multiple sexual partners
New sexual partner
Female sexual partner with BV symptoms
Shared use of sex toys
Copper intrauterine devices
Douching
Cigarette smoking
Obesity
Inconsistent condom use
Previous or concurrent STI
3.5
Complications
Major sequelae of BV include an increased risk of PTB, postpartum endometritis, post hyster-
ectomy vaginal cuff cellulitis, post abortal infection, pelvic inflammatory disease (PID), and
STIs (including increased risk of HIV acquisition and transmission).
Preterm birth
While BV has been associated with PTB and greater risk of early pregnancy loss, early identi-
fication and treatment of asymptomatic women has failed to impact those rates.
A 2020 meta-analysis of studies concerning pregnancy outcomes according to vaginal mi-
crobiota, in sub-Saharan Africa, showed that an association between BV and PTB was not
systematically reported (positive in seven out of nine studies).86 Another recent meta-anal-
ysis including 44 studies found no difference in the incidence of PTB and related outcomes
from treatment of asymptomatic women with BV in a general obstetric population.87 Thus,
the United States Preventative Services Task Force recommends against screening for BV in
pregnant women not at increased risk for preterm delivery and concludes there is insuffi-
cient evidence to assess the benefits and harms of screening for BV in pregnant persons at
increased risk for preterm delivery.88
Endometritis/ postpartum fever
As with PTB, BV has been associated with adverse postpartum complications including
endometritis and postpartum fever. A Cochrane Systematic Review found that antibiotic
prophylaxis given during the second or third trimester reduced the risk of postpartum endo-
metritis when routinely administered to all pregnant women.89 However, the authors noted
67
substantial bias due to the small numbers of studies available for analysis and high rate of
loss to follow-up and concluded there was insufficient evidence to support routine use of
antibiotics during pregnancy to prevent infectious adverse effects.
Post hysterectomy vaginal cuff cellulitis
Vaginal bacterial contamination is a major cause of febrile morbidity including post vaginal
cuff cellulitis and pelvic abscess following total hysterectomy.90 Thus, current standard of
therapy includes vaginal cleaning with hexachlorophene or povidone-iodine in addition to
the standard surgical preparation and prophylactic antibiotics.90 Despite these measures,
infections related to vaginal contamination persist and there remains a need to improve
vaginal antisepsis for hysterectomy.
Post abortal infection
Risk factors for post abortal infection include history of PID, lower genital tract infection,
BV, and age less than 20.91 As with hysterectomy, vaginal bacterial contamination is a major
cause of febrile morbidity with operative vaginal procedures despite current standard of
therapy to reduce bacterial load (including prophylactic antibiotics and standard procedural
preparation).
Pelvic inflammatory disease
PID has a multimicrobial etiology. BV (by Nugent score) has been associated with clinical as
well as subclinical PID infections.92, 93
Other sexually transmitted infections
BV is associated with increased risk of acquisition of HIV and other STIs.76-78, 94 Alteration in the
vaginal microenvironment, by the lack of hydrogen peroxide producing lactobacilli in wom-
en with BV, has been postulated to increase the risk for STI acquisition. BV is frequently seen
as a co-infection with cervical and vaginal STIs.95, 96 Schwebke et al. conducted a randomized
trial of metronidazole gel vs. observation in women with asymptomatic BV and found sig-
nificantly fewer cases of Chlamydia in the treated group (1.58 vs. 2.29 per person-year).95
3.6
Signs and symptoms
BV is confined to an asymptomatic state in at least half of the cases, though it still is a leading
cause of vulvovaginal symptoms worldwide.97 Unlike single pathogen vaginal infections, BV
is in fact thought of as a set of common clinical signs and symptoms that can be provoked
by a plethora of bacterial species and communities, with different bacterial species showing
different associations with presenting signs and symptoms, hence explaining considerable
variability in clinical presentation.8, 16 Key symptoms in BV, if present, do not seem to result pri-
marily from inflammation, but rather from bacterial strategies deployed by the BV microbiota to
colonize the vaginal niche. Breakdown of the protein backbone as well as of the sugar coating of
68
cervical mucins through bacterial mucinase action is thought to underlie at least in part vaginal
discharge in this setting.98 Symptomatic patients with BV may typically complain of copious vag-
inal discharge, that is thin, off-white to greyish, and sometimes described as foamy.99 (Figure 3.1)
Figure 3.1 Typical discharge associated with bacterial vaginosis
The production of biogenic polyamines by a few bacterial species associated with BV, relates
to the “fishy odor” or “fishy smelling” vaginal discharge, which is as a rather specific symptom
of BV.100, 101 Symptomatic patients may describe an even stronger smell after sexual intercourse
and some perceive the odor also to be more noticeable during and following their period.
Notably, a lack of perceived odor in patients with vulvovaginal symptoms makes the diagno-
sis of BV rather unlikely.101 Conversely, symptoms other than fishy-smelling discharge, such as
itch, dyspareunia, and dysuria, are not typically expected in BV, though such complaints may be
present with mixed vaginitis, i.e. combined BV and vulvovaginal candidiasis.102 To the attending
health care provider, there are usually no overt or even appreciable signs of disease in patients
with BV, though the typical “discharge” and “odor” may also be apparent as signs on clinical exam.
3.7
Diagnosis
While patient history and symptoms of “fishy odor” or “fishy smelling” vaginal discharge in
particular, may be highly suggestive of BV, ultimate diagnosis will rely on microbiological
confirmation, in addition to clinical presentation. Several point-of-care tests (POCTs) as well
laboratory tests for diagnosis of BV are available to that purpose.
69
Clinical diagnosis
In clinical practice, the diagnostic criteria originally described by Amsel et al. have proven a
useful diagnostic tool.33 Clinical diagnosis of BV according to Amsel’s criteria is convention-
ally made if three of the four following signs are present: (1) adherent and homogenous
grayish-white vaginal discharge; (2) a vaginal pH exceeding a value of 4.5; (3) the presence
of so-called clue cells (vaginal epithelial cells with such a heavy coating of BVAB that the
peripheral borders are obscured) on saline wet mount (Figure 3.2); and (4) a fishy or amine
odor after the addition of a 10% potassium hydroxide solution (positive whiff or sniff test).33
Despite being widely used, Amsel’s criteria have been criticized, particularly because the
appearance of the discharge and, to some extent, the appraisal of the odor, tend to be sub-
jective, difficult to standardize, and hence prone to mis-
diagnosis. It has been suggested however, that Amsel’s
approach can be simplified to a modified combination
of merely two criteria, without significant loss of overall
sensitivity and specificity.103-105 Elevated vaginal pH (>4.5)
in particular, is consistently found as the most sensitive
of all Amsel’s criteria. It is important to acknowledge that
many other factors may alter vaginal pH and/or interfere
with vaginal pH assessment, notably menses/blood (even
when not obvious on clinical exam), but also semen, as
well as any inserted product (lubricants, creams, suppos-
itories, etc.). Such artifacts generally involve a pH eleva-
tion and hence threaten specificity, rather than diagnostic
sensitivity. Positioning of the pH strip close the external
os of the cervix and close to the cervical mucus flow may
also distort the picture. The presence of clue cells, in turn,
is considered the single most specific predictor of BV.97
The clue cell criterion is often cited as the presence of clue
cells representing ≥20% of epithelial cells on microscopic
examination, although the 20% cut-off was not originally
mentioned by Amsel et al., but added later on as to in-
crease specificity and overall accuracy.33, 104, 106
Amsel’s approach has also been critiqued for its low sen-
sitivity when compared to Gram-stain based or molecular
diagnosis of BV. Across comparative studies, a wide range
of sensitivities and specificities has been reported, with
a sensitivity as low as 37% in one study and as high as
98.2% for the presence of clue cells on wet mount exami-
nation alone.103, 107 Clearly, the performance of Amsel’s cri-
teria, whether modified or not, is highly dependent on the
assessor’s experience, time, and equipment. Furthermore,
it should be acknowledged that Amsel’s criteria have not
Figure 3.2 Wet mount microscopy
(400x, phase contrast).
A-C– Bacterial vaginosis: absence of lacto-
bacilli, granular microbiota
and presence of clue cells (seen in A)
70
been developed as a screening tool, but rather as a diagnostic aid in case of vulvovaginal
symptoms suggestive for BV. Overall, in the absence of molecular or biochemic POCTs, Amsel’s
clinical criteria remain the best option for in-office testing for BV. Clinical diagnosis of BV can be
quickly obtained at very low cost, but does require the presence of a microscope and micros-
copy skills. In the presence of the latter, the diagnosis can be established using wet mount mi-
croscopy, with sensitivity and specificity ranging between 82-100% and 93-97%, respectively.108
Gram-stain diagnosis
Gram-stain based diagnosis has been the mainstay of BV diagnosis, especially in research
settings, and is broadly accepted as the gold standard in that respect. (Figure 3.3)
This approach has several advantages, including a high frequency of interpretable results,
a permanent record, and low cost. Gram-stained vaginal smears can also be interpreted
repeatedly or independently by more than one assessor, thereby increasing diagnostic re-
liability.97 Gram-stain diagnosis of BV does, however, require a specific laboratory setting
and considerable skill, experience, and time. The most widely performed Gram-stain based
method is the scoring system developed by Nugent et al..14 Briefly, the Nugent scoring sys-
tem accounts for three bacterial cell morphotypes – that is, Lactobacillus spp. morphotypes
(large Gram-positive rods), Gardnerella
spp. and Bacteroides spp. morphotypes
(small Gram-variable or Gram-negative
rods), and curved Gram-variable rods
(such as Mobiluncus spp.). Although
the taxonomic assignment of the mor-
photypes has been revised since then,
the overall approach remains valid.109
In particular, Srinivasan et al. found that
the “Bacteroides morphotype,” was pri-
marily represented by Prevotella spp.
and Porphyromonas spp., whereas Mo-
biluncus spp. morphotypes are more
likely BVAB1 (Candidatus Lachnocurva
vaginae). 109, 110 Based on the abundance
of each of the aforementioned morpho-
types per oil immersion field, quantitated
from 0 to 4+, a summary score is obtained
which equates the overall Nugent score
(See section 2.5 and table 2.2).
Diagnosis of BV is accepted when a
score of 7 or higher is obtained. A score
of 4–6 corresponds to intermediate vaginal microbiota, and a score of 0–3 is considered
to represent non-BV microbiota. Hay and Ison developed a similar, simplified scoring system,
which is known as the Ison-Hay criteria, in which smears are graded qualitatively as normal (grade
Figure 3.3 Gram stain (1000x, oil immersion).
A and B– Bacterial vaginosis (clue cell seen in B)
71
I), intermediate (grade II), or consistent with BV (grade III), further complemented with two addi-
tional grades, i.e. grade 0 for epithelial cells only with no bacteria, and grade IV for Gram positive
cocci only. The Ison-Hay criteria tends to perform equally well, but is less widely used than the
Nugent system.111
Overall, the Nugent scoring system for Gram-stained vaginal smears shows a high degree of
accuracy and high reliability, as well as high intraobserver and interobserver reproducibility.
This does not imply however that the Nugent scoring system is without shortcomings. First-
ly, of concern is the defined lack of standardized pre-analytical and analytical conditions.
Forsum et al. have reported on that account: different sampling devices and procedures,
different ways of spreading the vaginal specimen on the glass slide, hence with variable
homogeneity of the sample and thickness of the smear, different fixation methods and time,
and differences in the area of the high-power oil immersion field at magnification ×1000, all
of which may affect Gram stain interpretation.112-114 Secondly, no definite criteria have been
proposed to distinguish between the three basic morphotypes handled in the Nugent scor-
ing system.114 It may be added here, that the significance of the intermediate Nugent (scores
4 to 6) or grade II Ison-Hay category remains undetermined, although studies have suggest-
ed that about one third to one half of women with this category actually do have BV.21, 115
Cultures
Culture of Gardnerella spp. has no role in BV diagnostics. Most laboratories will report to the
clinician who obtained a vaginal swab the results of Gram stain analysis as well as of culture
on general and more specific media. A positive culture for Gardnerella spp. per se does not
provide any information on the community state of the vaginal microbiota, as Gardnerella spp.
is commonly part of the latter, also in women who do not have BV or intermediate microbiota.
Point-of-care tests (non-molecular)
As the most commonly used diagnostic approaches, Amsel’s and Nugent’s method, require
time, skill, equipment, and experience, a defined need for rapid POCTs is perceived by many
healthcare providers workers. Several such tests have been developed and commercialized,
primarily in the US, though none are widely used. POCTs would ideally also allow for diag-
nosis of self-collected vaginal swabs or even self-diagnosis. Commercial POCTs typically rely
on the detection of metabolites, specifically biogenic polyamines, such as trimethylamine,
or short-chain fatty acids (SCFAs), either on the detection of the enzymes proline amino-
peptidase and sialidase produced by several bacteria in women with BV.21, 97 Among the
better documented POCTs in this respect are the OSOM BV Blue test (Genzyme Diagnostics,
Cambridge, MA, USA) and the FemExam card (Cooper Surgical, Shelton, CT, USA).
The OSOM BV Blue assay is a chromogenic dipstick test, based on the measurement of sial-
idase levels in vaginal fluid. The test is particularly fast, with results available within 10 min-
utes, and performs rather accurate, with reported sensitivities of 88 to 94% and specificities
of 91 to 98%, compared to Nugent and Amsel’s criteria, respectively.21, 97
The FemExam card, in turn, is a POCT that consists of two plastic cards, one for pH assessment
and detection of trimethylamine, and a second one for proline aminopeptidase measurement.
72
The FemExam card is even faster than the OSOM BV Blue test, with results within two minutes,
comparable sensitivity (91%), however with significantly lower specificity (estimated 61%).21, 97
Molecular diagnostics
Over the past two decades vaginal microbiome and BV research, has witnessed a marked
shift towards molecular characterization techniques, primarily 16S rRNA gene, and to a
smaller extent cpn60 gene amplification-based methods, such as next-generation sequenc-
ing. Molecular techniques will likely also replace existing in-office and laboratory BV diag-
nostics in the future. Current molecular diagnostic assays can be broadly divided in direct
probe assays and nucleic acid amplification techniques (NAATs).21 Their use is recommended
only in symptomatic women.116
Direct probe assays make use of DNA probes that directly bind bacterial sequences, i.e. with-
out an intermediate amplification step. The best-known example of a direct probe assay for
BV is the so-called Affirm VPIII assay (Becton Dickinson, Sparks, MD, USA), which can provide
results in less than one hour. This assay specifically targets Gardnerella spp. with a detection
limit of 5x105 colony forming units/mL of vaginal fluid. The Affirm VPIII assay performs well
in comparison to the detection of clue cells with a sensitivity of 90% and a specificity of 97%,
respectively, and in comparison with Nugent score-based BV diagnosis with a sensitivity of
94% and a specificity of 81%, respectively. A related test by the same company, Affirm VPIII
microbial identification test (Becton Dickinson, Sparks, MD, USA) allows for simultaneous diag-
nosis of other common causes of vaginitis, such as Candida spp. and Trichomonas vaginalis.21
Nucleic acid amplification tests (NAATs), in turn, include an amplification step in which a spe-
cific nucleic acid sequence is enzymatically exponentially multiplied, before being detected
by DNA probes. Hence, NAATs have very low detection limits and are theoretically capable
of detecting as little as one organism in a sample.21 Several such tests have been marketed.
These NAATs for BV diagnosis will typically target multiple BV-related species (positive pre-
dictors), and most one or more vaginal Lactobacillus spp., as negative predictors. Some of
these test have data validating the use of self sampling. (Table 3.2)
Differential diagnosis
An increased pH is not specific of BV; it can also be found in cases of trichomoniasis, vaginal
atrophy and aerobic vaginitis/desquamative inflammatory vaginitis (AV/DIV).
Moreover, patients with trichomoniasis, atrophy and AV/DIV frequently have symptoms and
findings of vaginal inflammation and dyspareunia. Women with BV usually lack these in-
flammatory symptoms and signs. Furthermore, parabasal cells are often increased in vaginal
atrophy or AV/DIV and it can be easily detected in a wet mount preparation.
At times, patients can present with mixed “infections”, such as, for example BV and Tricho-
monas vaginalis or Candida spp..
73
TABLE 3.2 Commercial nucleic acid amplification tests (NAATs) for BV (adapted from Coleman et al.
and Muzny C et al.21, 117). Not all available tests are listed and not all listed tests widely available.
BVAB – Bacterial Vaginosis Associated Bacteria. a Compared with a combination of Nugent score
and Amsel’s criteria; b Compared with Nugent score; c Clinician collected sample (similar data for
self-collected samples); d Compared with the BD Max™ Vaginal Panel; e personal communication
from Barbara Van Der Pol
Gardnerella
spp. F. vaginae Mobilluncus
spp. Megasphaera BVAB Lactoba-
cillus
Sensitivity
(%)
(95% CI)
Specificity
(%)
(95% CI)
NuSwab®
(Laboratory
Corporation
of America
Holdings,
NC, USA)
x x x x 96.7a92.2a
BD Max™
Vaginal
Panel
(Becton
Dickinson,
MD, USA)
x x x x x 90.5a
(88.3-
92.2)
85.8 a
(83-88.3)
MDL BV
Panel
(Medical
Diagnostic
Laboratory,
NJ, USA)
x x x x x 99a94a
Allplex™
Vaginitis
(Seegene,
Seoul,
Korea)
x x x x 91.7b
(86.49-
95.40)
86.6b
(83.57-
89.24)
Aptima®
BV
(Hologic,
MA, USA)
x x x 95.0a,c
(93.1–
96.4)
89.6a.c
(87.1–
91.6)
Xpert®
Xpress
MVP
(Cepheid,
CA, USA)e
x x x 93.8c,d
(91.5-
95.5%)
93.8% c,d
(92.0-
95.3)
3.8
Treatment
The treatment targets include symptom alleviation, infection prevention and control fol-
lowing surgery, and reduction of STIs. Regarding non-pregnant women, the established
advantages of treatment include alleviating vulvo vaginal symptoms. Currently, there is no
evidence to recommend the treatment for asymptomatic non-pregnant women.81 BV can
spontaneously clear without treatment in both pregnant and non-pregnant women.118, 119
Screening and treatment may, nevertheless, be explored in high-risk groups for STIs since it
74
has been shown to increase the risk of infection with HIV, HPV, herpes simplex virus (HSV) 2,
T. vaginalis, Chlamydia trachomatis, Neisseria gonorrhea, and Mycoplasma genitalium.
The principles of treatment for non-pregnant women;
drug selection, dosing, adverse effects, and efficacy
The prescribing rational should be based on cost-effectiveness, availability of alternatives,
adverse effects, and patient factors (request, previous response history). Oral and topical an-
tibiotics (metronidazole, clindamycin, tinidazole, and secnidazole) and antiseptics (dequa-
linium chloride) are available to treat BV. (Table 3.3) Although the cure rates are about 80% for
all medications and methods, relapses are frequent.120 When metronidazole or clindamycin are
not available, dequalinium chloride, tinidazole or secnidazole are acceptable alternatives.
Oral versus vaginal treatment
Both metronidazole and clindamycin are available in oral and vaginal forms. Oral treatment
produces greater systemic adverse effects, including headache, nausea, abdominal pain,
and diarrhea. The vaginal levels attained with topical treatment can be up to 30 times those
of oral medication. This results in cure rates comparable to or slightly greater than those
obtained by the oral route, with the added benefit of fewer side effects. 120
Metronidazole
Dosing: Oral 500 mg metronidazole twice daily for seven days or vaginal metronidazole
0.75% gel once daily for five days.116, 121 A novel single-dose 1.3% metronidazole gel is avail-
able.122 However, we suggest the multiday treatments because it remains unknown if the
1.3% single-day dose is as effective as the multiday oral or vaginal regimens.
Adverse effects: A metallic taste, nausea, neutropenia, elevated international normalized
ratio in patients receiving vitamin K antagonists (i.e. warfarin), peripheral neuropathy, and
candidiasis are all possible side effects of oral and vaginal metronidazole.123 Metronidazole
allergy is uncommon, presenting as a rash, urticaria, and pruritus. Compared to clindamycin,
metronidazole is less frequently related to Clostridioides difficile infection.124
Efficacy: The majority of comparative trials utilizing divided-dose oral regimens for one
week obtained cure rates greater than 90% in the first week and up to 80% in the fourth
week (based on Amsel criteria).120, 125-127
Special considerations:
There is less evidence that continuing treatment beyond seven days is beneficial.81
A novel metronidazole-loaded solid lipid nanoparticles vaginal emulgel showed a sub-
stantial therapeutic benefit for BV treatment in a randomized controlled trial.128
Clindamycin
Physicians should be aware of the potential for drug interactions and ensure the antibiotic
is effective; drug interactions may occur when any form of clindamycin is used with medi-
75
cations that impact the functioning of CYP3A4 (clarithromycin, erythromycin, rifampin, ta-
moxifen, glucocorticoids, etc.).
Dosing: The recommended regimen is 5 g of 2% clindamycin cream intravaginally for seven days.
Alternatives include clindamycin 300 mg twice a day orally for seven days or clindamycin 100 mg
vaginal suppositories for three days.116, 129 In some countries a formulation of clindamycin phos-
phate vaginal cream 2% is available; it is a sustained release formulation, used as a single dose.130
Adverse effects: Overgrowth of Candida spp. and gastrointestinal side effects are the most fre-
quently reported adverse effects, and pseudomembranous colitis has been rarely reported.
Efficacy: A meta-analysis of randomized studies, both comparative and placebo-controlled,
revealed the effectiveness of oral and vaginal clindamycin regimens.120
Special considerations:
Clindamycin cream is oil-based and has the potential to weaken latex condoms and dia-
phragms for five days following application.
Overview of second-line and alternative treatments
Dequalinium chloride
Dosing: One 10 mg vaginal tablet daily for six days is the recommended regimen.
Adverse effects: The majority of adverse effects were local reactions, including vulvovaginal
pruritus, vaginal discharge and burning sensation.Dequalinium chloride, unlike antibiotics,
is less toxic to lactobacilli and does not increase the risk of candidiasis.131
Efficacy: In one report the cure rate was non-inferior to those attained with clindamycin.132
It is not anticipated that bacteria can acquire resistance to it, and it is effective against causes
of vaginitis other than BV, making it at least partially beneficial for “mixed” infections. Never-
theless, long-term research on recurrences is currently limited. Dequalinium is not available
worldwide, including in the US.
Tinidazole
Tinidazole is a second-generation nitroimidazole that may be used in place of metronida-
zole or clindamycin if they are not accessible or tolerated.133 It has an extended half-life (12
to 14 hours).
Dosing: We recommend taking 1 g orally once a day for five days. The effectiveness is somewhat
higher, and the adverse effects are slightly lower than with tinidazole 2 g orally daily for two days.134
Adverse effects: The most often reported symptoms included a metallic taste, nausea,
andfatigue and are comparable to oral metronidazole.135
Efficacy: Similar to metronidazole.
76
Secnidazole
Secnidazole is a nitroimidazole antibiotic with a longer half-life (17-19h) than metronidazole
that is used as a substitute for metronidazole in BV treatment.
Dosing: Secnidazole is administered orally in a single 2 g packet of granules, which can be
dissolved in a serving of pudding, applesauce, or yogurt.136, 137
Adverse effects: Secnidazole treatment is associated with an increased risk of candidiasis,
nausea, diarrhea, and abdominal pain.138
Efficacy: Although single-dose secnidazole was superior to placebo and similar to metroni-
dazole, there is no indication that it is superior to multidose metronidazole treatment.138, 139
Focused assessments of experimental/investigational treatments
Triple-sulfa creams, tetracycline, erythromycin, azithromycin, ampicillin, amoxicillin, 5%
monolaurin vaginal gel, vaginal boric acid, lactic acid or acetic acid gel are not recommend-
ed. They are significantly less effective than metronidazole or clindamycin.81, 140-148
Although there is limited research on the use of lactobacilli,estriol, and sucrose gel in addition to
antibiotics, there is not adequate evidence to include these approaches in treatment guidelines.149
A recent meta-analysis of three randomized controlled trials indicated that the novel drug
Astodrimer 1% gel is superior to placebo and safe for BV treatment.150 Future studies should
compare it to antibiotic treatment.
Efficacy of probiotics for bacterial vaginosis treatment
Probiotics as a supplement to medication may be beneficial in the short term for treating
recurrent vaginal infections in women. In some studies, probiotics decreased the recurrence
rate of BV and the frequency of adverse effects and increased the cure rate of BV compared
to antibiotics.151 However, there is inadequate evidence that probiotics alone effectively
treat acute symptomatic BV.152, 153
In pregnant women, oral probiotic preparations do not prevent BV.154 Vaginal probiotics, in-
cluding lactobacilli, offer the potential to treat and prevent BV.155 However, vaginal probiotic
capsules do not increase BV cure rates nor reduce recurrence.156
Follow-up
If symptoms resolve, follow-up is not indicated following sporadic infections.
Treatment regimens during pregnancy and lactation
Metronidazole 500 mg orally twice daily for seven days, metronidazole 250 mg orally three
times daily for seven days, or clindamycin 300 mg orally twice daily for seven days are all
efficacious and have been related with no significant fetal or obstetric complications.157-159
77
Topical regimens are not inferior to oral medication at treating or preventing adverse BV
outcomes. Topical therapy includes metronidazole 0.75% gel intravaginally once daily for
five days or clindamycin cream 2% intravaginally for seven nights.
For breastfeeding women, we recommend primarily oral metronidazole 500 mg twice a day
for seven days or metronidazole 0.75% gel 5 g once daily intravaginally for five days. Clin-
damycin may have an adverse effect on the gastrointestinal microbiota of breastfed infants.
In animal models the use of high doses of dequalinium chloride was not detected in the blood
stream. Studies using other quaternary ammonium compounds showed no embryofetal toxicity.
Thus, it is assumed to be safe both during pregnancy and breastfeeding, but data are limited.132
Special considerations:
Although some authors reported teratogenicity concerns in the past regarding metroni-
dazole use during the first trimester, a meta-analysis concluded that there is no relation-
ship between metronidazole exposure during the first trimester and congenital malfor-
mations.160
The rate of vaginal lactobacilli colonization was low in pregnant women with normal vag-
inal microbiota at risk for preterm delivery following two months of oral L. reuteri RC-14
and L. rhamnosus GR-1 supplementation.161
Counseling-supportive management in infertile women attending
fertility treatment
In individuals with tubal infertility, the prevalence of BV is considerably increased, and BV
has been related to early spontaneous abortion. Nevertheless, the data was of extremely
poor quality, and the inconclusive results suggest the need for more studies.162
Currently, a recommendation for screening of BV prior to fertility treatment cannot be made.163, 164
Screening and treatment of asymptomatic bacterial vaginosis in pregnancy
To avoid PTB and its associated complications, we do not recommend routine screening or
treatment of pregnant women with asymptomatic BV. Although early diagnosis and treat-
ment of asymptomatic pregnant women with a history of preterm delivery may have advan-
tages, there is insufficient evidence to advocate this as a standard.88, 89, 157, 165-168
Approache to preoperative screening strategies for bacterial vaginosis
We recommend antibiotic therapy before transvaginal surgery for womenwith verified BV. The
treatment alternatives are identical to those available to symptomatic non pregnant women.169-174
Management of sexual partners
A meta-analysis demonstrated that antibiotic treatment of male sexual partners did not in-
crease the rate of clinical or symptomatic recovery, nor did it decrease the recurrence rate
78
throughout a four-week research period.175 However, sexual activity with an untreated reg-
ular sexual partner following BV treatment was related to the development of a suboptimal
vaginal microbiome.176 In a large recent, randomized control study treatment of male sexual
partners of women with recurrent BV, the use of oral metronidazole failed to reduce the
recurrence rate.177 Recently, Plummer et al. showed that the concomitant treatment of male
partners of women with recurrent BV, using oral metronidazole and 2% clindamycin cream
applied topically to penile skin, both twice daily for seven days, was not only well tolerated
but also lead to higher than expected rates of cure.178
Despite the scarcity of data, treatment of female partners of women with BV may be con-
sidered, as there is high agreement rate concerning the BV status, even if asymptomatic.179
Management of recurrent and refractory bacterial vaginosis
Recurrent BV is defined as a confirmed diagnosis of BV occurring three or more times within
one year.180
Within 12 months, following a successful treatment of BV, more than half of women expe-
rience a recurrence.181 There is a lack of guidance for the optimal treatment of women with
recurrent BV.182 We recommend preventing symptomatic relapses with metronidazole 0.75%
gel twice a week for 4–6 months, immediately following successful induction therapy. It is one
of the most widely utilized regimens, with one study demonstrating a 70% success rate while
on this maintenance prophylactic regimen. However, recurrence may still occur when medi-
cation is discontinued, and candidiasis is common throughout this treatment regimen.183
In women living with HIV 1, monthly treatment consisting of 2 g of oral metronidazole and
150 mg of oral fluconazole was useful in reducing recurrence, and the risk of candidiasis
was also lower than in the placebo group.184 In another study, in HIV negative women, with
recurrent BV, a triple phase regimen consisting of oral induction metronidazole or tinida-
zole, followed by 30 consecutive days of vaginal boric acid, and then twice weekly vaginal
metronidazole has been demonstrated to have a therapeutic efficacy of 65% at 28 weeks,
but a 50% failure rate following cessation of drugs when followed at 36 weeks.185 This triple
phase therapy was further improved upon by using vaginal boric acid simultaneously with
oral nitroimidazole and then followed by twice weekly vaginal metronidazole gel.186
BV recurrence and therapeutic failures may be linked to the failure to eradicate the biofilm.
Vaginal boric acid is one of the drugs able to eradicate the biofilm; others include tobramy-
cin, octenidine, and retrocyclin.187 When prescribing boric acid women need to be warned
of toxicity if ingested; ideally, the boric acid should be administered from a compounded
formulation, not bought over the counter. Following effective treatment of BV, using 250 mg
ascorbic acid vaginal tablets six times per month for six months in one study decreased the
probability of recurrence from 32.4% to 16.2% (p=0.024).188 More data however are needed.
Mechanical removal of the biofilm by acidic, antiseptic or disinfectant vaginal washes has
been proposed, but has not been adequately studied.186, 189
The utilization of L. crispatus CTV-05 (Lactin-V) following vaginal metronidazole treatment
resulted in a substantially reduced BV recurrence compared to placebo at 12 weeks but,
although encouraging, is not commercially available.190
79
The notion of utilizing probiotics to restore vaginal health is intuitive and tempting, but it
has not been confirmed by adequate evidence. It does not affect the cure rate, although it
may prolong the period between recurrences in up to 50%.191, 192 Failure to restore the vaginal
microbiota with probiotics may be related to the use of insufficient species andthe failure
of exogenous lactobacilli to colonize the vagina. Oral consumption of yogurt preparation
was also suggested but has not been studied adequately.193 Combined oral contraceptive
pill alone does not lower the likelihood of BV recurrence; vaginal contraceptive ring use may
promote a favorable microbiome after successful treatment.194 Other risk factors worth ad-
dressing include cessation of smoking, use of condoms, and removal of intrauterine devices.
Refractory BV is a less common problem and is more likely due do antibiotic microbial resistance.
It is more common in non-compliant women and with the use of single- dose therapy. Currently,
there is a lack of guidelines for the management of refractory BV. Possible strategies may include
ensuring compliance, changing the original route of treatment and always using a multidose
scheme. In case of failure, changing the drug class should be tried. If the response is still inad-
equate, increasing dose (usually of vaginal formulations, as with the oral route there may be
tolerance and safety issues) or combination therapy including vaginal boric acid can be tried.182
TABLE 3.3 BV treatment algorithm for first-line, second-line, and alternative medications in
the current clinical practice
First-line
Metronidazole tablets 500 mg oral twice daily for 7 days
Metronidazole 0.75% gel 5 g intravaginally once daily for 5 days
Clindamycin cream 2% 5 g intravaginally once daily for 7 days
Second-line
Tinidazole 1 g oral once daily for 5 days
Tinidazole 2 g oral once daily for 2 days
Clindamycin 300 mg oral 2 once daily for 7 days
Clindamycin 100 mg vaginal suppositories once daily for 3 days
Secnidazole 2 g oral, single dose (dissolved in a serving of
pudding, applesauce, or yogurt)
Alternatives
Dequalinium chloride 10 mg tablets intravaginally once daily for 6 days
Clindamycin phosphate 2% cream Single vaginal dose
Metronidazole 1.3% gel Single vaginal dose
Recurrent BV
Metronidazole 0.75% gel 2 times/week for 4–6 months
Triple phase regimen: oral nitroimid-
azole, vaginal boric acid, and vaginal
metronidazole
Oral nitroimidazole once daily for 7 days
Vaginal boric acid once daily for 3 weeks
Vaginal metronidazole gel twice a week for 16 weeks
Metronidazole 2 g + fluconazole 150 mg Once a month
BV during
pregnancy and
lactation
Metronidazole tablets 500 mg oral twice daily for 7 days
250 mg oral 3 times daily for 7 days
Clindamycin capsules 300 mg oral twice daily for 7 days
Metronidazole 0.75% gel 5 g intravaginally once daily 5 days
Clindamycin cream 2% 5 g intravaginally once daily 7 days
80
3.9
Special situations
Infancy
An extensive review of the literature did not identify any specific information pertaining
to BV in infancy. This is not unexpected as BV is a disruption of the normal post-menarchal
VMB, which has not yet been established in infancy. As such, BV is not a condition usually
associated with infancy.
Postmenopausal women
An extensive review of the literature identified a paucity of reliable data pertaining to the in-
cidence of BV in postmenopausal women. This is not unexpected as BV is a disruption of the
normal VMB, which is often not present in postmenopausal women, especially those who
are not receiving estrogen therapy. The absence of local vaginal estrogen, especially in those
women who are not newly postmenopausal, will create conditions such as increased pH,
loss of lactobacilli, and vaginal dryness which will result in a dysbiotic picture in both wet
mount and Nugent’s score. In addition, it may take many years after onset of amenorrhea
before developing changes in vaginal microbiota, regardless of age of onset of menopause
or its duration. Also important are changes in well recognized lifestyle factors which influ-
ence the incidence of BV such as sexual activity, frequency of coitus, number of partners,
and absence of contraception.
In postmenopausal women with a normal microbiome with or without use of estrogen sup-
plementation, BV would be anticipated to occur at a similar rate as premenopausal wom-
en. However, in a systemic review and meta-analysis of BV in postmenopausal women by
Stewart L et al., prevalence estimates ranged from 2.0 to 57.1% and the overall summary
prevalence estimate was 16.93% (95% CI 8.45–27.4%; I2 = 97.9%; p< 0.01) but with marked
heterogeneity. This was based on a total of 328 full-text articles assessed for eligibility, with
only 13 studies found to be eligible for inclusion in the review. In addition, only three studies
focused on postmenopausal women, with all other studies including adult women of all
ages and none of the studies reported any sample size calculations. Also, only one popu-
lation-based study was identified, and the pooled estimate had marked heterogeneity. All
these limitations reinforce the scarcity of reliable BV prevalence data in the postmenopausal
population.65
Finally, we must consider the tools used to diagnose BV as well. Nugent scores and Amsel’s
criteria were developed to diagnose BV in premenopausal women and whether or not these
tests are reliable for the diagnosis of BV in postmenopausal women has not been validated.
As such, it is uncertain if studies using these tests to diagnose BV provide accurate rates in
the postmenopausal population. Molecular studies in postmenopausal women which can
assess quantity and identity of suspected pathogens have not been conducted.195
81
Immunosuppression
Immunosuppression is a widely used term which is nonspecific and not frequently quanti-
fied, thus standardizing study criteria to assess its impact on disease states is challenging.
Other than the complex relationship between BV and HIV, there are limited data discussing
BV in the immunosuppressed population, however, BV is not a common problem in immu-
nocompromised patients in general. A small study by Demirbilek M et al., using the Nugent
score, diagnosed BV in 42% of kidney transplant recipients compared with 9% of healthy
women.196 Murphy et al. reviewed the relationship of host immunity, environment and the
risk of BV and concluded that individuals with genetic variations which lowers their mucosal
innate immune response are at a higher risk of developing BV.84 It has long been known that
women infected with HIV and in whom the disease is not well controlled have an increased
risk of BV due to alterations in mucosal immunity.72 In a review of the vaginal microbiome’s
relationship to various urogenital disorders de Seta et al. noted that multiple cross sectional
studies have shown that independent of behavioral variables, HIV is often correlated with
the presence of BV. They proposed that hydrogen peroxide producing lactobacilli were pro-
tective against HIV acquisition due to reduced recruitment of CD4+ cells to the vaginal mu-
cosa. In addition, increased HIV-1 replication has been noted in a dysbiotic VMB due to the
presence of HIV-inducing factor (HIF) in vaginal secretions.5 Onderdonk AB et al. reviewed
the human microbiome during BV and noted an increased risk of HIV acquisition in women
with BV due to lower levels of antiviral factors such as secretory leukocyte protease inhibitor
(SLPI). They also reported that women with BV, compared to controls, had cervicovaginal
secretions which were lower in innate anti-HIV activity. 8
Despite these associations between BV and HIV, the presence of BV or even recurrent BV is
not considered an indication to screen for HIV.
Bacterial vaginosis in pregnancy
A dysbiotic vaginal microbiota has been linked to poor pregnancy outcomes, including PTB,
PROM, fetal growth restriction/low birth weight, abortion, stillbirth, as well as to neonatal
and puerperal infection.4, 86 Nevertheless, a dysbiotic microbiota cannot be assumed as syn-
onymous of BV.
The dominance by lactobacilli and consequent low pH out of pregancy is a unique feature of
women, that is not shared with other mammals, including other primates. In these species,
dominance by lactobacilli is seen merely during pregnancy, which led to the theory that this
feature is needed for the success of pregnancy (leaving unanswered the question of why most
women during their fertile years have dominance by lactobacilli, even out of pregnancy).197
Pregnancies with good outcomes tend to be dominated from an early phase by lactobacilli,
have a stable microbiota, and low diversity during the whole pregnancy.198-200 This profile is
most probably a consequence of the marked increase in circulating estrogens during this
phase. The shift from a less to a more favorable microbiota is more evident in women of
African descendent, who when non-pregnant more often have non-lactobacilli dominated
vaginal microbiota.200 The success of pregnancy is associated with lactobacilli dominance,
and not necessarily with a specific species within the genus.201
82
Nevertheless, the evidence of a relationship between PTB and BV has not been proved, with
meta-analysis in different populations not showing a clear association.86, 87
The same is true for low birth weight, infants which in a 2020 meta-analysis of studies per-
formed in Sub-Saharan women, was reported in two out of six studies; PROM was reported
in two out of four studies and none showed an association with pregnancy loss.86
In 2020, based on the available evidence (review of 44 studies), the US Preventive Services
Task Force issued a recommendation on the screening of BV in pregnant adolescents and
women. They concluded that, in a general population, there was no benefit of screening
and treating asymptomatic BV; however, in women with a previous PTB, the results were
inconsistent, with three studies showing benefit, while two failed to do so. The question of
whether to screen for BV in asymptomatic women remains unanswered.87
Part of the inconsistencies may be due to the diagnostic criteria and tests used for the diag-
nosis of BV (i.e. increased pH used as a surrogate of BV), the gestational age at the time of
diagnosis, and the outcome evaluated (i.e. early or late PTB).
3.10
Future perspectives
BV is a field in which much work is still needed, including a better understanding of its eti-
ology and complications. For instance, BV is a condition clearly sex-related but not defined
as a STI. One of the interesting theories worth exploring in the future is that a phage may be
the cause of BV and would explain its “transmission”.202
There is a clear idea that, while common, BV is not an ideal or optimal type of microbiota.
Most women with BV are asymptomatic, but it may constitute a disadvantage anyway. A
better understanding of the relation between BV and STIs (including HPV infection and con-
sequent cervical dysplasia), infertility, and obstetrical complications is essential to establish
recommendations on eventual screening and treatment in asymptomatic populations.163 An
association may not necessarily be a cause-effect relationship. Also, even if indeed it is a
cause-effect type of relationship, the direction of such may not always be obvious: for in-
stance, it still is not clear if dysbiosis is a risk factor for HPV infection or if HPV infection leads
to changes in the cervical and vaginal microbiota.203
As with other “vaginitis” in general, the assumption that an empirical diagnosis is easy and
that no exams are needed must be challenged and changed.204 This approach must be
moved to the standard use of wet mount microscopy in office or, in the lack of expertise,
the use the Amsel criteria or of Gram stain and Nugent score (despite the delay in obtaining
the diagnosis). While point-of-care tests seem a reasonable intermediate option, molecular
tests likely will be a major part of the future of the diagnosis of vaginitis. These tests have
good performance and are already commercially available, despite lack of general access
and having a significant associated cost. These open new perspectives, including “profiling”
BV (i.e. the risk associated with BV may be different according to the specific bacteria – or
even clades - present) and in the future antibiotic resistance testing is very likely to be pos-
83
sible.205, 206 Molecular tests may become the gold standard for the diagnosis of BV. However,
before that can be assumed, agreement on a bacterial profile (or profiles) of “molecular BV”
will have to be established.207
The available treatments are very effective in the treatment of acute episodes, but recurrence
is common. New therapeutic approaches, probably targeting the biofilm, are needed to im-
prove the rates of recurrence. While probiotics use seems, from a theoretical point of view, a
logical approach, the available results are not promising (for further details see chapter 10).208
A very promising area in the field of treatment is the vaginal microbiome transplant. The
concept of transplanting vaginal microbiome from healthy women to women with intracta-
ble BV is still investigational, but the results are encouraging.209, 210
The establishment of core outcome sets to be evaluated in studies is also a topic that de-
serves attention in the near future. This will allow direct comparison between studies in the
short term and in the medium and long term, the performance of meta-analysis with lower
heterogeneity.
Recommendations
Recommendation Quality of
evidence
Strength of
recommendation
Screening and treatment of bacterial vaginosis to prevent preterm
birth is currently not recommended. 1a A
The Amsel criteria may be useful in clinical practice, in the ab-
sence of expertise or availability of a microscope or other tests. 1b A
The Amsel criteria are not suitable for the screening of bacterial
vaginosis. 2b C
The Nugent score is the gold standard for the diagnosis of bacte-
rial vaginosis. 2a B
The Ison-Hay criteria can be used as an alternative to the Nugent
score. 4 C
Wet mount microscopy is a good tool for in office diagnosis of
bacterial vaginosis. 2b B
Cultures should not be used for the diagnosis of bacterial vaginosis. 4 D
Point of care tests, such as the OSOM BV Blue test and the FexEx-
am card, can be used for the diagnosis of bacterial vaginosis. 3b C
Direct DNA probe assays (Affirm VP) can be used for the diagnosis
of bacterial vaginosis (as well as of candidiasis and trichomononi-
sis).
2b B
Nucleic acid amplification tests (Allplex vaginitis, BD Max vaginal
panel, NuSwab, MDL BV panel) are recommended for the diagno-
sis of bacterial vaginosis.
2b B
There is no evidence to recommend the treatment of bacterial
vaginosis in asymptomatic non-pregnant women. 2b B
84
There is no evidence to recommend screening and treatment of
bacterial vaginosis prior to fertility treatments. 2b B
Screening and treatment of bacterial vaginosis can be considered
in women at high risk for sexually transmitted infections. 4 C
Treatment of asymptomatic bacterial vaginosis is recommended
prior to transvaginal surgery. 2b B
Topical or oral metronidazole or clindamycin are considered first
line treatments for bacterial vaginosis. 1b A
Tinidazole or secnidazole are acceptable oral alternatives. 2a B
Vaginal dequalinium chloride can be considered as as option for
the treatment of bacterial vaginosis. 2b B
Women using vaginal clindamycin must be warned that it weak-
ens condoms for up to 5 days after finishing the treatment. 5 D
Astodrimer 1% vaginal gel may be useful in the treatment of
vaginal bacteriosis. 3a B
Probiotics alone are not recommended as a treatment for bacte-
rial vaginosis. 1a A
Probiotics may decrease the rate of recurrence of bacterial vagino-
sis. 2a B
The first line treatment options used in non-pregnant women can
be used during pregnancy. 2a B
In breastfeeding women, metronidazole may be preferable to
clindamycin. 5 D
Treatment of partners is currently not recommended. 2a B
References
1. Moosa, Y.; Kwon, D.; de Oliveira, T.; Wong, E. B., Determinants of Vaginal Microbiota Composition. Front Cell Infect
Microbiol 2020, 10, 467.
2. Verstraelen, H.; Vieira-Baptista, P.; De Seta, F.; Ventolini, G.; Lonnee-Hoffmann, R.; Lev-Sagie, A., The Vaginal Micro-
biome: I. Research Development, Lexicon, Defining “Normal” and the Dynamics Throughout Women’s Lives. J Low
Genit Tract Dis 2022, 26, (1), 73-78.
3. Anahtar, M. N.; Gootenberg, D. B.; Mitchell, C. M.; Kwon, D. S., Cervicovaginal Microbiota and Reproductive Health:
The Virtue of Simplicity. Cell Host Microbe 2018, 23, (2), 159-168.
4. Ventolini, G.; Vieira-Baptista, P.; De Seta, F.; Verstraelen, H.; Lonnee-Hoffmann, R.; Lev-Sagie, A., The Vaginal Microbiome:
IV. The Role of Vaginal Microbiome in Reproduction and in Gynecologic Cancers. J Low Genit Tract Dis 2022, 26, (1), 93-98.
5. De Seta, F.; Lonnee-Hoffmann, R.; Campisciano, G.; Comar, M.; Verstraelen, H.; Vieira-Baptista, P.; Ventolini, G.;
Lev-Sagie, A., The Vaginal Microbiome: III. The Vaginal Microbiome in Various Urogenital Disorders. J Low Genit Tract
Dis 2022, 26, (1), 85-92.
6. Lev-Sagie, A.; De Seta, F.; Verstraelen, H.; Ventolini, G.; Lonnee-Hoffmann, R.; Vieira-Baptista, P., The Vaginal Microbi-
ome: II. Vaginal Dysbiotic Conditions. J Low Genit Tract Dis 2022, 26, (1), 79-84.
7. Zozaya-Hinchliffe, M.; Lillis, R.; Martin, D. H.; Ferris, M. J., Quantitative PCR assessments of bacterial species in women
with and without bacterial vaginosis. J Clin Microbiol 2010, 48, (5), 1812-9.
8. Onderdonk, A. B.; Delaney, M. L.; Fichorova, R. N., The Human Microbiome during Bacterial Vaginosis. Clin Microbiol
Rev 2016, 29, (2), 223-38.
9. Boskey, E. R.; Telsch, K. M.; Whaley, K. J.; Moench, T. R.; Cone, R. A., Acid production by vaginal flora in vitro is consist-
ent with the rate and extent of vaginal acidification. Infect. Immun. 1999, 67, (10), 5170-5.
85
10. Boskey, E. R.; Cone, R. A.; Whaley, K. J.; Moench, T. R., Origins of vaginal acidity: high D/L lactate ratio is consistent
with bacteria being the primary source. Hum. Reprod. 2001, 16, (9), 1809-13.
11. Fredricks, D. N.; Fiedler, T. L.; Marrazzo, J. M., Molecular identification of bacteria associated with bacterial vaginosis.
N. Engl. J. Med. 2005, 353, (18), 1899-911.
12. Ravel, J.; Gajer, P.; Abdo, Z.; Schneider, G. M.; Koenig, S. S.; McCulle, S. L.; Karlebach, S.; Gorle, R.; Russell, J.; Tacket, C.
O.; Brotman, R. M.; Davis, C. C.; Ault, K.; Peralta, L.; Forney, L. J., Vaginal microbiome of reproductive-age women. Proc.
Natl. Acad. Sci. U. S. A. 2011, 108 Suppl 1, 4680-7.
13. McKinnon, L. R.; Achilles, S. L.; Bradshaw, C. S.; Burgener, A.; Crucitti, T.; Fredricks, D. N.; Jaspan, H. B.; Kaul, R.; Kaushic,
C.; Klatt, N.; Kwon, D. S.; Marrazzo, J. M.; Masson, L.; McClelland, R. S.; Ravel, J.; van de Wijgert, J.; Vodstrcil, L. A.;
Tachedjian, G., The Evolving Facets of Bacterial Vaginosis: Implications for HIV Transmission. AIDS Res Hum Retrovi-
ruses 2019, 35, (3), 219-228.
14. Nugent, R. P.; Krohn, M. A.; Hillier, S. L., Reliability of diagnosing bacterial vaginosis is improved by a standardized
method of gram stain interpretation. J Clin Microbiol 1991, 29, (2), 297-301.
15. Hillier, S. L., Diagnostic microbiology of bacterial vaginosis. Am. J. Obstet. Gynecol. 1993, 169, (2 Pt 2), 455-9.
16. Srinivasan, S.; Hoffman, N. G.; Morgan, M. T.; Matsen, F. A.; Fiedler, T. L.; Hall, R. W.; Ross, F. J.; McCoy, C. O.; Bumgarner, R.;
Marrazzo, J. M.; Fredricks, D. N., Bacterial communities in women with bacterial vaginosis: high resolution phylogenetic
analyses reveal relationships of microbiota to clinical criteria. PLoS One 2012, 7, (6), e37818.
17. Gajer, P.; Brotman, R. M.; Bai, G.; Sakamoto, J.; Schütte, U. M.; Zhong, X.; Koenig, S. S.; Fu, L.; Ma, Z. S.; Zhou, X.; Abdo,
Z.; Forney, L. J.; Ravel, J., Temporal dynamics of the human vaginal microbiota. Sci Transl Med 2012, 4, (132), 132ra52.
18. Muzny, C. A.; Taylor, C. M.; Swords, W. E.; Tamhane, A.; Chattopadhyay, D.; Cerca, N.; Schwebke, J. R., An updated
conceptual model on the pathogenesis of bacterial vaginosis. J. Infect. Dis. 2019, 220, (9), 1399-1405.
19. Muzny, C. A.; Schwebke, J. R., Gardnerella vaginalis: Still a Prime Suspect in the Pathogenesis of Bacterial Vaginosis.
Curr. Infect. Dis. Rep. 2013, 15, (2), 130-5.
20. Schwebke, J. R.; Muzny, C. A.; Josey, W. E., Role of Gardnerella vaginalis in the pathogenesis of bacterial vaginosis: A
conceptual model. J. Infect. Dis. 2014, 210, (3), 338-343.
21. Coleman, J. S.; Gaydos, C. A., Molecular Diagnosis of Bacterial Vaginosis: an Update. J Clin Microbiol 2018, 56, (9).
22. Vaneechoutte, M.; Guschin, A.; Van Simaey, L.; Gansemans, Y.; Van Nieuwerburgh, F.; Cools, P., Emended description
of Gardnerella vaginalis and description of Gardnerella leopoldii sp. nov., Gardnerella piotii sp. nov. and Gardnerella
swidsinskii sp. nov., with delineation of 13 genomic species within the genus Gardnerella. Int. J. Syst. Evol. Microbiol.
2019, 69, (3), 679-687.
23. Konschuh, S.; Jayaprakash, T.; Dolatabadi, A.; Dayo, E.; Ramay, H.; Sycuro, L., O02.3 Reclassification of Atopobium
vaginae as three novel Fannyhessea species: implications for understanding their role in bacterial vaginosis. Sexually
Transmitted Infections 2021, 97, (Suppl 1), A18-A18.
24. Swidsinski, A.; Mendling, W.; Loening-Baucke, V.; Ladhoff, A.; Swidsinski, S.; Hale, L. P.; Lochs, H., Adherent biofilms in
bacterial vaginosis. Obstet. Gynecol. 2005, 106, (5 Pt 1), 1013-23.
25. Patterson, J. L.; Stull-Lane, A.; Girerd, P. H.; Jefferson, K. K., Analysis of adherence, biofilm formation and cytotoxicity
suggests a greater virulence potential of Gardnerella vaginalis relative to other bacterial-vaginosis-associated an-
aerobes. Microbiology 2010, 156, (Pt 2), 392-9.
26. Alves, P.; Castro, J.; Sousa, C.; Cereija, T. B.; Cerca, N., Gardnerella vaginalis outcompetes 29 other bacterial species isolat-
ed from patients with bacterial vaginosis, using in an in vitro biofilm formation model. J. Infect. Dis. 2014, 210, (4), 593-6.
27. Udayalaxmi, J.; Bhat, G.; Kotigadde, S.; Kotian, S., Effect of pH on the adherence, surface hydrophobicity and the biofilm
formation of Gardnerella Vaginalis. Journal of Clinical and Diagnostic Research 2012, 6, (6), 967-969.
28. Pybus, V.; Onderdonk, A. B., Evidence for a commensal, symbiotic relationship between Gardnerella vaginalis and
Prevotella bivia involving ammonia: potential significance for bacterial vaginosis. J. Infect. Dis. 1997, 175, (2), 406-13.
29. Hardy, L.; Jespers, V.; Dahchour, N.; Mwambarangwe, L.; Musengamana, V.; Vaneechoutte, M.; Crucitti, T., Unravelling
the bacterial vaginosis-associated biofilm: a multiplex Gardnerella vaginalis and Atopobium vaginae fluorescence
in situ hybridization assay using peptide nucleic acid probes. PLoS One 2015, 10, (8), e0136658.
30. Briselden, A. M.; Moncla, B. J.; Stevens, C. E.; Hillier, S. L., Sialidases (neuraminidases) in bacterial vaginosis and bac-
terial vaginosis-associated microflora. J. Clin. Microbiol. 1992, 30, (3), 663-6.
31. Hardy, L.; Jespers, V.; Van den Bulck, M.; Buyze, J.; Mwambarangwe, L.; Musengamana, V.; Vaneechoutte, M.; Crucitti,
T., The presence of the putative Gardnerella vaginalis sialidase A gene in vaginal specimens is associated with bac-
terial vaginosis biofilm. PLoS One 2017, 12, (2), e0172522.
32. Muzny, C. A.; Laniewski, P.; Schwebke, J. R.; Herbst-Kralovetz, M. M., Host-vaginal microbiota interactions in the
pathogenesis of bacterial vaginosis. Curr. Opin. Infect. Dis. 2019.
86
33. Amsel, R.; Totten, P. A.; Spiegel, C. A.; Chen, K. C.; Eschenbach, D.; Holmes, K. K., Nonspecific vaginitis. Diagnostic
criteria and microbial and epidemiologic associations. Am J Med 1983, 74, (1), 14-22.
34. Brand, J. M.; Galask, R. P., Trimethylamine: the substance mainly responsible for the fishy odor often associated with
bacterial vaginosis. Obstet. Gynecol. 1986, 68, (5), 682-5.
35. Chen, K. C.; Forsyth, P. S.; Buchanan, T. M.; Holmes, K. K., Amine content of vaginal fluid from untreated and treated
patients with nonspecific vaginitis. J. Clin. Invest. 1979, 63, (5), 828-35.
36. Srinivasan, S.; Morgan, M. T.; Fiedler, T. L.; Djukovic, D.; Hoffman, N. G.; Raftery, D.; Marrazzo, J. M.; Fredricks, D. N.,
Metabolic signatures of bacterial vaginosis. mBio 2015, 6, (2).
37. Muzny, C. A.; Schwebke, J. R., Pathogenesis of Bacterial Vaginosis: Discussion of Current Hypotheses. J. Infect. Dis.
2016, 214 Suppl 1, S1-5.
38. Mehta, S. D.; Zhao, D.; Green, S. J.; Agingu, W.; Otieno, F.; Bhaumik, R.; Bhaumik, D.; Bailey, R. C., The microbiome
composition of a man’s penis predicts incident bacterial vaginosis in his female sex partner with high accuracy.
Front Cell Infect Microbiol 2020, 10, 433.
39. Vodstrcil, L. A.; Walker, S. M.; Hocking, J. S.; Law, M.; Forcey, D. S.; Fehler, G.; Bilardi, J. E.; Chen, M. Y.; Fethers, K. A.;
Fairley, C. K.; Bradshaw, C. S., Incident bacterial vaginosis (BV) in women who have sex with women is associated
with behaviors that suggest sexual transmission of BV. Clin. Infect. Dis. 2015, 60, (7), 1042-53.
40. Zozaya, M.; Ferris, M. J.; Siren, J. D.; Lillis, R.; Myers, L.; Nsuami, M. J.; Eren, A. M.; Brown, J.; Taylor, C. M.; Martin, D. H.,
Bacterial communities in penile skin, male urethra, and vaginas of heterosexual couples with and without bacterial
vaginosis. Microbiome 2016, 4, 16.
41. Muzny, C. A.; Lensing, S. Y.; Aaron, K. J.; Schwebke, J. R., Incubation period and risk factors support sexual transmis-
sion of bacterial vaginosis in women who have sex with women. Sex. Transm. Infect. 2019, 0, (0), 1-5.
42. Bradshaw, C. S.; Vodstrcil, L. A.; Hocking, J. S.; Law, M.; Pirotta, M.; Garland, S. M.; De Guingand, D.; Morton, A. N.;
Fairley, C. K., Recurrence of bacterial vaginosis is significantly associated with posttreatment sexual activities and
hormonal contraceptive use. Clin. Infect. Dis. 2013, 56, (6), 777-86.
43. Bradshaw, C. S.; Walker, J.; Fairley, C. K.; Chen, M. Y.; Tabrizi, S. N.; Donovan, B.; Kaldor, J. M.; McNamee, K.; Urban, E.; Walker, S.;
Currie, M.; Birden, H.; Bowden, F.; Gar land, S.; Pirotta, M.; Gurrin, L.; Hocking, J. S., Prevalent and incident bacterial vaginosis
are associated with sexual and contraceptive behaviours in young Australian women. PLoS One 2013, 8, (3), e57688.
44. Muzny, C. A.; Schwebke, J. R., Suspected heterosexual transmission of bacterial vaginosis without seminal fluid
exposure. Sex. Transm. Dis. 2014, 41, (1), 58-60.
45. Marrazzo, J. M.; Thomas, K. K.; Fiedler, T. L.; Ringwood, K.; Fredricks, D. N., Risks for acquisition of bacterial vaginosis
among women who report sex with women: a cohort study. PLoS One 2010, 5, (6), e11139.
46. Fethers, K. A.; Fairley, C. K.; Hocking, J. S.; Gurrin, L. C.; Bradshaw, C. S., Sexual risk factors and bacterial vaginosis: a
systematic review and meta-analysis. Clin Infect Dis 2008, 47, (11), 1426-35.
47. Larsson, P. G.; Platz-Christensen, J. J.; Sundstrom, E., Is bacterial vaginosis a sexually transmitted disease? Int. J. STD
AIDS 1991, 2, (5), 362-4.
48. Fethers, K. A.; Fairley, C. K.; Morton, A.; Hocking, J. S.; Hopkins, C.; Kennedy, L. J.; Fehler, G.; Bradshaw, C. S., Early sexual
experiences and risk factors for bacterial vaginosis. J. Infect. Dis. 2009, 200, (11), 1662-70.
49. Hellberg, D.; Nilsson, S.; Mardh, P. A., Bacterial vaginosis and smoking. Int. J. STD AIDS 2000, 11, (9), 603-6.
50. Bradshaw, C. S.; Walker, S. M.; Vodstrcil, L. A.; Bilardi, J. E.; Law, M.; Hocking, J. S.; Fethers, K. A.; Fehler, G.; Petersen,
S.; Tabrizi, S. N.; Chen, M. Y.; Garland, S. M.; Fairley, C. K., The influence of behaviors and relationships on the vaginal
microbiota of women and their female partners: the WOW Health Study. J Infect Dis 2014, 209, (10), 1562-72.
51. Brotman, R. M.; He, X.; Gajer, P.; Fadrosh, D.; Sharma, E.; Mongodin, E. F.; Ravel, J.; Glover, E. D.; Rath, J. M., Association
between cigarette smoking and the vaginal microbiota: a pilot study. BMC Infect Dis 2014, 14, 471.
52. Mehta, S. D.; Donovan, B.; Weber, K. M.; Cohen, M.; Ravel, J.; Gajer, P.; Gilbert, D.; Burgad, D.; Spear, G. T., The vaginal
microbiota over an 8- to 10-year period in a cohort of HIV-infected and HIV-uninfected women. PLoS One 2015, 10,
(2), e0116894.
53. Ness, R. B.; Hillier, S. L.; Richter, H. E.; Soper, D. E.; Stamm, C.; McGregor, J.; Bass, D. C.; Sweet, R. L.; Rice, P., Douching in
relation to bacterial vaginosis, lactobacilli, and facultative bacteria in the vagina. Obstet Gynecol 2002, 100, (4), 765.
54. Hutchinson, K. B.; Kip, K. E.; Ness, R. B.; Gynecologic Infection Follow-Through, I., Vaginal douching and development of
bacterial vaginosis among women with normal and abnormal vaginal microflora. Sex. Transm. Dis. 2007, 34, (9), 671-5.
55. Brotman, R. M.; Klebanoff, M. A.; Nansel, T. R.; Andrews, W. W.; Schwebke, J. R.; Zhang, J.; Yu, K. F.; Zenilman, J. M.;
Scharfstein, D. O., A longitudinal study of vaginal douching and bacterial vaginosis--a marginal structural modeling
analysis. Am. J. Epidemiol. 2008, 168, (2), 188-96.
87
56. Van der Veer, C.; Bruisten, S. M.; van Houdt, R.; Matser, A. A.; Tachedjian, G.; van de Wijgert, J.; de Vries, H. J. C.; van
der Helm, J. J., Effects of an over-the-counter lactic-acid containing intra-vaginal douching product on the vaginal
microbiota. BMC Microbiol 2019, 19, (1), 168.
57. Peebles, K.; Velloza, J.; Balkus, J. E.; McClelland, R. S.; Barnabas, R. V., High Global Burden and Costs of Bacterial Vagi-
nosis: A Systematic Review and Meta-Analysis. Sex Transm Dis 2019, 46, (5), 304-311.
58. Torrone, E. A.; Morrison, C. S.; Chen, P. L.; Kwok, C.; Francis, S. C.; Hayes, R. J.; Looker, K. J.; McCormack, S.; McGrath, N.;
van de Wijgert, J.; Watson-Jones, D.; Low, N.; Gottlieb, S. L.; Group, S. W., Prevalence of sexually transmitted infections
and bacterial vaginosis among women in sub-Saharan Africa: An individual participant data meta-analysis of 18 HIV
prevention studies. PLoS Med. 2018, 15, (2), e1002511.
59. Chico, R. M.; Mayaud, P.; Ariti, C.; Mabey, D.; Ronsmans, C.; Chandramohan, D., Prevalence of malaria and sexually
transmitted and reproductive tract infections in pregnancy in sub-Saharan Africa: a systematic review. JAMA 2012,
307, (19), 2079-86.
60. Nyemba, D. C.; Haddison, E. C.; Wang, C.; Johnson, L. F.; Myer, L.; Davey, D. J., Prevalence of curable STIs and bacterial
vaginosis during pregnancy in sub-Saharan Africa: a systematic review and meta-analysis. Sex. Transm. Infect. 2021.
61. Peebles, K.; Kiweewa, F. M.; Palanee-Phillips, T.; Chappell, C.; Singh, D.; Bunge, K. E.; Naidoo, L.; Makanani, B.; Jeenarain,
N.; Reynolds, D.; Hillier, S. L.; Brown, E. R.; Baeten, J. M.; Balkus, J. E.; team, M. T. N. A. s., Elevated Risk of Bacterial Vagi-
nosis among Users of the Copper Intrauterine Device: A Prospective Longitudinal Cohort Study. Clin. Infect. Dis. 2020.
62. Sabour, S.; Arzanlou, M.; Vaez, H.; Rahimi, G.; Sahebkar, A.; Khademi, F., Prevalence of bacterial vaginosis in pregnant and
non-pregnant Iranian women: a systematic review and meta-analysis. Arch. Gynecol. Obstet. 2018, 297, (5), 1101-1113.
63. Takemoto, M. L. S.; Menezes, M. O.; Polido, C. B. A.; Santos, D. S.; Leonello, V. M.; Magalhaes, C. G.; Cirelli, J. F.; Knobel,
R., Prevalence of sexually transmitted infections and bacterial vaginosis among lesbian women: systematic review
and recommendations to improve care. Cad. Saude Publica 2019, 35, (3), e00118118.
64. Skafte-Holm, A.; Humaidan, P.; Bernabeu, A.; Lledo, B.; Jensen, J. S.; Haahr, T., The Association between Vaginal Dys-
biosis and Reproductive Outcomes in Sub-Fertile Women Undergoing IVF-Treatment: A Systematic PRISMA Review
and Meta-Analysis. Pathogens 2021, 10, (3).
65. Stewart, L. L.; Vodstrcil, L. A.; Coombe, J.; Bradshaw, C. S.; Hocking, J. S., Prevalence of bacterial vaginosis in postmen-
opausal women: a systematic review and meta-analysis. Sex Health 2022, 19, (1), 17-26.
66. Allsworth, J. E.; Peipert, J. F., Prevalence of bacterial vaginosis: 2001-2004 National Health and Nutrition Examination
Survey data. Obstet Gynecol 2007, 109, (1), 114-20.
67. Fethers, K.; Marks, C.; Mindel, A.; Estcourt, C. S., Sexually transmitted infections and risk behaviours in women who
have sex with women. Sex Transm Infect 2000, 76, (5), 345-9.
68. Marrazzo, J. M.; Antonio, M.; Agnew, K.; Hillier, S. L., Distribution of genital Lactobacillus strains shared by female sex
partners. J Infect Dis 2009, 199, (5), 680-3.
69. Marrazzo, J. M.; Koutsky, L. A.; Eschenbach, D. A.; Agnew, K.; Stine, K.; Hillier, S. L., Characterization of vaginal flora and
bacterial vaginosis in women who have sex with women. J Infect Dis 2002, 185, (9), 1307-13.
70. Evans, A. L.; Scally, A. J.; Wellard, S. J.; Wilson, J. D., Prevalence of bacterial vaginosis in lesbians and heterosexual
women in a community setting. Sex Transm Infect 2007, 83, (6), 470-5.
71. Esber, A.; Vicetti Miguel, R. D.; Cherpes, T. L.; Klebanoff, M. A.; Gallo, M. F.; Turner, A. N., Risk of Bacterial Vaginosis
Among Women With Herpes Simplex Virus Type 2 Infection: A Systematic Review and Meta-analysis. J Infect Dis
2015, 212, (1), 8-17.
72. Jamieson, D. J.; Duerr, A.; Klein, R. S.; Paramsothy, P.; Brown, W.; Cu-Uvin, S.; Rompalo, A.; Sobel, J., Longitudinal anal-
ysis of bacterial vaginosis: findings from the HIV epidemiology research study. Obstet Gynecol 2001, 98, (4), 656-63.
73. Myer, L.; Denny, L.; Telerant, R.; Souza, M.; Wright, T. C., Jr.; Kuhn, L., Bacterial vaginosis and susceptibility to HIV infec-
tion in South African women: a nested case-control study. J Infect Dis 2005, 192, (8), 1372-80.
74. Atashili, J.; Poole, C.; Ndumbe, P. M.; Adimora, A. A.; Smith, J. S., Bacterial vaginosis and HIV acquisition: a meta-anal-
ysis of published studies. Aids 2008, 22, (12), 1493-501.
75. Brotman, R. M.; Klebanoff, M. A.; Nansel, T. R.; Yu, K. F.; Andrews, W. W.; Zhang, J.; Schwebke, J. R., Bacterial vaginosis
assessed by gram stain and diminished colonization resistance to incident gonococcal, chlamydial, and trichomon-
al genital infection. J Infect Dis 2010, 202, (12), 1907-15.
76. Abbai, N. S.; Reddy, T.; Ramjee, G., Prevalent bacterial vaginosis infection - a risk factor for incident sexually transmit-
ted infections in women in Durban, South Africa. Int J STD AIDS 2016, 27, (14), 1283-1288.
77. Lokken, E. M.; Balkus, J. E.; Kiarie, J.; Hughes, J. P.; Jaoko, W.; Totten, P. A.; McClelland, R. S.; Manhart, L. E., Association
of Recent Bacterial Vaginosis With Acquisition of Mycoplasma genitalium. Am J Epidemiol 2017, 186, (2), 194-201.
88
78. Brusselaers, N.; Shrestha, S.; van de Wijgert, J.; Verstraelen, H., Vaginal dysbiosis and the risk of human papillomavirus
and cervical cancer: systematic review and meta-analysis. Am J Obstet Gynecol 2019, 221, (1), 9-18.e8.
79. Schwebke, J. R.; Desmond, R. A.; Oh, M. K., Predictors of bacterial vaginosis in adolescent women who douche. Sex
Transm Dis 2004, 31, (7), 433-6.
80. Ness, R. B.; Kip, K. E.; Soper, D. E.; Stamm, C. A.; Rice, P.; Richter, H. E., Variability of bacterial vaginosis over 6- to
12-month intervals. Sex Transm Dis 2006, 33, (6), 381-5.
81. Schwebke, J. R.; Desmond, R. A., A randomized trial of the duration of therapy with metronidazole plus or minus
azithromycin for treatment of symptomatic bacterial vaginosis. Clin Infect Dis 2007, 44, (2), 213-9.
82. Klebanoff, M. A.; Nansel, T. R.; Brotman, R. M.; Zhang, J.; Yu, K. F.; Schwebke, J. R.; Andrews, W. W., Personal hygienic
behaviors and bacterial vaginosis. Sex Transm Dis 2010, 37, (2), 94-9.
83. Brookheart, R. T.; Lewis, W. G.; Peipert, J. F.; Lewis, A. L.; Allsworth, J. E., Association between obesity and bacterial
vaginosis as assessed by Nugent score. Am J Obstet Gynecol 2019, 220, (5), 476.e1-476.e11.
84. Murphy, K.; Mitchell, C. M., The Interplay of Host Immunity, Environment and the Risk of Bacterial Vaginosis and
Associated Reproductive Health Outcomes. J Infect Dis 2016, 214 Suppl 1, (Suppl 1), S29-35.
85. Peebles, K.; Kiweewa, F. M.; Palanee-Phillips, T.; Chappell, C.; Singh, D.; Bunge, K. E.; Naidoo, L.; Makanani, B.; Jeenarain,
N.; Reynolds, D.; Hillier, S. L.; Brown, E. R.; Baeten, J. M.; Balkus, J. E., Elevated Risk of Bacterial Vaginosis Among Users
of the Copper Intrauterine Device: A Prospective Longitudinal Cohort Study. Clin Infect Dis 2021, 73, (3), 513-520.
86. Juliana, N. C. A.; Suiters, M. J. M.; Al-Nasiry, S.; Morré, S. A.; Peters, R. P. H.; Ambrosino, E., The Association Between Vag-
inal Microbiota Dysbiosis, Bacterial Vaginosis, and Aerobic Vaginitis, and Adverse Pregnancy Outcomes of Women
Living in Sub-Saharan Africa: A Systematic Review. Front Public Health 2020, 8, 567885.
87. Kahwati, L. C.; Clark, R.; Berkman, N.; Urrutia, R.; Patel, S. V.; Zeng, J.; Viswanathan, M., Screening for Bacterial Vaginosis
in Pregnant Adolescents and Women to Prevent Preterm Delivery: Updated Evidence Report and Systematic Re-
view for the US Preventive Services Task Force. JAMA 2020, 323, (13), 1293-1309.
88. Owens, D. K.; Davidson, K. W.; Krist, A. H.; Barr y, M. J.; Cabana, M.; Caughey, A. B.; Donahue, K.; Doubeni, C. A.; Epling, J.
W., Jr.; Kubik, M.; Ogedegbe, G.; Pbert, L.; Silverstein, M.; Simon, M. A.; Tseng, C. W.; Wong, J. B., Screening for Bacterial
Vaginosis in Pregnant Persons to Prevent Preterm Delivery: US Preventive Services Task Force Recommendation
Statement. JAMA 2020, 323, (13), 1286-1292.
89. Thinkhamrop, J.; Hofmeyr, G. J.; Adetoro, O.; Lumbiganon, P.; Ota, E., Antibiotic prophylaxis during the second and third
trimester to reduce adverse pregnancy outcomes and morbidity. Cochrane Database Syst Rev 2015, 2015, (6), Cd002250.
90. Eason, E.; Wells, G.; Garber, G.; Hemmings, R.; Luskey, G.; Gillett, P.; Martin, M., Antisepsis for abdominal hysterectomy:
a randomised controlled trial of povidone-iodine gel. Bjog 2004, 111, (7), 695-9.
91. Russo, J. A.; Achilles, S.; DePineres, T.; Gil, L., Controversies in family planning: postabortal pelvic inflammatory dis-
ease. Contraception 2013, 87, (4), 497-503.
92. Haggerty, C. L.; Hillier, S. L.; Bass, D. C.; Ness, R. B., Bacterial vaginosis and anaerobic bacteria are associated with
endometritis. Clin Infect Dis 2004, 39, (7), 990-5.
93. Wiesenfeld, H. C.; Hillier, S. L.; Krohn, M. A.; Amortegui, A. J.; Heine, R. P.; Landers, D. V.; Sweet, R. L., Lower genital tract
infection and endometritis: insight into subclinical pelvic inflammatory disease. Obstet Gynecol 2002, 100, (3), 456-63.
94. Abbai, N. S.; Nyirenda, M.; Naidoo, S.; Ramjee, G., Prevalent Herpes Simplex Virus-2 Increases the Risk of Incident
Bacterial Vaginosis in Women from South Africa. AIDS Behav 2018, 22, (7), 2172-2180.
95. Schwebke, J. R.; Desmond, R., A randomized trial of metronidazole in asymptomatic bacterial vaginosis to prevent
the acquisition of sexually transmitted diseases. Am J Obstet Gynecol 2007, 196, (6), 517.e1-6.
96. Wiesenfeld, H. C.; Hillier, S. L.; Krohn, M. A.; Landers, D. V.; Sweet, R. L., Bacterial vaginosis is a strong predictor of
Neisseria gonorrhoeae and Chlamydia trachomatis infection. Clin Infect Dis 2003, 36, (5), 663-8.
97. Verstraelen, H.; Verhelst, R., Bacterial vaginosis: an update on diagnosis and treatment. Expert Rev Anti Infect Ther
2009, 7, (9), 1109-24.
98. Olmsted, S. S.; Meyn, L. A.; Rohan, L. C.; Hillier, S. L., Glycosidase and proteinase activity of anaerobic gram-negative
bacteria isolated from women with bacterial vaginosis. Sex Transm Dis 2003, 30, (3), 257-61.
99. Spence, D.; Melville, C., Vaginal discharge. BMJ 2007, 335, (7630), 1147-51.
100. Nelson, T. M.; Borgogna, J. L.; Brotman, R. M.; Ravel, J.; Walk, S. T.; Yeoman, C. J., Vaginal biogenic amines: biomarkers
of bacterial vaginosis or precursors to vaginal dysbiosis? Front Physiol 2015, 6, 253.
101. Anderson, M. R.; Klink, K.; Cohrssen, A., Evaluation of vaginal complaints. JAMA 2004, 291, (11), 1368-79.
102. Benyas, D.; Sobel, J. D., Mixed Vaginitis Due to Bacterial Vaginosis and Candidiasis. J Low Genit Tract Dis 2022, 26, (1), 68-70.
103. Thomason, J. L.; Gelbart, S. M.; Anderson, R. J.; Walt, A. K.; Osypowski, P. J.; Broekhuizen, F. F., Statistical evaluation of
diagnostic criteria for bacterial vaginosis. Am J Obstet Gynecol 1990, 162, (1), 155-60.
89
104. Gutman, R. E.; Peipert, J. F.; Weitzen, S.; Blume, J., Evaluation of clinical methods for diagnosing bacterial vaginosis.
Obstet Gynecol 2005, 105, (3), 551-6.
105. Vieira-Baptista, P.; Silva, A. R.; Costa, M.; Figueiredo, R.; Saldanha, C.; Sousa, C., Diagnosis of bacterial vaginosis: Clinical
or microscopic? A cross-sectional study. Int J Gynaecol Obstet 2022, 156, (3), 552-559.
106. Eschenbach, D. A.; Hillier, S.; Critchlow, C.; Stevens, C.; DeRouen, T.; Holmes, K. K., Diagnosis and clinical manifesta-
tions of bacterial vaginosis. Am J Obstet Gynecol 1988, 158, (4), 819-28.
107. Sha, B. E.; Chen, H. Y.; Wang, Q. J.; Zariffard, M. R.; Cohen, M. H.; Spear, G. T., Utility of Amsel criteria, Nugent score, and
quantitative PCR for Gardnerella vaginalis, Mycoplasma hominis, and Lactobacillus spp. for diagnosis of bacterial
vaginosis in human immunodeficiency virus-infected women. J Clin Microbiol 2005, 43, (9), 4607-12.
108. Vieira-Baptista, P.; Grincevičienė, Š.; Oliveira, C.; Fonseca-Moutinho, J.; Cherey, F.; Stockdale, C. K., The International
Society for the Study of Vulvovaginal Disease Vaginal Wet Mount Microscopy Guidelines: How to Perform, Applica-
tions, and Interpretation. J Low Genit Tract Dis 2021, 25, (2), 172-180.
109. Srinivasan, S.; Morgan, M. T.; Liu, C.; Matsen, F. A.; Hoffman, N. G.; Fiedler, T. L.; Agnew, K. J.; Marrazzo, J. M.; Fredricks,
D. N., More than meets the eye: associations of vaginal bacteria with gram stain morphotypes using molecular
phylogenetic analysis. PLoS One 2013, 8, (10), e78633.
110. Holm, J. B.; France, M. T.; Ma, B.; McComb, E.; Robinson, C. K.; Mehta, A.; Tallon, L. J.; Brotman, R. M.; Ravel, J., Compara-
tive Metagenome-Assembled Genome Analysis of “Candidatus Lachnocurva vaginae, Formerly Known as Bacterial
Vaginosis-Associated Bacterium-1 (BVAB1). Front Cell Infect Microbiol 2020, 10, 117.
111. Ison, C. A.; Hay, P. E., Validation of a simplified grading of Gram stained vaginal smears for use in genitourinary med-
icine clinics. Sex Transm Infect 2002, 78, (6), 413-5.
112. Forsum, U.; Larsson, P. G.; Spiegel, C., Scoring vaginal fluid smears for diagnosis of bacterial vaginosis: need for qual-
ity specifications. Apmis 2008, 116, (2), 156-9.
113. Forsum, U.; Jakobsson, T.; Larsson, P. G.; Schmidt, H.; Beverly, A.; Bjørnerem, A.; Carlsson, B.; Csango, P.; Donders, G.; Hay,
P.; Ison, C.; Keane, F.; McDonald, H.; Moi, H.; Platz-Christensen, J. J.; Schwebke, J., An international study of the interob-
server variation between interpretations of vaginal smear criteria of bacterial vaginosis. Apmis 2002, 110, (11), 811-8.
114. Larsson, P. G.; Carlsson, B.; Fåhraeus, L.; Jakobsson, T.; Forsum, U., Diagnosis of bacterial vaginosis: need for validation
of microscopic image area used for scoring bacterial morphotypes. Sex Transm Infect 2004, 80, (1), 63-7.
115. Taylor-Robinson, D.; Morgan, D. J.; Sheehan, M.; Rosenstein, I. J.; Lamont, R. F., Relation between Gram-stain and
clinical criteria for diagnosing bacterial vaginosis with special reference to Gram grade II evaluation. Int J STD AIDS
2003, 14, (1), 6-10.
116. Workowski, K. A.; Bachmann, L. H.; Chan, P. A.; Johnston, C. M.; Muzny, C. A.; Park, I.; Reno, H.; Zenilman, J. M.; Bolan,
G. A., Sexually Transmitted Infections Treatment Guidelines, 2021. MMWR Recomm Rep 2021, 70, (4), 1-187.
117. Muzny, C. A.; Balkus, J.; Mitchell, C.; Sobel, J. D.; Workowski, K.; Marrazzo, J.; Schwebke, J. R., Diagnosis and Manage-
ment of Bacterial Vaginosis: Summary of Evidence Reviewed for the 2021 Centers for Disease Control and Preven-
tion Sexually Transmitted Infections Treatment Guidelines. Clin Infect Dis 2022, 74, (Suppl_2), S144-s151.
118. Klebanoff, M. A.; Hauth, J. C.; MacPherson, C. A.; Carey, J. C.; Heine, R. P.; Wapner, R. J.; Iams, J. D.; Moawad, A.;
Miodovnik, M.; Sibai, B. M.; vanDorsten, J. P.; Dombrowski, M. P., Time course of the regression of asymptomatic
bacterial vaginosis in pregnancy with and without treatment. Am J Obstet Gynecol 2004, 190, (2), 363-70.
119. Sherrard, J.; Wilson, J.; Donders, G.; Mendling, W.; Jensen, J. S., 2018 European (IUSTI/WHO) International Union
against sexually transmitted infections (IUSTI) World Health Organisation (WHO) guideline on the management of
vaginal discharge. Int J STD AIDS 2018, 29, (13), 1258-1272.
120. Oduyebo, O. O.; Anorlu, R. I.; Ogunsola, F. T., The effects of antimicrobial therapy on bacterial vaginosis in non-preg-
nant women. Cochrane Database Syst Rev 2009, (3), Cd006055.
121. Livengood, C. H., 3rd; Soper, D. E.; Sheehan, K. L.; Fenner, D. E.; Martens, M. G.; Nelson, A. L.; Ismail, M.; Thorp, J. M.; Lappin,
M.; Long, B. J.; Blackwelder, T.; Sweet, R. L.; Sagov, S., Comparison of once-daily and twice-daily dosing of 0.75% metro-
nidazole gel in the treatment of bacterial vaginosis. Sex Transm Dis 1999, 26, (3), 137-42.
122. Schwebke, J. R.; Marrazzo, J.; Beelen, A. P.; Sobel, J. D., A Phase 3, Multicenter, Randomized, Double-Blind, Vehi-
cle-Controlled Study Evaluating the Safety and Efficacy of Metronidazole Vaginal Gel 1.3% in the Treatment of Bac-
terial Vaginosis. Sex Transm Dis 2015, 42, (7), 376-81.
123. Retamal-Valdes, B.; Tavares, A. P. L.; Monique, S.; Pereira da Silva, H. D.; Mestnik, M. J.; Duarte, P. M.; Miranda, T. S.;
Borges, I.; Soares, G. M. S.; Faveri, M.; Castro Dos Santos, N.; Graças, Y. T. D.; Souto, M. L. S.; Giudicissi, M.; Romito, G. A.;
Saraiva, L.; Pannuti, C. M.; Figueiredo, L. C.; Feres, M., Adverse events of metronidazole and amoxicillin: Retrospective
analysis of a large data set of five randomized clinical trials. J Clin Periodontol 2022.
90
124. Slimings, C.; Riley, T. V., Antibiotics and hospital-acquired Clostridium difficile infection: update of systematic review
and meta-analysis. J Antimicrob Chemother 2014, 69, (4), 881-91.
125. Hanson, J. M.; McGregor, J. A.; Hillier, S. L.; Eschenbach, D. A.; Kreutner, A. K.; Galask, R. P.; Martens, M., Metronidazole
for bacterial vaginosis. A comparison of vaginal gel vs. oral therapy. J Reprod Med 2000, 45, (11), 889-96.
126. Joesoef, M. R.; Schmid, G. P.; Hillier, S. L., Bacterial vaginosis: review of treatment options and potential clinical indi-
cations for therapy. Clin Infect Dis 1999, 28 Suppl 1, S57-65.
127. Joesoef, M. R.; Schmid, G. P., Bacterial vaginosis: review of treatment options and potential clinical indications for
therapy. Clin Infect Dis 1995, 20 Suppl 1, S72-9.
128. Badawi, N. M.; Elkafrawy, M. A.; Yehia, R. M.; Attia, D. A., Clinical comparative study of optimized metronidazole loaded lipid
nanocarrier vaginal emulgel for management of bacterial vaginosis and its recurrence. Drug Deliv 2021, 28, (1), 814-825.
129. Paavonen, J.; Mangioni, C.; Martin, M. A.; Wajszczuk, C. P., Vaginal clindamycin and oral metronidazole for bacterial
vaginosis: a randomized trial. Obstet Gynecol 2000, 96, (2), 256-60.
130. Faro, S.; Skokos, C. K., The efficacy and safety of a single dose of Clindesse vaginal cream versus a seven-dose reg-
imen of Cleocin vaginal cream in patients with bacterial vaginosis. Infect Dis Obstet Gynecol 2005, 13, (3), 155-60.
131. Donders, G.; Bellen, G.; Donders, F.; Pinget, J.; Vandevelde, I.; Michiels, T.; Byamughisa, J., Improvement of abnormal
vaginal flora in Ugandan women by self-testing and short use of intravaginal antimicrobials. Eur J Clin Microbiol
Infect Dis 2017, 36, (4), 731-738.
132. Mendling, W.; Weissenbacher, E. R.; Gerber, S.; Prasauskas, V.; Grob, P., Use of locally delivered dequalinium chloride
in the treatment of vaginal infections: a review. Arch Gynecol Obstet 2016, 293, (3), 469-84.
133. Tinidazole (Tindamax)--a new option for treatment of bacterial vaginosis. Med Lett Drugs Ther 2007, 49, (1269), 73-4.
134. FDA Tindamax® (tinidazole) tablets for oral use. https://www.accessdata.fda.gov/drugsatfda_docs/la-
bel/2007/021618s003lbl.pdf
135. Schwebke, J. R.; Desmond, R. A., Tinidazole vs metronidazole for the treatment of bacterial vaginosis. Am J Obstet
Gynecol 2011, 204, (3), 211.e1-6.
136. Abd El Aziz, M. A.; Sharifipour, F.; Abedi, P.; Jahanfar, S.; Judge, H. M., Secnidazole for treatment of bacterial vaginosis:
a systematic review. BMC Womens Health 2019, 19, (1), 121.
137. Pentikis, H.; Adetoro, N.; Tipping, D.; Levy, S., An Integrated Efficacy and Safety Analysis of Single-Dose Secnidazole
2g in the Treatment of Bacterial Vaginosis. Reprod Sci 2020, 27, (2), 523-528.
138. Elghazaly, S. M.; Hamam, K. M.; Badawy, M. M.; Yakoub Agha, N. A.; Samy, A.; Abbas, A. M., Efficacy and safety of single
dose of oral secnidazole 2 g in treatment of bacterial vaginosis: A systematic review and meta-analysis. Eur J Obstet
Gynecol Reprod Biol 2019, 238, 125-131.
139. Schwebke, J. R.; Morgan, F. G., Jr.; Koltun, W.; Nyirjesy, P., A phase-3, double-blind, placebo-controlled study of the
effectiveness and safety of single oral doses of secnidazole 2 g for the treatment of women with bacterial vaginosis.
Am J Obstet Gynecol 2017, 217, (6), 678.e1-678.e9.
140. Plummer, E. L.; Bradshaw, C. S.; Doyle, M.; Fairley, C. K.; Murray, G. L.; Bateson, D.; Masson, L.; Slifirski, J.; Tachedjian, G.;
Vodstrcil, L. A., Lactic acid-containing products for bacterial vaginosis and their impact on the vaginal microbiota: A
systematic review. PLoS One 2021, 16, (2), e0246953.
141. McCormack, W. M.; Covino, J. M.; Thomason, J. L.; Eschenbach, D. A.; Mou, S.; Kapernick, P.; McGregor, J.; Rein, M.
F.; Hillier, S. L., Comparison of clindamycin phosphate vaginal cream with triple sulfonamide vaginal cream in the
treatment of bacterial vaginosis. Sex Transm Dis 2001, 28, (10), 569-75.
142. Wathne, B.; Holst, E.; Hovelius, B.; Mårdh, P. A., Erythromycin versus metronidazole in the treatment of bacterial
vaginosis. Acta Obstet Gynecol Scand 1993, 72, (6), 470-4.
143. Piot, P., Bacterial vaginosis. An evaluation of treatment. Scand J Urol Nephrol Suppl 1984, 86, 229-35.
144. Wewalka, G.; Stary, A.; Bosse, B.; Duerr, H. E.; Reimer, K., Efficacy of povidone-iodine vaginal suppositories in the
treatment of bacterial vaginosis. Dermatology 2002, 204 Suppl 1, 79-85.
145. Duff, P.; Lee, M. L.; Hillier, S. L.; Herd, L. M.; Krohn, M. A.; Eschenbach, D. A., Amoxicillin treatment of bacterial vaginosis
during pregnancy. Obstet Gynecol 1991, 77, (3), 431-5.
146. Mancuso, A. C.; Widdice, L. E.; Hughes, B. L.; Schlievert, P.; Swamy, G. K.; Stockdale, C. K.; Bernstein, D. I.; Winokur, P.
L., Five Percent Monolaurin Vaginal Gel for the Treatment of Bacterial Vaginosis: A Randomized Placebo-Controlled
Trial. J Low Genit Tract Dis 2020, 24, (3), 277-283.
147. Schoeman, J.; Steyn, P. S.; Odendaal, H. J.; Grové, D., Bacterial vaginosis diagnosed at the first antenatal visit better
predicts preterm labour than diagnosis later in pregnancy. J Obstet Gynaecol 2005, 25, (8), 751-3.
148. Armstrong-Buisseret, L.; Brittain, C.; Kai, J.; David, M.; Anstey Watkins, J.; Ozolins, M.; Jackson, L.; Abdali, Z.; Hepburn,
T.; Griffiths, F.; Montgomery, A.; Daniels, J.; Manley, A.; Dean, G.; Ross, J. D., Lactic acid gel versus metronidazole for
recurrent bacterial vaginosis in women aged 16 years and over: the VITA RCT. Health Technol Assess 2022, 26, (2), 1-170.
91
149. Tidbury, F. D.; Langhart, A.; Weidlinger, S.; Stute, P., Non-antibiotic treatment of bacterial vaginosis-a systematic
review. Arch Gynecol Obstet 2021, 303, (1), 37-45.
150. Abu-Zaid, A.; Alshahrani, M. S.; Bakhsh, H.; Miski, N. T.; Abuzaid, M.; Alomar, O.; Jabrah, E.; Jamjoom, M. Z.; Salem,
H.; Al-Badawi, I. A.; Baradwan, S., Astodrimer gel for treatment of bacterial vaginosis: A systematic review and me-
ta-analysis of randomized controlled trials. Int J Clin Pract 2021, 75, (7), e14165.
151. Liu, H. F.; Yi, N., A systematic review and meta-analysis on the efficacy of probiotics for bacterial vaginosis. Eur Rev
Med Pharmacol Sci 2022, 26, (1), 90-98.
152. Jeng, H. S.; Yan, T. R.; Chen, J. Y., Treating vaginitis with probiotics in non-pregnant females: A systematic review and
meta-analysis. Exp Ther Med 2020, 20, (4), 3749-3765.
153. Vieira-Baptista, P.; De Seta, F.; Verstraelen, H.; Ventolini, G.; Lonnee-Hoffmann, R.; Lev-Sagie, A., The Vaginal Microbi-
ome: V. Therapeutic Modalities of Vaginal Microbiome Engineering and Research Challenges. J Low Genit Tract Dis
2022, 26, (1), 99-104.
154. Husain, S.; Allotey, J.; Drymoussi, Z.; Wilks, M.; Fernandez-Felix, B. M.; Whiley, A.; Dodds, J.; Thangaratinam, S.; Mc-
Court, C.; Prosdocimi, E. M.; Wade, W. G.; de Tejada, B. M.; Zamora, J.; Khan, K.; Millar, M., Effects of oral probiotic
supplements on vaginal microbiota during pregnancy: a randomised, double-blind, placebo-controlled trial with
microbiome analysis. Bjog 2020, 127, (2), 275-284.
155. Van de Wijgert, J.; Verwijs, M. C., Lactobacilli-containing vaginal probiotics to cure or prevent bacterial or fungal vaginal
dysbiosis: a systematic review and recommendations for future trial designs. Bjog 2020, 127, (2), 287-299.
156. Marcotte, H.; Larsson, P. G.; Andersen, K . K.; Zuo, F.; Mikkelsen, L. S.; Brandsborg, E.; Gray, G.; Laher, F.; Otwombe, K., An
exploratory pilot study evaluating the supplementation of standard antibiotic therapy with probiotic lactobacilli in
south African women with bacterial vaginosis. BMC Infect Dis 2019, 19, (1), 824.
157. Brocklehurst, P.; Gordon, A.; Heatley, E.; Milan, S. J., Antibiotics for treating bacterial vaginosis in pregnancy. Cochrane
Database Syst Rev 2013, (1), Cd000262.
158. Leitich, H.; Brunbauer, M.; Bodner-Adler, B.; Kaider, A.; Egarter, C.; Husslein, P., Antibiotic treatment of bacterial vagi-
nosis in pregnancy: a meta-analysis. Am J Obstet Gynecol 2003, 188, (3), 752-8.
159. Riggs, M. A.; Klebanoff, M. A., Treatment of vaginal infections to prevent preterm birth: a meta-analysis. Clin Obstet
Gynecol 2004, 47, (4), 796-807; discussion 881-2.
160. Caro-Patón, T.; Carvajal, A.; Martin de Diego, I.; Martin-Arias, L. H.; Alvarez Requejo, A.; Rodríguez Pinilla, E., Is metro-
nidazole teratogenic? A meta-analysis. Br J Clin Pharmacol 1997, 44, (2), 179-82.
161. Yefet, E.; Colodner, R.; Strauss, M.; Gam Ze Letova, Y.; Nachum, Z., A Randomized Controlled Open Label Crossover
Trial to Study Vaginal Colonization of Orally Administered Lactobacillus Reuteri RC-14 and Rhamnosus GR-1 in Preg-
nant Women at High Risk for Preterm Labor. Nutrients 2020, 12, (4).
162. Haahr, T.; Zacho, J.; Bräuner, M.; Shathmigha, K.; Skov Jensen, J.; Humaidan, P., Reproductive outcome of patients
undergoing in vitro fertilisation treatment and diagnosed with bacterial vaginosis or abnormal vaginal microbiota:
a systematic PRISMA review and meta-analysis. Bjog 2019, 126, (2), 200-207.
163. Vieira-Baptista, P.; Silva-Soares, S.; Lyra, J.; Falcão, V.; Póvoa, A. M.; Calejo, L.; Sousa, S., Wet Mount Microscopy of the
Vaginal Milieu Does Not Predict the Outcome of Fertility Treatments: A Cross-sectional Study. J Low Genit Tract Dis 2022.
164. Van Oostrum, N.; De Sutter, P.; Meys, J.; Verstraelen, H., Risks associated with bacterial vaginosis in infertility patients:
a systematic review and meta-analysis. Hum Reprod 2013, 28, (7), 1809-15.
165. Nygren, P.; Fu, R.; Freeman, M.; Bougatsos, C.; Klebanoff, M.; Guise, J. M., Evidence on the benefits and harms of
screening and treating pregnant women who are asymptomatic for bacterial vaginosis: an update review for the
U.S. Preventive Services Task Force. Ann Intern Med 2008, 148, (3), 220-33.
166. Rebouças, K. F.; Eleutério, J., Jr.; Peixoto, R. C.; Costa, A. P. F.; Cobucci, R. N.; Gonçalves, A. K., Treatment of bacterial vagino-
sis before 28weeks of pregnancy to reduce the incidence of preterm labor. Int J Gynaecol Obstet 2019, 146, (3), 271-276.
167. Subtil, D.; Brabant, G.; Tilloy, E.; Devos, P.; Canis, F.; Fruchart, A.; Bissinger, M. C.; Dugimont, J. C.; Nolf, C.; Hacot, C.;
Gautier, S.; Chantrel, J.; Jousse, M.; Desseauve, D.; Plennevaux, J. L.; Delaeter, C.; Deghilage, S.; Personne, A.; Joyez,
E.; Guinard, E.; Kipnis, E.; Faure, K.; Grandbastien, B.; Ancel, P. Y.; Goffinet, F.; Dessein, R., Early clindamycin for bacte-
rial vaginosis in pregnancy (PREMEVA): a multicentre, double-blind, randomised controlled trial. Lancet 2018, 392,
(10160), 2171-2179.
168. Yudin, M. H.; Money, D. M., No. 211-Screening and Management of Bacterial Vaginosis in Pregnancy. J Obstet Gynae-
col Can 2017, 39, (8), e184-e191.
169. Soper, D. E., Bacterial vaginosis and surgical site infections. Am J Obstet Gynecol 2020, 222, (3), 219-223.
170. Penney, G. C.; Thomson, M.; Norman, J.; McKenzie, H.; Vale, L.; Smith, R.; Imrie, M., A randomised comparison of
strategies for reducing infective complications of induced abortion. Br J Obstet Gynaecol 1998, 105, (6), 599-604.
92
171. Larsson, P. G.; Platz-Christensen, J. J.; Dalaker, K.; Eriksson, K.; Fåhraeus, L.; Irminger, K.; Jerve, F.; Stray-Pedersen, B.;
Wölner-Hanssen, P., Treatment with 2% clindamycin vaginal cream prior to first trimester surgical abortion to re-
duce signs of postoperative infection: a prospective, double-blinded, placebo-controlled, multicenter study. Acta
Obstet Gynecol Scand 2000, 79, (5), 390-6.
172. Miller, L.; Thomas, K.; Hughes, J. P.; Holmes, K. K.; Stout, S.; Eschenbach, D. A., Randomised treatment trial of bacterial
vaginosis to prevent post-abortion complication. Bjog 2004, 111, (9), 982-8.
173. Crowley, T.; Low, N.; Turner, A.; Harvey, I.; Bidgood, K.; Horner, P., Antibiotic prophylaxis to prevent post-abortal upper
genital tract infection in women with bacterial vaginosis: randomised controlled trial. Bjog 2001, 108, (4), 396-402.
174. Larsson, P. G.; Platz-Christensen, J. J.; Thejls, H.; Forsum, U.; Påhlson, C., Incidence of pelvic inflammatory disease after
first-trimester legal abortion in women with bacterial vaginosis after treatment with metronidazole: a double-blind,
randomized study. Am J Obstet Gynecol 1992, 166, (1 Pt 1), 100-3.
175. Amaya-Guio, J.; Viveros-Carreño, D. A.; Sierra-Barrios, E. M.; Martinez-Velasquez, M. Y.; Grillo-Ardila, C. F., Antibiotic treat-
ment for the sexual partners of women with bacterial vaginosis. Cochrane Database Syst Rev 2016, 10, (10), Cd011701.
176. Ratten, L. K.; Plummer, E. L.; Murray, G. L.; Danielewski, J.; Fairley, C. K.; Garland, S. M.; Hocking, J. S.; Tachedjian, G.;
Chow, E.; Bradshaw, C. S.; Vodstrcil, L. A., Sex is associated with the persistence of non-optimal vaginal microbiota
following treatment for bacterial vaginosis: a prospective cohort study. Bjog 2021, 128, (4), 756-767.
177. Schwebke, J. R.; Lensing, S. Y.; Lee, J.; Muzny, C. A.; Pontius, A.; Woznicki, N.; Aguin, T.; Sobel, J. D., Treatment of Male
Sexual Partners of Women With Bacterial Vaginosis: A Randomized, Double-Blind, Placebo-Controlled Trial. Clin In-
fect Dis 2021, 73, (3), e672-e679.
178. Plummer, E. L.; Vodstrcil, L. A.; Doyle, M.; Danielewski, J. A.; Murray, G. L.; Fehler, G.; Fairley, C. K.; Bulach, D. M.; Garland,
S. M.; Chow, E. P. F.; Hocking, J. S.; Bradshaw, C. S., A Prospective, Open-Label Pilot Study of Concurrent Male Partner
Treatment for Bacterial Vaginosis. mBio 2021, 12, (5), e0232321.
179. Muzny, C. A.; Schwebke, J. R., Asymptomatic Bacterial Vaginosis: To Treat or Not to Treat? Curr Infect Dis Rep 2020, 22, (12).
180. Bilardi, J. E.; Walker, S. M.; Temple-Smith, M. J.; McNair, R. P.; Mooney-Somers, J.; Vodstrcil, L. A.; Bellhouse, C. E.; Fairley,
C. K.; Bradshaw, C. S., Women view key sexual behaviours as the trigger for the onset and recurrence of bacterial
vaginosis. PLoS One 2017, 12, (3), e0173637.
181. Bradshaw, C. S.; Morton, A. N.; Hocking, J.; Garland, S. M.; Morris, M. B.; Moss, L. M.; Horvath, L. B.; Kuzevska, I.; Fairley,
C. K., High recurrence rates of bacterial vaginosis over the course of 12 months after oral metronidazole therapy and
factors associated with recurrence. J Infect Dis 2006, 193, (11), 1478-86.
182. Muzny, C. A.; Sobel, J. D., The Role of Antimicrobial Resistance in Refractory and Recurrent Bacterial Vaginosis and
Current Recommendations for Treatment. Antibiotics (Basel) 2022, 11, (4).
183. Sobel, J. D.; Ferris, D.; Schwebke, J.; Nyirjesy, P.; Wiesenfeld, H. C.; Peipert, J.; Soper, D.; Ohmit, S. E.; Hillier, S. L., Sup-
pressive antibacterial therapy with 0.75% metronidazole vaginal gel to prevent recurrent bacterial vaginosis. Am J
Obstet Gynecol 2006, 194, (5), 1283-9.
184. McClelland, R. S.; Richardson, B. A.; Hassan, W. M.; Chohan, V.; Lavreys, L.; Mandaliya, K.; Kiarie, J.; Jaoko, W.; Ndinya-Acho-
la, J. O.; Baeten, J. M.; Kurth, A. E.; Holmes, K. K., Improvement of vaginal health for Kenyan women at risk for acquisition
of human immunodeficiency virus type 1: results of a randomized trial. J Infect Dis 2008, 197, (10), 1361-8.
185. Reichman, O.; Akins, R.; Sobel, J. D., Boric acid addition to suppressive antimicrobial therapy for recurrent bacterial
vaginosis. Sex Transm Dis 2009, 36, (11), 732-4.
186. Surapaneni, S.; Akins, R.; Sobel, J. D., Recurrent Bacterial Vaginosis: An Unmet Therapeutic Challenge. Experience
With a Combination Pharmacotherapy Long-Term Suppressive Regimen. Sex Transm Dis 2021, 48, (10), 761-765.
187. Unemo, M.; Bradshaw, C. S.; Hocking, J. S.; de Vries, H. J. C.; Francis, S. C.; Mabey, D.; Marrazzo, J. M.; Sonder, G. J. B.;
Schwebke, J. R.; Hoornenborg, E.; Peeling, R. W.; Philip, S. S.; Low, N.; Fairley, C. K., Sexually transmitted infections:
challenges ahead. Lancet Infect Dis 2017, 17, (8), e235-e279.
188. Krasnopolsky, V. N.; Prilepskaya, V. N.; Polatti, F.; Zarochentseva, N. V.; Bayramova, G. R.; Caserini, M.; Palmieri, R.,
Efficacy of vitamin C vaginal tablets as prophylaxis for recurrent bacterial vaginosis: a randomised, double-blind,
placebo-controlled clinical trial. J Clin Med Res 2013, 5, (4), 309-15.
189. Verstraelen, H.; Verhelst, R.; Roelens, K.; Temmerman, M., Antiseptics and disinfectants for the treatment of bacterial
vaginosis: a systematic review. BMC Infect Dis 2012, 12, 148.
190. Cohen, C. R.; Wierzbicki, M. R.; French, A. L.; Morris, S.; Newmann, S.; Reno, H.; Green, L.; Miller, S.; Powell, J.; Parks, T.; Hemmer-
ling, A., Randomized Trial of Lactin-V to Prevent Recurrence of Bacterial Vaginosis. N Engl J Med 2020, 382, (20), 1906-1915.
191. Heczko, P. B.; Tomusiak, A.; Adamski, P.; Jakimiuk, A. J.; Stefański, G.; Mikołajczyk-Cichońska, A.; Suda-Szczurek, M.;
Strus, M., Supplementation of standard antibiotic therapy with oral probiotics for bacterial vaginosis and aerobic
vaginitis: a randomised, double-blind, placebo-controlled trial. BMC Womens Health 2015, 15, 115.
93
192. Xie, H. Y.; Feng, D.; Wei, D. M.; Mei, L.; Chen, H.; Wang, X.; Fang, F., Probiotics for vulvovaginal candidiasis in non-preg-
nant women. Cochrane Database Syst Rev 2017, 11, (11), Cd010496.
193. Laue, C.; Papazova, E.; Liesegang, A.; Pannenbeckers, A.; Arendarski, P.; Linnerth, B.; Domig, K. J.; Kneifel, W.; Petricevic,
L.; Schrezenmeir, J., Effect of a yoghurt drink containing Lactobacillus strains on bacterial vaginosis in women - a
double-blind, randomised, controlled clinical pilot trial. Benef Microbes 2018, 9, (1), 35-50.
194. Vodstrcil, L. A.; Plummer, M. E.; Fairley, C. K.; Tachedjian, G.; Law, M. G.; Hocking, J. S.; Worthington, M. K.; Grant, M. M.;
Okoko, N.; Bradshaw, C. S., Combined oral contraceptive pill-exposure alone does not reduce the risk of bacterial
vaginosis recurrence in a pilot randomised controlled trial. Sci Rep 2019, 9, (1), 3555.
195. Vieira-Baptista, P.; Silva, A. R.; Costa, M.; Aguiar, T.; Saldanha, C.; Sousa, C., Clinical validation of a new molecular test
(Seegene Allplex™ Vaginitis) for the diagnosis of vaginitis: a cross-sectional study. Bjog 2021, 128, (8), 1344-1352.
196. Demirbilek, M.; Can, F.; Güleç, A. T.; Kuşçu, E.; Kayhan, Z.; Haberal, M., Incidence of bacterial vaginosis in renal trans-
plant recipients. Transplant Proc 2003, 35, (7), 2696-7.
197. Miller, E. A.; Beasley, D. E.; Dunn, R. R.; Archie, E. A., Lactobacilli Dominance and Vaginal pH: Why Is the Human Vaginal
Microbiome Unique? Front Microbiol 2016, 7, 1936.
198. Bayar, E.; Bennett, P. R.; Chan, D.; Sykes, L.; MacIntyre, D. A., The pregnancy microbiome and preterm birth. Semin
Immunopathol 2020, 42, (4), 487-499.
199. Gupta, S.; Kakkar, V.; Bhushan, I., Crosstalk between Vaginal Microbiome and Female Health: A review. Microb Pathog
2019, 136, 103696.
200. Serrano, M. G.; Parikh, H. I.; Brooks, J. P.; Edwards, D. J.; Arodz, T. J.; Edupuganti, L.; Huang, B.; Girerd, P. H.; Bokhari, Y. A.;
Bradley, S. P.; Brooks, J. L.; Dickinson, M. R.; Drake, J. I.; Duckworth, R. A., 3rd; Fong, S. S.; Glascock, A. L.; Jean, S.; Jime-
nez, N. R.; Khoury, J.; Koparde, V. N.; Lara, A. M.; Lee, V.; Matveyev, A. V.; Milton, S. H.; Mistry, S. D.; Rozycki, S. K.; Sheth,
N. U.; Smirnova, E.; Vivadelli, S. C.; Wijesooriya, N. R.; Xu, J.; Xu, P.; Chaffin, D. O.; Sexton, A. L.; Gravett, M. G.; Rubens,
C. E.; Hendricks-Muñoz, K. D.; Jefferson, K. K.; Strauss, J. F., 3rd; Fettweis, J. M.; Buck, G. A., Racioethnic diversity in the
dynamics of the vaginal microbiome during pregnancy. Nat Med 2019, 25, (6), 1001-1011.
201. MacIntyre, D. A.; Chandiramani, M.; Lee, Y. S.; Kindinger, L.; Smith, A.; Angelopoulos, N.; Lehne, B.; Arulkumaran, S.;
Brown, R.; Teoh, T. G.; Holmes, E.; Nicoholson, J. K.; Marchesi, J. R.; Bennett, P. R., The vaginal microbiome during
pregnancy and the postpartum period in a European population. Sci Rep 2015, 5, 8988.
202. Hay, P., Bacterialvaginosis. F1000Res 2017, 6, 1761.
203. Kyrgiou, M.; Moscicki, A. B., Vaginal microbiome and cervical cancer. Semin Cancer Biol 2022, 86, (Pt 3), 189-198.
204. Vieira-Baptista, P.; Eleutério Jr., J., Diagnosis of vaginitis: time to improve and move on. DST - J bras Doenças Sex
Transm 2020, 32, (e203214), 1-3.
205. Castro, J.; Jefferson, K. K.; Cerca, N., Genetic Heterogeneity and Taxonomic Diversity among Gardnerella Species.
Trends Microbiol 2020, 28, (3), 202-211.
206. Le Roy, C.; Bébéar, C.; Pereyre, S., Performance of Three Commercial Molecular Diagnostic Assays for the Simultane-
ous Detection of Mycoplasma genitalium and Macrolide Resistance. J Clin Microbiol 2021, 59, (6).
207. Turpin, R.; Slopen, N.; Borgogna, J. C.; Yeoman, C. J.; He, X.; Miller, R. S.; Klebanoff, M. A.; Ravel, J.; Brotman, R. M.,
Perceived Stress and Molecular Bacterial Vaginosis in the National Institutes of Health Longitudinal Study of Vaginal
Flora. Am J Epidemiol 2021, 190, (11), 2374-2383.
208. Li, C.; Wang, T.; Li, Y.; Zhang, T.; Wang, Q.; He, J.; Wang, L.; Li, L.; Yang, N.; Fang, Y., Probiotics for the treatment of women with
bacterial vaginosis: A systematic review and meta-analysis of randomized clinical trials. Eur J Pharmacol 2019, 864, 172660.
209. Lev-Sagie, A.; Goldman-Wohl, D.; Cohen, Y.; Dori-Bachash, M.; Leshem, A.; Mor, U.; Strahilevitz, J.; Moses, A. E.; Shap-
iro, H.; Yagel, S.; Elinav, E., Vaginal microbiome transplantation in women with intractable bacterial vaginosis. Nat
Med 2019, 25, (10), 1500-1504.
210. Yockey, L. J.; Hussain, F. A.; Bergerat, A.; Reissis, A.; Worrall, D.; Xu, J.; Gomez, I.; Bloom, S. M.; Mafunda, N. A.; Kelly, J.;
Kwon, D. S.; Mitchell, C. M., Screening and characterization of vaginal fluid donations for vaginal microbiota trans-
plantation. Sci Rep 2022, 12, (1), 17948.
94
95
4.1
Introduction
Vulvovaginal candidiasis or candidosis (VVC) is a common condition. Frequently trivialized
and as a result mismanaged, VVC has significant adverse effects on women’s overall health.
When diagnosed and managed correctly, these ill effects can be minimized. This document
represents our current understanding of the condition and highlights the best practices for
diagnosis and treatment of uncomplicated and complicated VVC.
4.2
Etiology and pathophysiology
VVC is an inflammatory disease caused by Candida spp. that affects the female lower genital
tract. Less commonly, there may be other kinds of fungal organisms causing disease in
these organs which together with VVC represent female lower genital tract mycosis. Among
the various species, Candida albicans is by far the most frequent infecting organism, causing
more than 80% of cases. However, a broad range of other species may cause VVC, includ-
ing C. glabrata, C. guilliermondii and C. tropicalis.1-4 Finally, variants of C. albicans such as C.
africana and C. dubliniensis also play a role in producing VVC.5 Geographical and population
diversity may explain the differences in relative prevalence of the more common species.
Furthermore, reported experiences from tertiary care centers, which tend to treat more dif-
ficult cases, may exaggerate the contributions of non-albicans Candida species to the overall
burden of VVC. On a final note, it should be mentioned that, due to taxonomic changes,
some clinically important causes of VVC, including C. glabrata, are now being classified as
other species.6 For example, C. glabrata is now known as Nakaseomyces glabrata. To avoid
confusion and since these name changes have not been instituted in clinical settings, we
will be using the more widely used nomenclature.
4
CANDIDIASIS
(alphabetical order)
Sophia Ehrström
José Martinez de Oliveira
Werner Mendling
Paul Nyirjesy
96
In order to cause disease, Candida spp. must first colonize the vaginal epithelium. With col-
onization, the presence of yeast may be temporary and seems to frequently be followed
by its elimination by normal vaginal defense mechanisms involving neutrophils and mac-
rophages.7 In the absence of pro-inflammatory mediators, long-standing asymptomatic col-
onization may take place. At some point, in women with symptomatic VVC, the infecting or-
ganism becomes pathogenic. It is not fully understood what induces Candida spp. to cause
an actual infection. At some point, yeasts may express greater virulence factors through
different mechanisms, like morphologic changes (dimorphism), proteinase secretion and
cell surface composition changes.8 Histologic evaluation of the vaginal wall in women with
acute VVC has demonstrated evidence of superficial tissue invasion but has failed to show
any sign of biofilm formation.9 In some women, symptoms and signs may occur with a low
burden of Candida; it is thought that symptoms in these cases may be due to an allergic or
inflammatory response to the presence of the yeast, mediated by an immunologic, perhaps
allergic mechanism.10
In women with repeated episodes of VVC, the source of the organism remains controversial.
Proposed sources include an intestinal reservoir, reinfection from a sexual partner, or failure
to fully clear the organism after initial infection (the vaginal relapse theory). For women
with recurrent VVC where vaginal relapse seems to play a very large role, evaluating and
treating an intestinal or partner source remain controversial and seems unhelpful.
4.3
Prevalence and epidemiology
The estrogenized vagina is colonized by Candida spp. in at least 20% of pregnant wom-
en and 30% of immunocompromised patients, if examined by culture.11 When sequencing
methods are used, vaginal fungal colonization can be found in >60% of all premenopausal
women, of which the most abundant species is C. albicans in >80%.12 It is estimated, that
30 – 50 % of all women at least once during their life suffer from VVC and that most of these
women are in their reproductive years.13, 14
Without a correct diagnosis of VVC, it is difficult to estimate the prevalence and incidence
of this disease in different populations. Further complicating these estimates, VVC is not a re-
portable disease and treatment is often given based on symptoms. In some studies, more than
60% of US women who receive treatment for VVC are diagnosed without either microscopy or cul-
ture.15 This finding is probably similar in all countries. Thus, the historical estimate that 75%
of women will suffer at least one episode of VVC during her lifetime has been based on very
little data.16 Nevertheless, recent studies seem to confirm the historical estimations.17
Worldwide, recurrent vulvovaginal candidiasis (RVVC), which is currently defined as three or
more confirmed episodes per year, affects about 138 million women annually (range 103–
172 million) and causes substantial morbidity and economic burden, with a global annual
prevalence of 3871 per 100,000 women and 372 million women affected by RVVC over their
lifetime.18 The 19–35 year age group has the highest prevalence (9%). A qualified online
survey questionnaire with responses from 284 women between 2016 und 2018, in three
97
university-affiliated gynecological clinics in the US, revealed at least one lifetime episode in
77.5%, 1-3 in 29.0%, 4–10 in 28.4% and >10 episodes in 43.6% of the participants; 44.3%
with an age of 26–40 years suffered from >3 episodes per year at some point.17 The quality of
life of affected women is heavily impaired, especially in recurrent cases, with impacts similar
to those seen in women with chronic obstructive lung disease or asthma.19, 20
4.4
Risk factors
A series of conditions are considered to favor the establishment of symptomatic VVC. Among
them are non-pathological (pregnancy, estrogen-containing contraceptives, and meno-
pause hormone therapies), behavioral (frequent sexual intercourse or multiple partners),
or disease-related events and comorbidities (antibiotic use, innate reduced cellular immu-
nity, iatrogenic or spontaneous immunosuppressive conditions, poorly controlled diabetes
mellitus).4 With antibiotic use specifically, it is important to note that most women who take
antibiotics will not get VVC.21 When they do, the mechanism is thought to be either from
the induction of growth of Candida spp. in either the intestine or the vagina. In diabetics
the use of sodium-glucose cotransporter 2 (SGLT2) inhibitors (i.e. canaglifozin, dapaglifozin,
empaglifozin) to control the disease do promote VVC.22
Suggesting that certain women are simply more prone to VVC, a history of episodes of VVC
seems to put women at higher risk of new ones. In women with persistence of VVC, the cul-
ture-proven presence of yeast in the vagina, particularly during treatment or immediately
after the end of treatment, suggests clinical resistance, whereas clinical recurrence after an
asymptomatic and culture negative episode may represent re-infection.
4.5
Classification of vulvovaginal candidiasis
Because episodes of VVC can vary between affected women and because these variations may
affect treatment outcomes, providers should make every effort to classify infection. Of the var-
ious systems in use, perhaps the most widely known is the one first proposed in 199823 and still
recommended by the US Centers for Disease Control and Prevention.24 (Table 4.1)
TABLE 4.1 Classification of vulvovaginal candidiasis and its clinical implications.
Type of infection Clinical implications
Uncomplicated All treatments with similar efficacy
Choice can be individualized
Complicated
Severe or predisposing factors
Recurrent or chronic
Non-albicans Candida vulvovaginal candidiasis
More likely to fail short course therapy
50% idiopathic
Usually requires maintenance therapy
More likely to fail azole therapy
98
This system distinguishes between uncomplicated and complicated VVC. In general, uncompli-
cated VVC affects women with no predisposing factors for yeast infections, such as diabetes or im-
munosuppressive conditions, are sporadic (two or less per year) episodes, with mild or moderate
symptoms, and the infection is caused by C. albicans. In general, women with uncomplicated VVC
will respond to pretty much any of the treatment options available. Complicated VVC consists
of women with any one of the following: 1) severe infection, 2) recurrent episodes (defined
as three or more episodes in the previous year), 3) conditions such as diabetes, underlying
immunodeficiency, or immunosuppressive therapy, or 4) infections due to a non-albicans
Candida. In general, women with complicated VVC are less likely to respond to standard
regimens of antifungal therapy and will require closer follow-up and more aggressive treat-
ments. However, it is important to emphasize that each category of complicated VVC de-
serves its own individualized approach to management.
Severe infections
VVC-related symptoms and signs can be assessed on a semi-quantitative basis.25 Women
with severe VVC are more likely to fail standard treatment for VVC and should receive more
prolonged courses of treatment.
Recurrent infections
Based largely on consensus expert opinion, the current definition of RVVC is three or more ep-
isodes in the prior 12 months. In most women with RVVC, there are no underlying known pre-
disposing factors to infection, and no further work-up is indicated. In most affected women,
C. albicans is the causative organism. Women with chronic VVC may possibly be a separate
group from those with RVVC, but should also be considered complicated.26 Recently, flucona-
zole-resistant C. albicans organisms seem to be an emerging and particularly problematic
cause of RVVC.27 It is well accepted that certain conditions, including diabetes, treatments
for diabetes such as SGLT2 inhibitors and immunosuppressive conditions and medications
increase Candida spp. colonization and infection.22 HIV is not considered a predisposing con-
dition since HIV-positive and HIV-negative women have similar responses to treatment.28
Non-
albicans Candida
infections
In general, it is felt that 90-95% of women with VVC have infections due to C. albicans. How-
ever, other organisms such as C. glabrata, C. parapsilosis, C. krusei, C. tropicalis and Saccha-
romyces cerevisiae can sometimes be found in symptomatic women. There exists some con-
troversy into whether these species can cause true vulvovaginal infections, and the extent
to which their presence may simply represent asymptomatic colonization in a woman who
has a different cause for her symptoms. In these cases, treating the organism and having
the patient return after the organism has been suppressed may be the only way to decide
whether this organism is contributing to the patient’s symptoms.29
99
4.6
Signs and symptoms
The symptoms and signs of VVC can be relatively nonspecific. In the classic description of
VVC, affected women will complain of an abnormal thick white discharge, itching, irritation
and burning. They may also complain of external dysuria.30 If sexually active, they may note
dyspareunia. The presentation will often be relatively acute, but it may occur on a repeated ba-
sis in women with recurrent yeast infection or simply be chronic if they have chronic VVC, either
because of a missed diagnosis or an organism which is resistant to antifungal therapy. It should
be emphasized that the symptoms attributed to VVC can have many other potential causes,
including vaginal infections and vulvar dermatoses. The presence of a discharge described as
“cheesy” and of itching, which occurs in 70-90% of women with VVC, increases the likelihood of
VVC; the absence of itching or irritative symptoms makes it less likely.31 On physical examination,
women with VVC may have erythema of the labia majora, minora or vestibule. (Figure 4.1)
They may have swelling of these structures or fissures. If itching is severe, excoriations may
be noted. (Figure 4.2)
On speculum examination, there may be vaginal enanthema. Thrush, which is an adherent
white discharge on the sidewalls of the vagina, may be noted. 31 (Figure 4.3)
Many women with symptomatic VVC will have no significant physical findings.
With both symptoms and signs, VVC represents a spectrum. Affected women may have few
or all of them, and the symptoms and signs may affect the vulva, the vagina, or both.
Figure 4.1 Acute vulvovaginal candidiasis.
Exteriorization of white “cheesy” discharge,
vulvar erythema and edema.
Figure 4.2 Acute vulvovaginal candidiasis.
Erythema and fissures of the intelabial sulci.
100
4.7
Diagnosis
Since asymptomatic colonization with Can-
dida spp. is a common life event, accurate
diagnosis relies on obtaining an appropri-
ate history, as well as the detection of yeast
through some sort of testing modality. Any
sort of diagnosis which does not attempt
to detect the organism runs the risk of
misdiagnosis. In a study of women about
to self-treat for acute VVC, only 33.7% had
VVC alone.32 Similarly, diagnosis of vaginal
infections over the telephone by a nurse
showed poor agreement beyond chance for
diagnosing VVC.33 Thus, in cases where pa-
tients are treated without an examination
and obtaining vaginal samples to exclude
other infections and detect yeast, it is im-
portant to recognize that the likelihood of misdiagnosis is quite high.
In general, it is recommended that office laboratory testing, composed of checking the vag-
inal pH, mixing the vaginal secretions with 10% KOH to detect amines (the whiff test) and
performing microscopic examinations of the discharge mixed with saline and, separately,
10% KOH (important if phase contrast is not being used), be performed in all women with
vulvovaginal symptoms. In women with VVC, the pH will be often normal, the amine test will
Figure 4.3 Acute vulvovaginal candidiasis.
Adherent white discharge on the sidewalls
of the vagina and cervix.
Figure 4.4 Wet mount microscopy (400x, phase contrast).
A– Blastospores (culture positive for C. krusei) B– Hyphae and blastospores (culture positive for C. albicans)
101
be negative, and microscopic examination will reveal blastospores, pseudohyphae or hyphae,
or other fungal elements. (Figure 4.4 and 4.5)
Figure 4.5 Gram stain (1000x, oil immersion).
A– Blastospores B– Hyphae and blastospores
Microscopy, which by its nature is provider-specific, has two main limitations, underdiagnosis
and overdiagnosis. In general, the estimated sensitivity of microscopy, performed in research set-
tings, is about 56%; in community settings, the sensitivity is lower.34 There remain significant
concerns with false positive rates for detecting VVC, which may be as high as 49%.35 Phase
contrast microscopy and training courses can significantly improve the diagnostic competence.36
Nucleic acid amplifications tests (NAATs) for yeast have become widely available and can be
used in symptomatic women. Depending on the country, these tests may be either cleared
by governmental authorities or they may have only undergone validation by a local labora-
tory. In general, the NAATs are quite sensitive (>90%) for C. albicans.34, 37 When it comes to
non-albicans Candida species, performance data for NAAT tests may be difficult to measure
given the low rates of such infections in clinical studies.34, 37 Thus, if providers use NAATs for di-
agnosing VVC, they should be aware of the performance characteristics of the specific lab they
are using and also realize that NAATs may miss yeast infections by less common organisms.
Yeast culture remains the gold standard for confirming the diagnosis for VVC. In a woman
with uncomplicated VVC, culture is probably not necessary as most patients will get better
with treatment. However, in a woman with complicated VVC, culture will help to confirm
the diagnosis, permit speciation of the infecting organism, and make the organism available
for susceptibility testing. In cases where drug resistance is suspected, susceptibility testing
can be considered, but providers should be aware that susceptibility tests done at a pH of 7,
the standard for most clinical laboratories, may vary dramatically from those at lower pHs.38
Thus, susceptibility testing will only add a little insight into which drugs may be of use. In
comparing NAATs to culture, NAATs offers the advantages of sometimes wider availability
and a quicker result, but they are usually more expensive and may miss non-albicans Candida.
102
4.8
Treatment of vulvovaginal candidiasis
Asymptomatic colonization
Although asymptomatic colonization represents the first step to developing symptomatic
disease, asymptomatic women should not be screened for VVC and those who happen to
have a positive culture or its identification in a Pap test do not require any sort of treatment.39
Uncomplicated vulvovaginal candidiasis
Azole antifungals are the treatment of choice in uncomplicated cases of VVC. They come in
a large variety of formulations, such as topical vaginal creams, ointments, and supposito-
ries. Some of the many options available internationally are shown in Table 4.2; it should
be stressed that this list is not exhaustive. The most common treatments, such as local clo-
trimazole, miconazole, or econazole or oral single dose fluconazole resolve up to 80-85%
of cases.40-42 Topical azoles are well tolerated, although side-effects such as slight burning
have been reported in 1-10% of cases; allergic reactions are rare.41 The Centers for Disease
Control and Prevention (CDC) also recommends formulations of tioconazole, butaconazole,
and terconazole; these seem to be less available outside of the US. In general, since one can
expect similar efficacy with all of the available options, treatment should be individualized
depending on drug availability, tolerability, price and patient preference.
TABLE 4.2 Commonly available treatment options for patients with uncomplicated vulvovaginal
candidiasis. The list does not include all options and the options listed may not be available in all
countries
Local treatment (mild symptoms)
Clotrimazole
200 mg vaginal tablets, once daily (3 days)
100 mg vaginal tablets, once daily (7 days)
1% cream, 5 g vaginally once daily (7 days)
500 mg vaginal tablet, once daily (1 day)
1% cream, once daily for 7 days
Econazole 150 mg vaginal suppository, twice daily (1 day)
150 mg vaginal suppository, once daily (3 days)
Fenticonazole 600 mg vaginal capsule, once daily (1 day)
Isoconazole
150 mg vaginal suppository, twice daily (1 day)
150 mg vaginal suppository, once daily (3 days)
600 mg vaginal suppository, once daily (1 day)
Alternative treatment (severe symptoms)
Fluconazole
150 mg orally, single dose
50 mg orally, once daily (7–14 days)
100 mg orally, once daily (14 days) (if immunocompromised)
Itraconazole 100 mg orally 2×2 capsules daily (1 day)
100 mg orally 1×2 capsules daily (3 days)
Nystatin 100,000 units vaginal tablets (14 days)
200,000 units vaginal tablets (6 days)
103
Recurrent vulvovaginal candidiasis due to
C. albicans
Treatments of RVVC differ from country to country, depending on traditions and the im-
portance of dominating research fields, as well as regulatory factors. The treatment options
in RVVC are complex and include local and oral antifungals, as well as vaginal boric acid.
No matter the chosen treatment, a positive yeast culture is crucial prior to initiating any
plan. Yeast cultures help to establish the diagnosis, determine the species of the infecting
organism, and makes it available for susceptibility testing. Susceptibility testing, however,
does not correlate with clinical outcome, and the result of such testing depends on the pH
at which is the testing is performed.38
For women with RVVC, some form of maintenance therapy is the standard approach for
treating women with C. albicans infections. Standard treatment in culture-verified chronic
or recurrent vulvovaginal C. albicans is accepted in most countries. The regimen adopted by
many experts is oral fluconazole 150 mg every three days for three doses, followed by 150
mg weekly for six months. When on this regimen, 90% of women are well controlled43, and
this straightforward approach has been shown to improve quality of life in 96% of women.44
However, it is uncommonly curative and recurrence occurs frequently, with one study find-
ing that more than 63% of women who had completed maintenance therapy continued to
have ongoing infections.45
As discussed later, recent concerns about fluconazole use in women who are either attempting to
become or are pregnant limit its use in this population. Similarly, a small proportion of women may
be intolerant or allergic to fluconazole. Because fluconazole is a potent inhibitor of cytochrome
P450, there are a large number of possible drug interactions. Since even maintenance fluconazole
is a low dose, it is unclear whether these interactions are more than theoretical. As a s eparat e con-
cern, fluconazole may prolong the QT interval. However, QT effects seem to be dose related
and low relative to prevalence in the general population.46 In women where fluconazole is
ruled out as a course of treatment, options include clotrimazole 1% cream 5 grams nightly
for 14 days, followed by 5 grams vaginally twice a week for 6 months. One would expect that
similar topical maintenance regimens would also be successful at controlling RVVC. No mat-
ter the maintenance regimen, re-evaluating the patient after therapy is initiated, preferably
with repeated yeast cultures is very helpful to determine the response to treatment.
Ancillary measures can be considered to improve patient outcome. Treatment of the asymp-
tomatic sexual partner does not seem to be beneficial.39 Removal of intrauterine contracep-
tives should also be considered in women with chronic RVVC, as C. albicans may be more
likely to attach to it, possibly due to formation of biofilm on the intrauterine device.47 After
removal of the intrauterine device and treatment with fluconazole, affected women are more
likely to stay recurrence-free for a longer time period. In a small pilot study of administra-
tion of oral or depot formulations of medroxyprogesterone acetate (MPA) for RVVC, treated
women described a reduction of symptoms in the second year of therapy.48 For women who
have fewer than three annual episodes of VVC but are otherwise complicated (e.g. diabetics
with a severe infection), they should be treated in a manner similar to someone with RVVC.
Finally, it should be noted that azole-resistant C. albicans infections are being encountered
with increasing frequency by many tertiary care centers. Resistant infections are a clinical
104
definition, i.e. a women who remains symptomatic and culture positive while taking anti-
fungal treatment. Evidence for treatment options is relatively sparse.27 Since many non-al-
bicans Candida spp. are inherently azole resistant, providers should consider the treatment
options in the section on non-albicans Candida.
Non-
albicans Candida
vulvovaginitis
When a non-albicans Candida grows on a culture from a woman with vulvovaginal symp-
toms, providers should keep in mind that vulvovaginal symptoms are often non-specific
and that the patient ’s symptoms may be due to some other cause than the cultured organ-
ism (i.e. vulvodynia, lichen sclerosus). For example, it is estimated that 50% of women with
growth of non-albicans Candida are asymptomatic.49 Thus, it is recommended that treat-
ment only be offered to symptomatic women with no other identifiable cause. Treatment
with azoles is frequently unsuccessful in symptomatic C. glabrata; most of the options will
consist of compounded medications. Local administration of nystatin (100,000 units vagi-
nally nightly for a month), boric acid (600 mg capsules vaginally nightly for three weeks)50
or amphotericin B (50 mg suppositories vaginally nightly for two weeks)51 have been sug-
gested as reasonable options. In particularly refractory cases, a compounded cream with 1
gram flucytosine and 100 mg of amphotericin formulated in lubricating jelly base in a total
8 g delivered dose, inserted nightly for 14 days, was effective in two cases.52 Flucytosine is
both difficult to obtain and, in many countries, expensive.
C. krusei, C. tropicalis and S. cerevisiae are almost always inherently resistant to fluconazole.
Topical clotrimazole 100 mg daily for two weeks, nystatin53 or vaginal boric acid are the
treatments of choice in symptomatic women. Vaginal boric acid is suggested if azoles do
not work.29, 54 Providers should be aware that the European Chemicals Agency has issued a
warning against the use of boric acid, as it feels that it may impair fertility and might be em-
bryotoxic.55 On the other hand, the CDC do not mention any significant safety concerns with
boric acid.24 As a result of this controversy, boric acid is not readily available in many coun-
tries, despite the evidence that it may represent the best option for treating azole-resistant
yeast infections. Clearly, there remains a need for better treatment options for non-albicans
yeast and C. albicans infections resistant than azoles.
4.9
Special situations
Prepubertal children
It is assumed that the likelihood of term neonates to develop oral thrush or “diaper der-
matitis” during the first year of life is increased in those who are colonized through mater-
nal-to-neonatal transmission during vaginal delivery.56, 57 In these earlier papers, prophylac-
tic antimycotic treatment was suggested during the last weeks of pregnancy in women with
asymptomatic colonization to prevent transmission to the newborn during vaginal delivery.
However, given the absence of high-quality data and the difficulties involved with institut-
105
ing a huge screening program to identify women colonized with C. albicans, there are no
countries apart from Germany, to our knowledge, who have implemented such a program.11
After the initial postnatal period, due to reduced estrogenization of the vagina, premenstru-
al girls are much less likely to develop Candida spp. colonization and symptomatic VVC.14 Al-
though yeast cultures may be obtained in premenarchal girls with vulvovaginal complaints,
it should be expected that the vast majority will not have VVC and that antifungal treatment
will not be helpful.
Pregnancy
Pregnancy is a known risk factor for development of VVC, likely due to pregnancy-related
factors, including increased estrogen levels, increased vaginal glycogen, and alterations in
the immune system. Vaginal treatment with topical azoles, preferably clotrimazole, is rec-
ommended during pregnancy. We feel that it is reasonable to consider a longer course of
7-14 days, according to studies. Given the absence of safety data with regard to use in preg-
nancy, boric acid should not be used. Although there are few clinical studies on the use of
dequalinium chloride as an alternative treatment during pregnancy, available data suggest
good tolerability and effectiveness. Therefore, dequalinium chloride may be considered as a
therapeutic option for VVC during pregnancy in the countries where it is available.58-60
Oral fluconazole may be associated with malformations such as transposition of large ves-
sels and cleft palate and also with miscarriages. The US National Birth Defects Prevention
Study analyzed data from 43,257 women and found a significant association between low-
dose fluconazole use during the first trimester and incidence of fetal cleft lip and palate and
transposition of the large vessels61; similar results have been reported in a Danish registry
and a Canadian study.62, 63 Thus, oral fluconazole is not recommended in early pregnancy.
A more controversial question is whether treatment of VVC can improve pregnancy out-
come. Although some studies suggest that colonization with C. albicans is associated with
preterm birth and one prospective randomized study suggested that treatment with clo-
trimazole might decrease the risk, the data are insufficient to recommend screening or
treating for VVC in asymptomatic pregnant women.64, 65 Additional higher quality studies are
needed to further investigate a possible relationship between VVC and pregnancy outcome.
Postpartum and breastfeeding mothers
In general, women with symptomatic VVC who are postpartum or breastfeeding can be
treated similarly to other healthy women. Fluconazole is considered safe to use in breast-
feeding women.
Menopause
In general, because estrogen is thought to be the hormone associated with symptomatic
VVC, menopausal women are less likely to get VVC.66 Menopausal women who take meno-
pause hormone therapy seem to be more likely to get VVC than those who do not and are
more likely to recur.67
106
Immunosuppression
Women with immune deficiencies are more likely to be colonized by yeast and less likely to
clear infection. As such, consideration should be given to using regimens which are longer
in duration than those for uncomplicated infection and to recommend follow-up to make
sure that the treatment was effective.
4.10
Future perspectives
The current guidelines recommend fluconazole to suppress RVVC for most women, but
emerging resistance and therapeutic failure are creating a need for better treatment op-
tions. New medications may be on the horizon which may change current treatment algo-
rithms. Ibrexafungerp, an oral agent which affects the cell wall instead of the cell membrane
(the target of azoles) was recently approved as a single day two-dose regimen for acute VVC
in the US but is not available in Europe. Published data suggest that it is as effective as one
would expect with single dose fluconazole.68 Its role and the optimal dosing for managing
RVVC and resistant yeast infections remain to be elucidated. For RVVC, oteseconazole has
shown better results in reducing recurrences of C. albicans. In a double-blind placebo-con-
trolled randomized controlled trial of women with RVVC, patients treated with VT-1161 at
either a high or low dose for 12 or 24 weeks showed remarkably lower rates of recurrence
than those on placebo at the 48-week study time point (4 vs. 52%).69 Oteseconazole is soon
to be registered in Europe and was recently approved by the FDA in the US. However, the
FDA warning that it should not be used in women of reproductive potential because of the
combination of ocular abnormalities in the offspring of pregnant rats and a drug exposure
window of 690 days may sharply limit its use. Neither drug is approved for use in pregnancy.
As an alternative to antifungal medications, vaccination against C. albicans in a phase 2
study suggested that it was capable of reducing the frequency of symptomatic VVC for up
to 12 months, but only in a subset of women under 40 years of age.70
Beyond the ever present need for new therapeutics, there are many important research
questions which remain unanswered. Many of them relate to basic questions of pathophysi-
ology, such as the virulence or adherence factors of Candida spp. which play a role in causing
an infection, or the exact mechanisms which make certain women more prone to getting
VVC than others. Quick, easy and accurate methods of diagnosis, preferably available direct-
ly to patients, would greatly improve the care of affected women. Finally, understanding
the potential role of immune modulating medications and how they affect patient response
may shed further light on optimal patient management.
Despite these ongoing needs, it should be emphasized that applying our current state of knowl-
edge to women with VVC can frequently lead to excellent patient outcome and satisfaction.
107
Recommendations
Recommendation Quality of
evidence
Strength of
recommendation
For women with recurrent vulvovaginal candidiasis, evaluation and
treatment of an intestinal or partner source are not recommended. 1a A
Providers should classify candidiasis (complicated vs. uncomplicated,
and according to severity). 5 D
Asymptomatic colonization should not prompt treatment. 2a B
Self-diagnosis or empirical diagnosis are not recommended. 2b B
pH measurement, whiff test and wet mount microscopy are recom-
mended for evaluation of women with possible acute vulvovaginal
candidiasis.
3a B
Validated nucleic acid amplification tests can be used for the diagno-
sis of vulvovaginal candidiasis in symptomatic women. 1a A
Yeast culture remains the gold standard for confirming the diagnosis
of vulvovaginal candidiasis. 1a A
In women with uncomplicated vulvovaginal candidiasis, culture is
usually not necessary. 5 D
In women with suspected complicated vulvovaginal candidiasis,
culture is recommended. 5 D
In cases where drug resistance is suspected, susceptibility testing can
be considered. 5 D
Azole antifungals, including local clotrimazole, miconazole, or
econazole or oral single dose fluconazole, are the treatment of choice
in uncomplicated cases of vulvovaginal candidiasis.
1a A
A positive yeast culture is crucial prior to initiating suppressive treat-
ment for recurrent vulvovaginal candidiasis. 2b B
Susceptibility testing is not needed to start a suppressive treatment
for recurrent vulvovaginal candidiasis. 5 D
For recurrent vulvovaginal candidiasis, oral fluconazole 150 mg every
three days for three doses, followed by 150 mg weekly for six months
is the most commonly recommended regimen.
1a A
For recurrent vulvovaginal candidiasis, if fluconazole cannot be used,
clotrimazole 1% cream 5 grams nightly for 14 days, followed by 5
grams vaginally twice a week for 6 months can be considered.
1a A
Ancillary measures can be considered to improve patient outcome. 5 D
Removal of intrauterine contraceptives should be considered in
women with chronic or recurrent vulvovaginal candidiasis. 4 C
Medroxyprogesterone acetate can be considered in recurrent vulvo-
vaginal candidiasis. 5 D
Women who have fewer than three annual episodes of vulvovaginal
candidiasis but are otherwise complicated should be treated in a
manner similar to someone with recurrent vulvovaginal candidiasis.
1b A
108
In a symptomatic women in whom a non-albicans Candida is identi-
fied, it is recommended that treatment should only be offered if no
other causes are identified.
5 D
For Candida glabrata most of the options will consist of compounded
medications. 4 C
For Candida krusei, C. tropicalis and Saccharomyces cerevisiae topical
clotrimazole 100 mg daily for two weeks, nystatin or boric acid are the
treatments of choice in symptomatic women.
4 C
Vaginal boric acid may represent the best option for treating azole-re-
sistant yeast infections, despite safety concerns. 4 C
Vaginal treatment with longer courses of topical azoles, preferably
clotrimazole, is recommended during pregnancy. 1a A
Boric acid should not be used during pregnancy. 5 D
Oral fluconazole is not recommended in early pregnancy. 1a A
Women with symptomatic vulvovaginal candidiasis who are breast-
feeding can be treated similarly to other healthy women, including
with fluconazole.
1b A
Immunosuppressed women with vulvovaginal candidiasis should
be given longer regimens and followed-up to make sure that the
treatment was effective.
5 D
Ibrexafungerp and oteseconazole are new drugs that may change the
algorithms in the future, but are contra-indicated in pregnant women
or in those who may become pregnant in the near (ibrexafungerp) or
distant (oteseconazole) future.
1b A
References
1. Rolo, J.; Faria-Gonçalves, P.; Barata, T.; Oliveira, A. S.; Gaspar, C.; Ferreira, S. S.; Palmeira-de-Oliveira, R.; Martinez-de -Olivei-
ra, J.; Costa-de-Oliveira, S.; Palmeira-de-Oliveira, A., Species Distribution and Antifungal Susceptibility Profiles of Isolates
from Women with Nonrecurrent and Recurrent Vulvovaginal Candidiasis. Microb Drug Resist 2021, 27, (8), 1087-1095.
2. Sasani, E.; Rafat, Z.; Ashrafi, K.; Salimi, Y.; Zandi, M.; Soltani, S.; Hashemi, F.; Hashemi, S. J., Vulvovaginal candidiasis in
Iran: A systematic review and meta-analysis on the epidemiology, clinical manifestations, demographic character-
istics, risk factors, etiologic agents and laboratory diagnosis. Microb Pathog 2021, 154, 104802.
3. Sherrard, J.; Wilson, J.; Donders, G.; Mendling, W.; Jensen, J. S., 2018 European (IUSTI/WHO) International Union
against sexually transmitted infections (IUSTI) World Health Organisation (WHO) guideline on the management of
vaginal discharge. Int J STD AIDS 2018, 29, (13), 1258-1272.
4. Sobel, J. D., Genital candidiasis. Medicine 2014, 42, (7), 364-368.
5. Gharehbolagh, S. A.; Fallah, B.; Izadi, A.; Ardestani, Z. S.; Malekifar, P.; A, M. B.; Mahmoudi, S., Distribution, antifungal
susceptibility pattern and intra-Candida albicans species complex prevalence of Candida africana: A systematic
review and meta-analysis. PLoS One 2020, 15, (8), e0237046.
6. Borman, A. M.; Johnson, E. M., Name Changes for Fungi of Medical Importance, 2018 to 2019. J Clin Microbiol 2021, 59, (2).
7. Yano, J.; Peters, B. M.; Noverr, M. C.; Fidel, P. L., Jr., Novel Mechanism behind the Immunopathogenesis of Vulvovagi-
nal Candidiasis: “Neutrophil Anergy”. Infect Immun 2018, 86, (3).
8. De Bernardis, F.; Graziani, S.; Tirelli, F.; Antonopoulou, S., Candida vaginitis: virulence, host response and vaccine
prospects. Med Mycol 2018, 56, (suppl_1), 26-31.
109
9. Swidsinski, A.; Guschin, A.; Tang, Q.; Dörffel, Y.; Verstraelen, H.; Ter tychnyy, A.; Khayrullina, G.; Luo, X.; Sobel, J. D.; Jiang,
X., Vulvovaginal candidiasis: histologic lesions are primarily polymicrobial and invasive and do not containbiofilms.
Am J Obstet Gynecol 2019, 220, (1), 91.e1-91.e8.
10. Oliveira, A. S.; Rolo, J.; Gaspar, C.; Palmeira de Oliveira, R.; Martinez de Oliveira, J.; Palmeira de Oliveira, A., Allergic
vulvovaginitis: a systematic literature review. Arch Gynecol Obstet 2022, 306, (3), 593-622.
11. Farr, A.; Effendy, I.; Frey Tirri, B.; Hof, H.; Mayser, P.; Petricevic, L.; Ruhnke, M.; Schaller, M.; Schaefer, A. P. A.; Sustr, V.; Will-
inger, B.; Mendling, W., Guideline: Vulvovaginal candidosis (AWMF 015/072, level S2k). Mycoses 2021, 64, (6), 583-602.
12. Drell, T.; Lillsaar, T.; Tummeleht, L.; Simm, J.; Aaspõllu, A.; Väin, E.; Saarma, I.; Salumets, A.; Donders, G. G.; Metsis, M.,
Characterization of the vaginal micro- and mycobiome in asymptomatic reproductive-age Estonian women. PLoS
One 2013, 8, (1), e54379.
13. Foxman, B.; Muraglia, R.; Dietz, J. P.; Sobel, J. D.; Wagner, J., Prevalence of recurrent vulvovaginal candidiasis in 5 Euro-
pean countries and the United States: results from an internet panel survey. J Low Genit Tract Dis 2013, 17, (3), 340-5.
14. Dennerstein, G. J.; Ellis, D. H., Oestrogen, glycogen and vaginal candidiasis. Aust N Z J Obstet Gynaecol 2001, 41, (3), 326-8.
15. Benedict, K.; Lyman, M.; Jackson, B. R., Possible misdiagnosis, inappropriate empiric treatment, and opportunities for in-
creased diagnostic testing for patients with vulvovaginal candidiasis-United States, 2018. PLoS One 2022, 17, (4), e0267866.
16. Hurley, R., Candidal vaginitis. Proc R Soc Med 1977, 70 Suppl 4, (Suppl 4), 1-2.
17. Blostein, F.; Levin-Sparenberg, E.; Wagner, J.; Foxman, B., Recurrent vulvovaginal candidiasis. Ann Epidemiol 2017,
27, (9), 575-582.e3.
18. Denning, D. W.; Kneale, M.; Sobel, J. D.; R autemaa-Richardson, R., Global burden of recurrent vulvovaginal candidia-
sis: a systematic review. Lancet Infect Dis 2018, 18, (11), e339-e347.
19. Aballéa, S.; Guelfucci, F.; Wagner, J.; Khemiri, A.; Dietz, J. P.; Sobel, J.; Toumi, M., Subjective health status and health-re-
lated quality of life among women with Recurrent Vulvovaginal Candidosis (RVVC) in Europe and the USA. Health
Qual Life Outcomes 2013, 11, 169.
20. Zhu, Y. X.; Li, T.; Fan, S. R.; Liu, X. P.; Liang, Y. H.; Liu, P., Health-related quality of life as measured with the Short-Form 36
(SF-36) questionnaire in patients with recurrent vulvovaginal candidiasis. Health Qual Life Outcomes 2016, 14, 65.
21. Shukla, A.; Sobel, J. D., Vulvovaginitis Caused by Candida Species Following Antibiotic Exposure. Curr Infect Dis Rep
2019, 21, (11), 44.
22. Nyirjesy, P.; Sobel, J. D.; Fung, A.; Mayer, C.; Capuano, G.; Ways, K.; Usiskin, K., Genital mycotic infections with canag-
liflozin, a sodium glucose co-transporter 2 inhibitor, in patients with type 2 diabetes mellitus: a pooled analysis of
clinical studies. Curr Med Res Opin 2014, 30, (6), 1109-19.
23. Sobel, J. D.; Faro, S.; Force, R. W.; Foxman, B.; Ledger, W. J.; Nyirjesy, P. R.; Reed, B. D.; Summers, P. R., Vulvovaginal
candidiasis: epidemiologic, diagnostic, and therapeutic considerations. Am J Obstet Gynecol 1998, 178, (2), 203-11.
24. Workowski, K. A.; Bachmann, L. H.; Chan, P. A.; Johnston, C. M.; Muzny, C. A.; Park, I.; Reno, H.; Zenilman, J. M.; Bolan,
G. A., Sexually Transmitted Infections Treatment Guidelines, 2021. MMWR Recomm Rep 2021, 70, (4), 1-187.
25. Sobel, J. D.; Kapernick, P. S.; Zervos, M.; Reed, B. D.; Hooton, T.; Soper, D.; Nyirjesy, P.; Heine, M. W.; Willems, J.; Panzer, H.;
Wittes, H., Treatment of complicated Candida vaginitis: comparison of single and sequential doses of fluconazole.
Am J Obstet Gynecol 2001, 185, (2), 363-9.
26. Hong, E.; Dixit, S.; Fidel, P. L.; Bradford, J.; Fischer, G., Vulvovaginal candidiasis as a chronic disease: diagnostic criteria
and definition. J Low Genit Tract Dis 2014, 18, (1), 31-8.
27. Sobel, J. D.; Sobel, R., Current treatment options for vulvovaginal candidiasis caused by azole-resistant Candida
species. Expert Opin Pharmacother 2018, 19, (9), 971-977.
28. Pappas, P. G.; Kauffman, C. A.; Andes, D. R.; Clancy, C. J.; Marr, K. A.; Ostrosky-Zeichner, L.; Reboli, A. C.; Schuster, M.
G.; Vazquez, J. A.; Walsh, T. J.; Zaoutis, T. E.; Sobel, J. D., Clinical Practice Guideline for the Management of Candidiasis:
2016 Update by the Infectious Diseases Society of America. Clin Infect Dis 2016, 62, (4), e1-50.
29. Powell, A. M.; Gracely, E.; Nyirjesy, P., Non-albicans Candida Vulvovaginitis: Treatment Experience at a Tertiary Care
Vaginitis Center. J Low Genit Tract Dis 2016, 20, (1), 85-9.
30. Eckert, L. O.; Hawes, S. E.; Stevens, C. E.; Koutsky, L. A.; Eschenbach, D. A.; Holmes, K. K., Vulvovaginal candidiasis:
clinical manifestations, risk factors, management algorithm. Obstet Gynecol 1998, 92, (5), 757-65.
31. Anderson, M. R.; Klink, K.; Cohrssen, A., Evaluation of vaginal complaints. JAMA 2004, 291, (11), 1368-79.
32. Ferris, D. G.; Nyirjesy, P.; Sobel, J. D.; Soper, D.; Pavletic, A.; Litaker, M. S., Over-the-counter antifungal drug misuse
associated with patient-diagnosed vulvovaginal candidiasis. Obstet Gynecol 2002, 99, (3), 419-25.
33. Allen-Davis, J. T.; Beck, A.; Parker, R.; Ellis, J. L.; Polley, D., Assessment of vulvovaginal complaints: accuracy of tele-
phone triage and in-office diagnosis. Obstet Gynecol 2002, 99, (1), 18-22.
110
34. Schwebke, J. R.; Gaydos, C. A.; Nyirjesy, P.; Paradis, S.; Kodsi, S.; Cooper, C. K., Diagnostic Performance of a Molecular
Test versus Clinician Assessment of Vaginitis. J Clin Microbiol 2018, 56, (6).
35. Ledger, W. J.; Polaneczky, M. M.; Yih, M. C.; Jeremias, J.; Tolbert, V.; Witkin, S. S., Difficulties in the Diagnosis of Candida
Vaginitis. Infectious Diseases in Clinical Practice 2000, 9, (2), 66-69.
36. Donders, G. G.; Marconi, C.; Bellen, G.; Donders, F.; Michiels, T., Effect of short training on vaginal fluid microscopy
(wet mount) learning. J Low Genit Tract Dis 2015, 19, (2), 165-9.
37. Vieira-Baptista, P.; Silva, A. R.; Costa, M.; Aguiar, T.; Saldanha, C.; Sousa, C., Clinical validation of a new molecular test (Seegene
Allplex™ Vaginitis) for the diagnosis of vaginitis: a cross-sectional study. Bjog 2021, 128, (8), 1344-1352.
38. Danby, C. S.; Boikov, D.; Rautemaa-Richardson, R.; Sobel, J. D., Effect of pH on in vitro susceptibility of Candida glabrata and
Candida albicans to 11 antifungal agents and implications for clinical use. Antimicrob Agents Chemother 2012, 56, (3), 1403-6.
39. Sobel, J. D., Vulvovaginal candidosis. Lancet 2007, 369, (9577), 1961-71.
40. Pitsouni, E.; Iavazzo, C.; Falagas, M. E., Itraconazole vs fluconazole for the treatment of uncomplicated acute vaginal
and vulvovaginal candidiasis in nonpregnant women: a metaanalysis of randomized controlled trials. Am J Obstet
Gynecol 2008, 198, (2), 153-60.
41. Mendling, W.; Krauss, C.; Fladung, B., A clinical multicenter study comparing efficacy and tolerability of topical com-
bination therapy with clotrimazole (Canesten, two formats) with oral single dose fluconazole (Diflucan) in vulvo-
vaginal mycoses. Mycoses 2004, 47, (3-4), 136-42.
42. Nurbhai, M.; Grimshaw, J.; Watson, M.; Bond, C.; Mollison, J.; Ludbrook, A., Oral versus intra-vaginal imidazole and triazole
anti-fungal treatment of uncomplicated vulvovaginal candidiasis (thrush). Cochrane Database Syst Rev 2007, (4), Cd002845.
43. Sobel, J. D.; Wiesenfeld, H. C.; Martens, M.; Danna, P.; Hooton, T. M.; Rompalo, A.; Sperling, M.; Livengood, C., 3rd;
Horowitz, B.; Von Thron, J.; Edwards, L.; Panzer, H.; Chu, T. C., Maintenance fluconazole therapy for recurrent vulvo-
vaginal candidiasis. N Engl J Med 2004, 351, (9), 876-83.
44. Nguyen, Y.; Lee, A.; Fischer, G., Quality of life in patients with chronic vulvovaginal candidiasis: A before and after
study on the impact of oral fluconazole therapy. Australas J Dermatol 2017, 58, (4), e176-e181.
45. Crouss, T.; Sobel, J. D.; Smith, K.; Nyirjesy, P., Long-Term Outcomes of Women With Recurrent Vulvovaginal Candidia-
sis After a Course of Maintenance Antifungal Therapy. J Low Genit Tract Dis 2018, 22, (4), 382-386.
46. Berger, F. A.; Monadian, N.; de Groot, N. M. S.; Santbergen, B.; van der Sijs, H.; Becker, M. L.; Broers, A. E. C.; van Gelder,
T.; van den Bemt, P., QTc prolongation during ciprofloxacin and fluconazole combination therapy: prevalence and
associated risk factors. Br J Clin Pharmacol 2018, 84, (2), 369-378.
47. Cakiroglu, Y.; Caliskan, S.; Doger, E.; Ozcan, S.; Caliskan, E., Does removal of CU-IUD in patients with biofilm forming
candida really maintain regression of clinical symptoms? J Obstet Gynaecol 2015, 35, (6), 600-3.
48. Špaček, J.; Kestřánek, J.; Jílek, P.; Leško, D.; Plucnarová, S.; Buchta, V., Comparison of two long-term gestagen reg-
imens in the management of recurrent vulvovaginal candidiasis: A pilot study. Mycoses 2017, 60, (4), 260-265.
49. Kennedy, M. A.; Sobel, J. D., Vulvovaginal Candidiasis Caused by Non-albicans Candida Species: New Insights. Curr
Infect Dis Rep 2010, 12, (6), 465-70.
50. Sobel, J. D.; Chaim, W.; Nagappan, V.; Leaman, D., Treatment of vaginitis caused by Candida glabrata: use of topical
boric acid and flucytosine. Am J Obstet Gynecol 2003, 189, (5), 1297-300.
51. Phillips, A. J., Treatment of non-albicans Candida vaginitis with amphotericin B vaginal suppositories. Am J Obstet
Gynecol 2005, 192, (6), 2009-12; discussion 2012-3.
52. White, D. J.; Habib, A. R.; Vanthuyne, A.; Langford, S.; Symonds, M., Combined topical flucytosine and amphotericin
B for refractory vaginal Candida glabrata infections. Sex Transm Infect 2001, 77, (3), 212-3.
53. Singh, S.; Sobel, J. D.; Bhargava, P.; Boikov, D.; Vazquez, J. A., Vaginitis due to Candida krusei: epidemiology, clinical
aspects, and therapy. Clin Infect Dis 2002, 35, (9), 1066-70.
54. Sobel, J. D.; Suprapaneni, S., Candida parapsilosis Vaginal Infection-a New Site of Azole Drug Resistance. Curr Infect
Dis Rep 2018, 20, (11), 43.
55. Felix, T. C.; de Brito Röder, D. V. D.; Dos Santos Pedroso, R., Alternative and complementary therapies for vulvovaginal
candidiasis. Folia Microbiol (Praha) 2019, 64, (2), 133-141.
56. Blaschke-Hellmessen, R., [Epidemiological studies of the occurrence of yeasts in children and their mothers].
Mykosen 1968, 11, (9), 611-6.
57. Schnell, J. D., Epidemiology and the prevention of peripartal mycoses. Chemotherapy 1982, 28 Suppl 1, 66-72.
58. Mendling, W.; Weissenbacher, E. R.; Gerber, S.; Prasauskas, V.; Grob, P., Use of locally delivered dequalinium chloride
in the treatment of vaginal infections: a review. Arch Gynecol Obstet 2016, 293, (3), 469-84.
59. Frey Tirri, B., Antimicrobial topical agents used in the vagina. Curr Probl Dermatol 2011, 40, 36-47.
60. Della Casa, V.; Noll, H.; Gonser, S.; Grob, P.; Graf, F.; Pohlig, G., Antimicrobial activity of dequalinium chloride against
leading germs of vaginal infections. Arzneimittelforschung 2002, 52, (9), 699-705.
111
61. Howley, M. M.; Carter, T. C.; Browne, M. L.; Romitti, P. A.; Cunniff, C. M.; Druschel, C. M., Fluconazole use and birth
defects in the National Birth Defects Prevention Study. Am J Obstet Gynecol 2016, 214, (5), 657.e1-9.
62. Mølgaard-Nielsen, D.; Svanström, H.; Melbye, M.; Hviid, A.; Pasternak, B., Association Between Use of Oral Flucona-
zole During Pregnancy and Risk of Spontaneous Abortion and Stillbirth. JAMA 2016, 315, (1), 58-67.
63. Bérard, A.; Sheehy, O.; Zhao, J. P.; Gorgui, J.; Bernatsky, S.; de Moura, C. S.; Abrahamowicz, M., Associations between low-
and high-dose oral fluconazole and pregnancy outcomes: 3 nested case-control studies. Cmaj 2019, 191, (7), E179-e187.
64. Farr, A.; Kiss, H.; Holzer, I.; Husslein, P.; Hagmann, M.; Petricevic, L., Effect of asymptomatic vaginal colonization with
Candida albicans on pregnancy outcome. Acta Obstet Gynecol Scand 2015, 94, (9), 989-96.
65. Holzer, I.; Farr, A.; Kiss, H.; Hagmann, M.; Petricevic, L., The colonization with Candida species is more harmful in the
second trimester of pregnancy. Arch Gynecol Obstet 2017, 295, (4), 891-895.
66. Nyirjesy, P.; Leigh, R. D.; Mathew, L.; Lev-Sagie, A.; Culhane, J. F., Chronic vulvovaginitis in women older than 50 years:
analysis of a prospective database. J Low Genit Tract Dis 2012, 16, (1), 24-9.
67. Fischer, G.; Bradford, J., Vulvovaginal candidiasis in postmenopausal women: the role of hormone replacement
therapy. J Low Genit Tract Dis 2011, 15, (4), 263-7.
68. Schwebke, J. R.; Sobel, R.; Gersten, J. K.; Sussman, S. A.; Lederman, S. N.; Jacobs, M. A.; Chappell, B. T.; Weinstein, D.
L.; Moffett, A. H.; Azie, N. E.; Angulo, D. A.; Harriott, I. A.; Borroto-Esoda, K.; Ghannoum, M. A.; Nyirjesy, P.; Sobel, J. D.,
Ibrexafungerp Versus Placebo for Vulvovaginal Candidiasis Treatment: A Phase 3, Randomized, Controlled Superior-
ity Trial (VANISH 303). Clin Infect Dis 2022, 74, (11), 1979-1985.
69. Brand, S. R.; Degenhardt, T. P.; Person, K.; Sobel, J. D.; Nyirjesy, P.; Schotzinger, R. J.; Tavakkol, A., A phase 2, randomized,
double-blind, placebo-controlled, dose-ranging study to evaluate theefficacy and safety of orally administeredVT-
1161 inthe treatment of recurrentvulvovaginal candidiasis. Am J Obstet Gynecol 2018, 218, (6), 624.e1-624.e9.
70. Edwards, J. E., Jr.; Schwartz, M. M.; Schmidt, C. S.; Sobel, J. D.; Nyirjesy, P.; Schodel, F.; Marchus, E.; Lizakowski, M.;
DeMontigny, E. A.; Hoeg, J.; Holmberg, T.; Cooke, M. T.; Hoover, K.; Edwards, L.; Jacobs, M.; Sussman, S.; Augenbraun,
M.; Drusano, M.; Yeaman, M. R.; Ibrahim, A. S.; Filler, S. G.; Hennessey, J. P., Jr., A Fungal Immunotherapeutic Vaccine
(NDV-3A) for Treatment of Recurrent Vulvovaginal Candidiasis-A Phase 2 Randomized, Double-Blind, Placebo-Con-
trolled Trial. Clin Infect Dis 2018, 66, (12), 1928-1936.
112
113
5.1
Introduction
Trichomonas vaginalis is estimated to be the most common curable non-viral sexually trans-
mitted infection (STI) worldwide.1 It is associated with multiple adverse health outcomes
including adverse birth outcomes,2 increased risk of acquisition and transmission of human
immunodeficiency virus (HIV) and other STIs,3-5 pelvic inflammatory disease (PID),6, 7 infertil-
ity,8, 9 and cervical cancer.10 A predominant health disparity exists for T. vaginalis infection, as
African Americans are significantly more likely to be infected than persons of other races.11
Beyond screening recommendations at entry to care and annually thereafter for HIV-infect-
ed women,12 there are no established screening, surveillance, or control programs for this
infection. Due to this limited public health response, T. vaginalis is frequently considered a
“neglected” STI.13 This chapter reviews the etiology, pathophysiology, epidemiology, clinical
manifestations, diagnosis, and treatment recommendations for this common infection.
5.2
Etiology and pathophysiology
Trichomoniasis is caused by the parasitic pathogen, T. vaginalis, which primarily infects the
squamous epithelium of the genital tract and causes damage to host epithelial cells. (Figure 5.1)
5
TRICHOMONIASIS
(alphabetical order)
Patricia Kissinger
Christina Muzny
Colleen K. Stockdale
114
Figure 5.1 A and B– Trichomonads seen with Gram stain (1000x, oil immersion)
It typically infects the female lower genital tract (vagina, urethra, and endocervix) and the
male urethra and prostate. It is transmitted among humans, its only known natural host.14, 15
While transmission by fomites has been occasionally reported,16-19 transmission occurs pri-
marily by sexual contact.20 Based on in vitro studies, its incubation period is 4-28 days.21 T.
vaginalis does not exist in a cyst form and does not survive well in the environment, but has
been identified outside the human body in warm and wet locations (i.e. moist towels) for >3
hours.16 It has its own microbiota, harboring two Mycoplasma species and a double-strand-
ed RNA virus, T. vaginalis virus (TVV), which can contribute to its pathogenesis.22, 23 Of the
four known TVV viruses, TVV1 and TVV2 have been linked to genital symptom severity24 and
TVV2 and TVV3 to surface expression of an immunogenic protein P270 (associated with cy-
totoxicity, cytoadherence, and host immune evasion);25 the role of TVV4 is not yet elucidat-
ed. However, in a study of 355 US T. vaginalis isolates from women participating in a clinical
trial, of which 40% were positive for TVV, there was no association between TVV positivity
and genital symptoms, repeated infections, or metronidazole resistance, suggesting that
TVV may be commensal to T. vaginalis.26
T. vaginalis infection is more common in women than in men, which may be due to the
anatomy of the female genital tract.27-31 Other possibilities could be due to spontaneous
resolution in men (which may occur in 36-69% of cases)32, 33 or less effective testing among
asymptomatic men.11, 29, 34, 35
Notably, T. vaginalis may persist for long periods of time in women (months or years)36 while
persistence in men is typically shorter (less than a month in some cases).32 The greater like-
lihood of persistence in women has been linked to the greater availability of iron, an es-
sential nutrient for the parasite.15, 31, 37-40 Additionally, menstrual blood creates a rich growth
medium which, when combined with the high concentration of iron during menstruation,
promotes vaginal attachment and parasite growth.27, 28
115
5.3
Prevalence and epidemiology
While not a reportable STI, global estimates indicate that, among women and men, there
are 156 million new cases per year.1 In addition, the global prevalence of T. vaginalis among
women (5.3%) is higher than that of chlamydia (3.8%), gonorrhea (0.9%), and syphilis (0.5%)
combined.1 In the US, the prevalence of T. vaginalis by urine nucleic acid amplification test-
ing (NAAT) in a recent population-based study was 1.8% in women and 0.5% in men.11
African Americans had a 4-fold higher prevalence than other racial groups, constituting a
dramatic health disparity.11 Unlike many STIs, T. vaginalis prevalence can be higher among
older persons with rates ranging from 0.2-21.4% among persons >45 years of age.41 In addi-
tion, population-based studies have found that T. vaginalis rates are highest among those 25
years or older.42 The prevalence of urethral T. vaginalis in men who have sex with men (MSM)
is extremely low to non-existent.43 Although extra-genital (oral, rectal) T. vaginalis occasion-
ally occurs, it is much less common than genital infection;44, 45 testing is not recommended.
5.4
Risk factors
Table 5.1 lists common risk factors for T. vaginalis infection. While present in all races,11 infec-
tion is more common in African American women engaging in high-risk sexual behaviors in-
cluding having multiple sexual partners,46, 47 inconsistent condom use, illicit drug use during
TABLE 5.1 Risk factors for T. vaginalis infection.
BV – bacterial vaginosis, HIV - human immunodeficiency virus
Female sex
African American race
Multiple sexual partners
Early coitarche
Older age
Inconsistent condom use
Illicit drug use
Sex with partners using illicit drugs
Transactional sex
History of incarceration
Having less than a high school education
Living below the national poverty level
Concurrent BV
Concurrent HIV
116
sex,46 sex with partners using illicit drugs,29, 46, 48, 49 and transactional sex.50-53 Other risk factors
include early coitarche,54 older age,29, 48, 55 history of incarceration,56, 57 having less than a high
school education,11 and living below the national poverty level.12, 47
Women with bacterial vaginosis (BV) are at higher risk for acquiring T. vaginalis.12, 20, 58 While
vaginal dysbiosis has been associated with increased pathogenicity of T. vaginalis,59 it is not
clear if the presence of BV interferes with T. vaginalis treatment. In two prior randomized
controlled trials (RCTs), BV was found to increase metronidazole treatment failure in HIV-in-
fected women60 but not in HIV-uninfected women.61 This difference may be due to impaired
immunity in HIV-infected women,60 altered pharmacokinetics and pharmacodynamics of
metronidazole,62 or inadequate power in the studies conducted.61 Additionally, women with
HIV are at higher risk for T. vaginalis.50, 63 Several studies have also shown that women who
have sex with women and men are at higher risk for T. vaginalis than women who have sex
with women and women who have sex with men.29, 64
5.5
Complications
Table 5.2 lists major complications associated with T. vaginalis infection among women,
which are further detailed below.
Adverse birth outcomes
In a meta-analysis of 19 peer-reviewed studies, significant associations were found between
T. vaginalis and preterm delivery (odds ratio [OR] 1.27; 95% CI 1.08-1.50), pre-labor rupture
of membranes (OR 1.87; 95% CI 1.53-2.29), and low birth weight (OR 2.12; 95% CI 1.15-
3.91).2 The physiological mechanisms linking trichomoniasis and adverse birth outcomes
are not well understood. One hypothesis is that preterm delivery and premature rupture of
membranes in T. vaginalis-infected pregnant women are related to maternal innate immune
inflammatory responses to the parasite, which involve elevated cervical interleukin-8 (IL-8)
and vaginal defensin levels.8, 65 These cytokines are markers of neutrophil activation, which
has been associated with preterm delivery and other adverse birth events. For example,
cervical IL-8 is thought to trigger cervical ripening and dilatation.66, 67 Additionally, one study
has shown an association between maternal T. vaginalis infection and intellectual disability
in children born to infected mothers.68
117
TABLE 5.2 Complications of T. vaginalis infection in women.
CI - confidence interval, HIV - human immunodeficiency virus, OR - odds ratio, PID - pelvic
inflammatory disease, RCT- randomized controlled trial, RR - relative risk,
STI - sexually transmitted infection
Outcome Author (Year) Study Design Study Findings
Adverse birth
outcomes
Van Gerwen et al,
2021 2Meta-Analysis
Among 19 studies, significant associations were
found between T. vaginalis and preterm delivery
(OR 1.27; 95% CI 1.08-1.50), prelabor rupture of
membranes (OR 1.87; 95% CI 1.53-2.29) and low
birthweight (OR 2.12; 95% CI 1.15-3.91).
HIV acquisition
Masha et al, 2018 69
Barker et al, 2022 70
Meta-Analysis
Meta-Analysis
Among 19 studies, T. vaginalis-infected individ-
uals were 1.5 times more likely to acquire HIV
than non-infected individuals (95% CI 1.3-1.7;
p<0.001).
Of 32 studies reporting k=97 effect size esti-
mates of HIV acquisition risk due to non-viral STI
infections, HIV acquisition risk was statistically
significant for T. vaginalis-infected women (RR
1.54; 95% CI 1.31-1.82; k = 17).
Pelvic inflam-
matory
disease
Moodley et al,
2002 6
Wiringa et al, 2020 7
Cross-sectional
study
Secondary analysis
of RCT data
T. vaginalis was associated with PID in HIV-infected
women but not HIV-uninfected women (p=0.002).
The odds of endometritis at baseline were twice
as high among T. vaginalis-infected women
compared to uninfected women (adjusted OR
1.9, 95% CI 1.0-3.3).
Infertility Zhang et al, 2022 77 Meta-Analysis
Among 8 studies, T. vaginalis was associated with
a 1.7 times greater risk of infertility in women
(95% CI 1.25-2.31).
Cervical cancer Yang et al, 2018 10 Meta-Analysis
Among 17 studies, the odds of cervical cancer
for T. vaginalis-infected women was 2.06 (95% CI
1.77-2.39).
Risk of HIV
A meta-analysis of 19 peer-reviewed studies found that persons infected with T. vagina-
lis were 1.5 times more likely to acquire HIV than non-infected individuals (95% CI 1.3-1.7;
p<0.001).69 In another meta-analysis of 32 peer-reviewed studies reporting k=97 effect size
estimates of HIV acquisition risk due to non-viral STI infections among high-risk heterosexuals
dia gnosed with chlamydia, gonorrhea, syphilis, Mycoplasma genitalium, and/or T. vaginalis,
HIV acquisition risk was statistically significant for T. vaginalis-infected women (relative risk
[RR] 1.54; 95% CI 1.31-1.82; k=17).70 The greater susceptibility to HIV among T. vaginalis-in-
fected individuals is biologically plausible for several reasons: (1) T. vaginalis damages epi-
thelial cell membranes which act as a structural barrier to HIV, (2) the host immune response
to T. vaginalis stimulates an increased number of HIV target cells in the genital tract mucosa,
and (3) T. vaginalis alters the normal vaginal microbiota, rendering it more permissive to the
development of BV, which, in turn, increases HIV acquisition risk.69
118
There is less direct evidence suggesting that HIV-infected individuals with T. vaginalis are
more likely to transmit HIV. A review paper found that only seven of 14 studies demonstrat-
ed a higher likelihood of HIV shedding in the genital fluids of T. vaginalis coinfected individ-
uals compared to HIV-infected individuals without coinfection.71 In other studies, vaginal
shedding of HIV-1 RNA was decreased after T. vaginalis treatment in a cohort of women from
Kenya72 and New Orleans, LA.73 However, in the Kenyan cohort, the prevalence of vaginal
HIV-1 DNA remained unchanged despite T. vaginalis treatment.
Risk of other sexually transmitted infections
Concomitant infection with T. vaginalis has been associated with a higher incidence of gen-
ital herpes simplex virus (HSV) 2 infection74 as well as genital HSV2 shedding.75 It has also
been associated with the presence of other STIs including chlamydia, gonorrhea, and hu-
man papillomavirus (HPV).4, 76
Pelvic inflammatory disease
T. vaginalis is not traditionally considered STI associated with PID. However, in a 2002 study
of 119 South African women, those infected with T. vaginalis had a significantly higher risk of
PID than those without (p=0.03).6 When women were stratified according to their HIV status,
the risk of PID in HIV-infected women with T. vaginalis increased significantly (p=0.002); no
association was found in women without HIV.6 More recently, among 647 women in the PID
Evaluation and Clinical Health (PEACH) study, T. vaginalis was frequently isolated from the
vagina in 12.8% and the odds of having endometritis at baseline was twice as high among
women with T. vaginalis compared to those without (adjusted OR 1.9; 95% CI 1.0-3.3). Infer-
tility and recurrent PID were also more common among women with T. vaginalis.7
Infertility
A meta-analysis of eight peer-reviewed studies found that T. vaginalis was associated with a
1.7 times greater risk of infertility in women (95% CI 1.25-2.31).77 Similarly, a meta-analysis
of five peer-reviewed studies found that T. vaginalis was associated with a 1.91 times great-
er risk of infertility in men (95% CI 1.02-3.58).77 This is thought to be due to inflammatory
damage of female reproductive organs and changes in the vaginal environment which may
result in a decrease or loss of reproductive function in women.77 In men, T. vaginalis itself or
the induced inflammatory response can impair sperm cells, causing a decrease in cell viabil-
ity or death, which may result in a decrease or loss of reproductive function.77
Risk of cervical cancer
A study of Tanzanian women found that those infected with T. vaginalis were 6.5 times more
likely to have high-risk HPV, suggesting an indirect link between T. vaginalis and cervical
neoplasia.4 In addition, a meta-analysis of 17 peer-reviewed studies found that T. vagina-
lis-infected women had a higher risk of cervical neoplasia (OR 2.06, 95% CI 1.77-2.39), with
HPV co-infection playing a central role.10
119
5.6
Signs and symptoms
The “classic” symptoms of T. vaginalis infection
include vaginal odor and a yellow-green, frothy,
malodorous vaginal discharge.12, 27, 48, 78, 79 How-
ever, a large number of infected women have
minimal or no symptoms, with only 11-17%
presenting with typical symptoms.80 Half of
asymptomatic women infected with T. vag-
inalis may become symptomatic within six
months.14 Infected women can also develop cy-
clic symptoms that are worse during menstru-
ation.27 Symptomatic women with T. vaginalis
may note a wide range of additional symptoms
including genital pruritus, dysuria, and dys-
pareunia.
On exam, signs may include vaginal enanthema,
malodorous, frothy, vaginal discharge ranging from clear to yellow-green, colpitis macularis or
“strawberry cervix” (present in <5% of women;81, 82 rises to nearly 50% with colposcopy82), and
elevated vaginal pH >4.5.20 (Figure 5.2)
Infection may also be present in the setting of a normal vaginal pH.20
5.7
Diagnosis
T. vaginalis has been traditionally diagnosed at the point-of-care (POC) by wet mount mi-
croscopy (WMM) of vaginal discharge for motile trichomonads (sensitivity 44-68%; specific-
ity 100%).83 (Figure 5.3)
Figure 5.3 Wet mount microscopy.
A– Several trichomonads and inflammation (https://www.youtube.com/watch?v=pTL-_Q4S1Og) (200x)
B– Trichomonad with its typical structures: flagella on the outside and hydrogenosomes on the inside (400x)
C– Several trichomonads, inflammation and bacterial vaginosis (400x, phase contrast)
Figure 5.2 Strawberry cervix
120
Ideally, this test must be performed within 10-20 minutes after collection or the trichomon-
ads will lose their motility, increasing the likelihood of a false negative test. The OSOM® rap-
id test (Sekisui Diagnostics, California) is another POC test (results ≤10 minutes) that uses
antibodies to detect T. vaginalis protein antigens in vaginal discharge (sensitivity 82-95%
and specificity 97-100%, compared to WMM and culture). It is a qualitative test that should
primarily be used in symptomatic women or contacts to T. vaginalis.83 When present, T. vag-
inalis antigens bind the antibodies resulting in the formation of a blue line on the test strip.
This test does not require microscopy, however is more expensive than WMM.
Trichomonas culture (InPouch® system [BioMed Diagnostics, White City, OR]) has previously
been the gold standard for diagnosis (sensitivity 44-96%; specificity 100%).83-85 Specimens
from women (vaginal swabs) or men (urethral swabs, urine sediment, and/or semen; mul-
tiple specimens recommended to increase yield) should be used to inoculate the culture
medium <1 hour after collection.83 This test is categorized by the Clinical Laboratory Im-
provement Amendments (CLIA) as moderately complex, as it requires incubation at 37°C
and reading over multiple days.86
Over the past decade, the availability of highly sensitive and specific T. vaginalis molecular
diagnostic assays has grown rapidly. These assays can be sub-divided into molecular ampli-
fied assays (i.e. AmpliVue™ and Solana® assays),87, 88 instrument-based assays (i.e. Hologic
Aptima® T. vaginalis NAAT, Becton Dickinson [BD] ProbeTec™ Qx T. vaginalis NAAT, BD Max™
CT/GC/TV2 NAAT, Cepheid GeneXpert® T. vaginalis NAAT, Roche Cobas® MG/TV NAAT, and
the Abbott Alinity m STI assay [including T. vaginalis NAAT testing],89-93 and instrument-free
assays (i.e. Visby GC/CT/TV NAAT testing device).94 These assays, with their respective sen-
sitivities and specificities, specimen types in women, complexity, and time to results, are
detailed in Table 5.3. Several of these molecular assays can provide testing results within one
hour or less (i.e. AmpliVue™ assay [results in 45-50 minutes]; Solana® assay [results in <40
minutes]; Cepheid GeneXpert® T. vaginalis NAAT [results in 40-63 minutes], and Visby GC/CT/
TV NAAT testing device [results in 25 minutes]).
TABLE 5.3 T. vaginalis diagnostic tests in women.
CLIA - Clinical Laboratory Improvement Amendments, CT - Chlamydia trachomatis, GC - Neis-
seria gonorrhoeae, MG - Mycoplasma genitalium, NAAT - nucleic acid amplification test, POC
- point-of-care. STI - sexually transmitted infection, TV - Trichomonas vaginalis.
*FDA-approved 5/4/22; https://www.molecular.abbott/int/en/products/infectious-disease/
alinity-m-sti-assay
Test Sample Sensitivity/Specificity
for T. vaginalis Complexity/Time
Wet mount micros-
copy83 Vaginal specimens Sensitivity: 44-68%;
Specificity: 100%
CLIA waived; POC test (re-
sults in ≤10 minutes).
OSOM®83
Vaginal specimens
(Most useful in symptomatic
women)
Sensitivity: 83-92%;
Specificity: 99-100%
CLIA waived; POC test (re-
sults in ≤ 10 minutes).
121
BD Affirm™ VPIII83 Vaginal specimens Sensitivity: 91-100%;
Specificity: 93-96%
Moderate complexity.
Results <1 hour.
Culture83, 85 Vaginal specimens Sensitivity: 44-81%;
Specificity: 100%
Moderate complexity. Re-
quires incubation at 37°C;
should be read for 5 days
over a 7 day period. 86
AmpliVue™ 87
Vaginal specimens from
symptomatic and asymp-
tomatic women
Sensitivity 90.7%;
Specificity 98.9% Results in 45-50 minutes.
Solana® 88
Vaginal specimens from
symptomatic and asymp-
tomatic women; urine
specimens
Sensitivity/specificity
98.6%-100%/98.5%-
98.9% for vaginal
specimens and 92.9%-
98%/97.9%-98.4% for
urine specimens
Results in <40 minutes.
Hologic Aptima® T.
vaginalis NAAT89
Vaginal, endocervical, Thin-
Prep®, and urine specimens
from symptomatic and
asymptomatic women
Sensitivity: 95.2%-100%;
Specificity: 98.9%-99.6%
High complexity. Re-
quires Panther, Viper, or
Tigris system. Results in
<8 hours.
BD ProbeTec™ Qx T.
vaginalis NAAT90
Vaginal, endocervical, and
urine specimens from symp-
tomatic and asymptomatic
women
Sensitivity: 98%-100%;
Specificity: 98%-100%
High complexity.
Requires Viper system.
Results in <8 hours.
BD Max™ CT/GC/TV2
NAAT93
Vaginal, endocervical, and
urine specimens from symp-
tomatic and asymptomatic
women
Sensitivity: 86.6%-100%;
Specificity: 99.2%-99.8% High complexity.
Cepheid GeneXpert® 91
Self-collected vaginal, clini-
cian-collected endocervical,
and urine specimens from
symptomatic and asymp-
tomatic women
Sensitivity: 99.5%-100%;
Specificity: 99.4-99.9%
Moderate complexity.
Results in 40-63 minutes.
Roche Cobas® MG/TV
NAAT 92
Vaginal and endocervical
specimens from symptomat-
ic and asymptomatic women
Sensitivity: 96.4%-100%;
Specificity: 96.5%-98.8%
High complexity. For
use on Cobas 6800/8800
systems.
Abbott Alinity m STI
assay*
Vaginal, endocervical, Thin-
Prep® and urine specimens
from symptomatic and
asymptomatic women
Sensitivity, specificity
not yet published; refer
to Abbott Molecular
website*
Results in <115 minutes.
Visby GC/CT/TV NAAT
Testing Device94
Self-collected vaginal
specimens
Sensitivity: 99.2%;
Specificity 96.9%
CLIA waived; POC test
(results in 25 minutes).
122
5.8
Treatment and follow-up
The primary drug class used to treat T. vaginalis is the 5-nitroimidazoles (metronidazole, tini-
dazole, and secnidazole). For decades, the Centers for Disease Control and Prevention (CDC)
and World Health Organization have recommended single dose 2 g oral metronidazole as
the preferred treatment for T. vaginalis, with oral metronidazole 400-500 mg twice daily for
seven days or single-dose 2 g oral tinidazole as alternative therapies. The recommended
treatment was changed to the seven-day oral metronidazole dose for HIV-infected women
over a decade ago in response to a multi-center RCT demonstrating superiority of the sev-
en day oral metronidazole dose over singe-dose.60 A subsequent meta-analysis95 and mul-
ti-center RCT61 found similar results in HIV-uninfected women. In vivo pharmacokinetic and
pharmacodynamic effects of metronidazole may be playing a role in treatment failure with
the use of the single 2 g oral dose, necessitating a longer treatment regimen in women.62
Two hypotheses for this finding are: (1) competition for oral metronidazole by BV-associ-
ated bacteria in the vaginal microbiota of T. vaginalis-infected women and (2) inadequate
accumulation of the active metabolites of metronidazole when only a single oral dose is
given.62 Thus, the seven day oral metronidazole regimen has since become the recommend-
ed treatment regimen for all women with single dose 2 g oral tinidazole remaining as an
alternative;12, 96 single dose 2 g oral metronidazole is no longer recommended in women.
Given the lack of a comparable RCT in men, single dose 2 g oral metronidazole remains the
recommended therapy for men with single dose 2 g oral tinidazole as an alternate until
additional studies are conducted.12
If a woman is still infected with T. vaginalis after multi-dose oral metronidazole and has been
re-exposed to an untreated sexual partner, she should be re-dosed with the same seven day
treatment regimen. If she has not been re-exposed, she should be re-treated with either 2
g of oral metronidazole or tinidazole daily for seven days.12 If a male is still infected with T.
vaginalis after treatment with single dose 2 g oral metronidazole and has been re-exposed
to an untreated sexual partner, he should be re-dosed with another single dose 2 g oral met-
ronidazole. If he has not been re-exposed, he should be given a course of oral metronidazole
500 mg twice daily for seven days.12
Most recently, a randomized, double-blind, placebo controlled, delayed-treatment study
evaluating the efficacy and safety of a single 2 g dose of oral secnidazole, a second genera-
tion 5-nitroimidazole with a longer half-life (17-19 hours), in 147 women with trichomoniasis
was conducted.97 At the test-of-cure visit 6−12 days after randomization, the microbiologic
cure rate was 92.2% (95% CI 82.7-97.4) in the secnidazole group and 1.5% (95% CI 0.0-8.0) in
the placebo group (p<0.001).97 For women who received placebo at baseline, the opposite
treatment was given at test-of-cure to ensure all participants were treated per standard of
care. Overall, secnidazole was well tolerated. The most frequent adverse events were vul-
vovaginal candidiasis and nausea (2.7% each); no serious adverse events were observed.
Secnidazole has since been FDA approved for T. vaginalis treatment in adolescent and adult
women and men aged ≥12 years. It is also FDA approved for BV treatment in women12 and
is the only oral single-dose treatment currently available for both BV and trichomoniasis.98
123
Re-testing for T. vaginalis is recommended, preferably by NAAT, for all sexually active women
between three weeks to three months after the end of treatment, regardless of whether
their sexual partner(s) were treated or not.12 The optimal time for repeat T. vaginalis NAAT
testing after completion of multi-dose oral metronidazole was three weeks or greater in a
recent study, informing this recommendation;99 repeat NAAT testing before this time carries
the risk of detecting remnant T. vaginalis nucleic acid that can still exist even if no viable
organism persists. If re-testing by three months is not possible, women should be re-tested
whenever they next seek medical care <12 months after treatment.12
5.9
Special situations
Infants
T. vaginalis has been documented to be transmitted perinatally in case reports,100 although
this is rare. In female newborns, T. vaginalis acquisition during birth may cause vaginal dis-
charge during the first week of life.101 Respiratory infection in newborns is also possible.102
Pregnant and lactating women
Several meta-analyses have found metronidazole to be safe for use in pregnant women in all
stages of pregnancy;103, 104 this is supported by current US guidelines.12 Tinidazole use should
be avoided in pregnant women based on preclinical data suggesting that it poses a moder-
ate risk.12 Limited data are available on the use of secnidazole in pregnant women however
there is no evidence of adverse developmental outcomes in animal studies.98
In lactating women who are administered metronidazole, withholding breastfeeding dur-
ing treatment and for 12–24 hours after the last dose will reduce the exposure of the infant
to metronidazole. For women treated with tinidazole, interruption of breastfeeding is rec-
ommended during treatment and for three days after the last dose.105
5-nitroimidazole hypersensitivity
The most common reactions associated with 5-nitroimidazoles (primarily metronidazole)
are immediate, type I IgE-mediated hypersensitivity reactions, occurring within 1-2 hours of
drug exposure. This includes urticaria and hives with potential life-threatening manifesta-
tions such as angioedema, bronchospasm, and anaphylaxis.106 Type II and IV hypersensitivity
reactions have also been described, although less commonly.107-110 The prevalence of met-
ronidazole hypersensitivity has been found to be approximately 0.15% in a study of 2,375,424
Kaiser Permanente health plan members (a representative sample of 1% of the US population).111
Although uncommon, treatment for T. vaginalis-infected patients with a history of 5-nitro-
imidazole hypersensitivity is challenging.112 If a prior IgE-mediated hypersensitivity reaction
has been confirmed based on patient history and/or graded oral challenge (drug provo-
cation test),112 desensitization to the 5-nitroimidazole performed by an allergist is the first
124
line of treatment.12 Oral107, 113 and intravenous114 protocols for desensitization have been
published, primarily involving metronidazole. Intensive monitoring is required during the
desensitization process due to the need for frequent drug administration and close moni-
toring for reactions; thus, it should typically be performed in the inpatient setting.112 After
completion of a desensitization protocol, patients are able to safely take oral metronida-
zole for 4-5 half-lives of the drug (half-life = 7-8 hours), approximately two days.111 If the
drug is not continued at regular intervals after successful completion of the desensitization
protocol, it will need to be restarted from the beginning to avoid breakthrough of a type I
hypersensitivity reaction.111
For patients in which metronidazole desensitization is not an option, use of other 5-nitro-
imidazoles such as tinidazole or secnidazole is not recommended because of the risk of
cross-reactivity within the same drug class.109 In this case, alternative treatment options
outside of the 5-nitroimidazoles should be used.12, 112 Use of these alternative treatments is
anecdotal, limited to vaginal formulations (the majority of which have to be compounded),
and may not reach all sites infected with T. vaginalis (i.e. Bartholin’s and Skene’s glands).12, 112
One option based on case reports is a prolonged course of vaginal boric acid 600 mg twice
daily for 60 days, either alone115, 116 or in combination with vaginal clotrimazole.117 Another
option based on case reports and case series data is vaginal paromomycin 6.25% cream dai-
ly for 8-14 days;118-120 topical use of this medication can result in painful vulvar ulcers that are
self-limited and resolve once treatment is discontinued. Use of lubricating jelly to the vulva
before use has been successful in preventing the development of these ulcers in some women.120
Persistent
T. vaginalis
infection
For patients who are experiencing persistent infection not due to sexual re-exposure, clinicians in
the US can request a trichomonas culture kit from the CDC to perform drug resistance testing (404-
718-4141; (https://www.cdc.gov/laboratory/specimen-submission/detail.html?CDCTestCode=CDC-10239).
CDC has experience with susceptibility testing for 5-nitroimidazole-resistant T. vaginalis as well
as management of infected patients and can provide guidance on treatment in these cases.
Based on resistance testing results, an alternative treatment regimen may be recommended.
Resistance rates of T. vaginalis for metronidazole and tinidazole have ranged from 4.3-
10%,121 although these data are not contemporary; resistance rates of secnidazole among
clinical T. vaginalis isolates are unknown. In vitro resistance may not always correlate with
clinical treatment failure,122 especially in pregnant women,53 but use of alternative treatment
regimens following drug resistance testing results in cure of resistant infections in >80% of
cases, suggesting that there is a benefit to drug resistance testing.123
Alternative treatment regimens for infections demonstrating in vitro drug resistance may
include 2 g oral metronidazole or tinidazole daily for seven days.12 If a patient fails the sev-
en day regimen of high dose oral metronidazole or tinidazole, two additional treatment
options have had successful results in women. The first is high dose oral tinidazole 2 g daily
plus vaginal tinidazole 500 mg twice daily for 14 days.124 If this fails, high dose oral tinidazole (1
g three times daily) plus vaginal paromomycin (4 g of 6.25% cream nightly) for 14 days could be
considered.125
125
HIV-infected women
In a RCT of HIV-infected women co-infected with T. vaginalis, seven day oral metronidazole
was found to be superior to single dose 2 g oral metronidazole.60 Further analysis revealed
that this superiority only occurred in the presence of BV.126 Studies have also found that pro-
tease inhibitors used for the treatment of HIV may interfere with the efficacy of single-dose
2 g oral metronidazole among HIV-infected women.127, 128
As previously mentioned, T. vaginalis screening (and treatment for positive cases) at entry
to care and annually is recommended for HIV-infected women.12 It has been estimated that if
this recommendation for T. vaginalis screening and treatment among HIV-infected women were
followed, the lifetime cost of new HIV infections prevented would approximate U.S. $159,264,000
and could potentially prevent new HIV cases secondary to female-to-male transmission.129
Partner management
Sexual partners of patients with T. vaginalis infection should be treated. Commonly, patients
are told by their providers to tell their partners to seek testing and treatment. Providers
should consider treating the partner(s) of positive patients presumptively. One method of
presumptive partner treatment is expedited partner therapy (EPT). EPT is the clinical prac-
tice of treating sexual partner(s) of patients diagnosed with an STI by providing prescrip-
tions or medications to the patient to take to his/her partner(s) without the health care pro-
vider first examining the partner(s).
One RCT demonstrated that partner treatment with single dose 2 g oral tinidazole resulted
in a >4 fold reduction in repeat infections among T. vaginalis-infected index women.130 Two
other studies using single dose 2-gram oral metronidazole for male partners of T. vagina-
lis-infected women found no effect of EPT131 or a borderline effect.132 While it is possible that
the two studies that used metronidazole were either underpowered or did not use a correct
control arm, it is also possible that oral tinidazole is a better treatment for men.
5.10
Future perspectives
Given that most studies have examined outcomes of symptomatic T. vaginalis infection,
additional studies are needed to examine the importance of asymptomatic infection. This
is particularly important given the proliferation of T. vaginalis molecular diagnostic tests,
including those available at the POC.94 Additional investigation of the role of T. vaginalis
in the etiology of PID is also needed, especially among HIV-uninfected women. Regarding
treatment, contemporary data on rates of T. vaginalis resistance among 5-nitroimidazoles,
including secnidazole, are needed. In addition, the role of oral secnidazole in the treatment
of persistent T. vaginalis infection should be elucidated. Finally, far less is known about T.
vaginalis infection in men, particularly the most optimal treatment.
126
Recommendations
Recommendation Quality of
evidence
Strength of
recommendation
Annual screening is recommended in women living with HIV. 4 C
Testing for non-genital trichomoniasis is not recommended. 5 D
A normal pH is not enough to exclude T. vaginalis infection. 2b B
Wet mount microscopy should be performed when trichomoniasis
is suspected, but a negative result does not exclude the diagnosis. 1b A
Molecular tests are currently the gold standard for the diagnosis of
trichomoniasis. 1a A
Oral metronidazole 400-500 mg twice daily for 7 days is the
recommended standard treatment for trichomoniasis in women,
regardless of the HIV status.
1a A
Single dose 2 g oral tinidazole or secnidazole can be considered as
an alternative. 2b B
Single dose 2 g oral metronidazole is no longer recommended in
women for treatment for trichomoniasis. 1a A
If a male is still infected with T. vaginalis after treatment with single
dose 2 g oral metronidazole and has been re-exposed to an un-
treated sexual partner, he should be re-dosed with another single
dose 2 g oral metronidazole.
5 D
If a male is still infected with T. vaginalis and he has not been re-ex-
posed, he should be given a course of oral metronidazole 500 mg
twice daily for 7 days.
5 D
Re-testing for T. vaginalis is recommended, preferably by nucleic
acid amplification test, for all sexually active women between 3
weeks to 3 months after the end of treatment regardless of wheth-
er or not their sexual partner(s) was/were treated.
4 C
If re-testing by 3 months is not possible, women should be
re-tested whenever they next seek medical care <12 months after
treatment.
5 D
Metronidazole is safe for use in pregnant women in all stages of
pregnancy. 1a A
Tinidazole use should be avoided in pregnant women. 4 C
Limited data are available on the use of secnidazole in pregnant
women, but there is no evidence of adverse developmental out-
comes in animal studies.
4 C
In lactating women who are administered metronidazole, with-
holding breastfeeding during treatment and for 12–24 hours after
the last dose is recommended.
4 C
In lactating women who are administered tinidazole, interruption
of breastfeeding is recommended during treatment and for 3 days
after the last dose.
4 C
127
In case of 5-nitroimidazoles hypersensitivity, desensitization is the
first line option of treatment. 4 C
For patients in which metronidazole desensitization is not an
option, use of other 5-nitroimidazoles such as tinidazole
or secnidazole is not recommended because of the risk of cross-re-
activity within the same drug class.
5 D
For patients who are experiencing persistent infection not due
to sexual re-exposure, culture and drug resistance testing are
recommended.
5 D
Sexual partners of patients with T. vaginalis infection should be
treated. 5 D
Providers should consider treating partner(s) of positive patients
presumptively, without the need of observing or testing them. 5 D
A single dose 2 g oral tinidazole as expedited partner therapy
may be superior to single dose metronidazole for male partners of
infected women.
1b A
References
1. Rowley, J.; Vander Hoorn, S.; Korenromp, E.; Low, N.; Unemo, M.; Abu-Raddad, L. J.; Chico, R. M.; Smolak, A.; Newman,
L.; Gottlieb, S.; Thwin, S. S.; Broutet, N.; Taylor, M. M., Chlamydia, gonorrhoea, trichomoniasis and syphilis: global
prevalence and incidence estimates, 2016. Bull World Health Organ 2019, 97, (8), 548-562P.
2. Van Gerwen, O. T.; Craig-Kuhn, M. C.; Jones, A. T.; Schroeder, J. A.; Deaver, J.; Buekens, P.; Kissinger, P. J.; Muzny, C. A.,
Trichomoniasis and adverse birth outcomes: a systematic review and meta-analysis. BJOG 2021, 128, (12), 1907-1915.
3. Sorvillo, F.; Smith, L.; Kerndt, P.; Ash, L., Trichomonas vaginalis, HIV, and African-Americans. Emerg Infect Dis 2001, 7, (6), 927-32.
4. Lazenby, G. B.; Taylor, P. T.; Badman, B. S.; McHaki, E.; Kor te, J. E.; Soper, D. E.; Young Pierce, J., An association between
Trichomonas vaginalis and high-risk human papillomavirus in rural Tanzanian women undergoing cervical cancer
screening. Clin Ther 2014, 36, (1), 38-45.
5. Allsworth, J. E.; Ratner, J. A.; Peipert, J. F., Trichomoniasis and other sexually transmitted infections: results from the
2001-2004 National Health and Nutrition Examination Surveys. Sex Transm Dis 2009, 36, (12), 738-44.
6. Moodley, P.; Wilkinson, D.; Connolly, C.; Moodley, J.; Sturm, A. W., Trichomonas vaginalis is associated with pelvic
inflammatory disease in women infected with human immunodeficiency virus. Clin Infect Dis 2002, 34, (4), 519-22.
7. Wiringa, A. E.; Ness, R. B.; Darville, T.; Beigi, R. H.; Haggerty, C. L., Trichomonas vaginalis, endometritis and sequelae
among women with clinically suspected pelvic inflammatory disease. Sex Transm Infect 2020, 96, (6), 436-438.
8. Mielczarek, E.; Blaszkowska, J., Trichomonas vaginalis: pathogenicity and potential role in human reproductive fail-
ure. Infection 2016, 44, (4), 447-58.
9. Van Gerwen, O. T.; Camino, A. F.; Sharma, J.; Kissinger, P. J.; Muzny, C. A., Epidemiology, Natural History, Diagnosis, and
Treatment of Trichomonas vaginalis in Men. Clin Infect Dis 2021, 73, (6), 1119-1124.
10. Yang, S.; Zhao, W.; Wang, H.; Wang, Y.; Li, J.; Wu, X., Trichomonas vaginalis infection-associated risk of cervical cancer:
A meta-analysis. European journal of obstetrics, gynecology, and reproductive biology 2018, 228, 166-173.
11. Patel, E. U.; Gaydos, C. A.; Packman, Z. R.; Quinn, T. C.; Tobian, A. A. R., Prevalence and Correlates of Trichomonas
vaginalis Infection Among Men and Women in the United States. Clin Infect Dis 2018, 67, (2), 211-217.
12. Workowski, K. A.; Bachmann, L. H.; Chan, P. A.; Johnston, C. M.; Muzny, C. A.; Park, I.; Reno, H.; Zenilman, J. M.; Bolan,
G. A., Sexually Transmitted Infections Treatment Guidelines, 2021. MMWR Recomm Rep 2021, 70, (4), 1-187.
13. Muzny, C. A., Why Does Trichomonas vaginalis Continue to be a “Neglected” Sexually Transmitted Infection? Clin
Infect Dis 2018, 67, (2), 218-220.
14. Petrin, D.; Delgaty, K.; Bhatt, R.; Garber, G., Clinical and microbiological aspects of Trichomonas vaginalis. Clin Micro-
biol Rev 1998, 11, (2), 300-17.
15. Lizarraga, A.; Munoz, D.; Strobl-Mazzulla, P. H.; de Miguel, N., Toward incorporating epigenetics into regulation of
gene expression in the parasite Trichomonas vaginalis. Mol Microbiol 2021, 115, (5), 959-967.
16. Burch, T. A.; Rees, C. W.; Reardon, L. V., Epidemiological studies on human trichomoniasis. Am J Trop Med Hyg 1959,
8, (3), 312-8.
128
17. Muzny, C. A.; Rivers, C. A.; Mena, L. A.; Schwebke, J. R., Genotypic characterization of Trichomonas vaginalis isolates
among women who have sex with women in sexual partnerships. Sex Transm Dis 2012, 39, (7), 556-8.
18. Crucitti, T.; Jespers, V.; Mulenga, C.; Khondowe, S.; Vandepitte, J.; Buve, A., Non-sexual transmission of Trichomonas
vaginalis in adolescent girls attending school in Ndola, Zambia. PLoS One 2011, 6, (1), e16310.
19. Charles, S. X., Epidemiology of trichomonas vaginalis (TV) in rural adolescent and juvenile children. J Trop Pediatr
1991, 37, (2), 90.
20. Kissinger, P. J.; Gaydos, C. A.; Sena, A. C.; Scott McClelland, R.; Soper, D.; Secor, W. E.; Legendre, D.; Workowski, K. A.;
Muzny, C. A., Diagnosis and Management of Trichomonas vaginalis: Summary of Evidence Reviewed for the 2021
Centers for Disease Control and Prevention Sexually Transmitted Infections Treatment Guidelines. Clin Infect Dis
2022, 74, (Supplement_2), S152-S161.
21. Schwebke, J. R.; Burgess, D., Trichomoniasis. Clin Microbiol Rev 2004, 17, (4), 794-803, table of contents.
22. Fichorova, R.; Fraga, J.; Rappelli, P.; Fiori, P. L., Trichomonas vaginalis infection in symbiosis with Trichomonasvirus and
Mycoplasma. Res Microbiol 2017, 168, (9-10), 882-891.
23. Mercer, F.; Johnson, P. J., Trichomonas vaginalis: Pathogenesis, Symbiont Interactions, and Host Cell Immune Re-
sponses. Trends Parasitol 2018, 34, (8), 683-693.
24. Fraga, J.; Rojas, L.; Sariego, I.; Fernandez-Calienes, A.; Nunez, F. A., Species typing of Cuban Trichomonas vaginalis
virus by RT-PCR, and association of TVV-2 with high parasite adhesion levels and high pathogenicity in patients.
Arch Virol 2012, 157, (9), 1789-95.
25. Bessarab, I. N.; Nakajima, R.; Liu, H. W.; Tai, J. H., Identification and characterization of a type III Trichomonas vaginalis
virus in the protozoan pathogen Trichomonas vaginalis. Arch Virol 2011, 156, (2), 285-94.
26. Graves, K. J.; Ghosh, A. P.; Schmidt, N.; Augostini, P.; Secor, W. E.; Schwebke, J. R.; Martin, D. H.; Kissinger, P. J.; Muzny, C.
A., Trichomonas vaginalis Virus Among Women With Trichomoniasis and Associations With Demographics, Clinical
Outcomes, and Metronidazole Resistance. Clin Infect Dis 2019, 69, (12), 2170-2176.
27. Bouchemal, K.; Bories, C.; Loiseau, P. M., Strategies for Prevention and Treatment of Trichomonas vaginalis Infections.
Clin Microbiol Rev 2017, 30, (3), 811-825.
28. Harp, D. F.; Chowdhury, I., Trichomoniasis: evaluation to execution. European journal of obstetrics, gynecology, and
reproductive biology 2011, 157, (1), 3-9.
29. Bassey, G. B.; Clarke, A. I. L.; Elhelu, O. K.; Lee, C. M., Trichomoniasis, a new look at a common but neglected STI in
African descendance population in the United States and the Black Diaspora. A review of its incidence, research
prioritization, and the resulting health disparities. J Natl Med Assoc 2022, 114, (1), 78-89.
30. Ryan, C. M.; de Miguel, N.; Johnson, P. J., Trichomonas vaginalis: current understanding of host-parasite interactions.
Essays Biochem 2011, 51, 161-75.
31. Secor, W. E.; Meites, E.; Starr, M. C.; Workowski, K. A., Neglected parasitic infections in the United States: trichomoni-
asis. Am J Trop Med Hyg 2014, 90, (5), 800-804.
32. Krieger, J. N.; Verdon, M.; Siegel, N.; Holmes, K. K., Natural history of urogenital trichomoniasis in men. J Urol 1993,
149, (6), 1455-8.
33. Schwebke, J. R.; Rompalo, A.; Taylor, S.; Sena, A. C.; Martin, D. H.; Lopez, L. M.; Lensing, S.; Lee, J. Y., Re- evaluating the
treatment of nongonococcal urethritis: emphasizing emerging pathogens--a randomized clinical trial. Clin Infect
Dis 2011, 52, (2), 163-70.
34. Van der Pol, B., Trichomonas vaginalis infection: the most prevalent nonviral sexually transmitted infection receives
the least public health attention. Clin Infect Dis 2007, 44, (1), 23-5.
35. Poole, D. N.; McClelland, R. S., Global epidemiology of Trichomonas vaginalis. Sex Transm Infect 2013, 89, (6), 418-22.
36. Van Der Pol, B.; Williams, J. A.; Orr, D. P.; Batteiger, B. E.; Fortenberry, J. D., Prevalence, incidence, natural history, and re-
sponse to treatment of Trichomonas vaginalis infection among adolescent women. J Infect Dis 2005, 192, (12), 2039-44.
37. Figueroa-Angulo, E. E.; Rendon-Gandarilla, F. J.; Puente-Rivera, J.; Calla-Choque, J. S.; Cardenas-Guerra, R. E.; Orte-
ga-Lopez, J.; Quintas-Granados, L. I.; Alvarez-Sanchez, M. E.; Arroyo, R., The effects of environmental factors on the
virulence of Trichomonas vaginalis. Microbes Infect 2012, 14, (15), 1411-27.
38. Beltran, N. C.; Horvathova, L.; Jedelsky, P. L.; Sedinova, M.; Rada, P.; Marcincikova, M.; Hrdy, I.; Tachezy, J., Iron-induced
changes in the proteome of Trichomonas vaginalis hydrogenosomes. PLoS One 2013, 8, (5), e65148.
39. Lehker, M. W.; Alderete, J. F., Iron regulates growth of Trichomonas vaginalis and the expression of immunogenic
trichomonad proteins. Mol Microbiol 1992, 6, (1), 123-32.
40. Lehker, M. W.; Arroyo, R.; Alderete, J. F., The regulation by iron of the synthesis of adhesins and cytoadherence levels
in the protozoan Trichomonas vaginalis. J Exp Med 1991, 174, (2), 311-8.
41. Lindrose, A. R.; Htet, K. Z.; O’Connell, S.; Marsh, J.; Kissinger, P. J., Burden of trichomoniasis among older adults in the
United States: a systematic review. Sex Health 2022, 19, (3), 151-156.
129
42. Lewis, F. M. T.; Spicknall, I. H.; Flagg, E. W.; Papp, J. R.; Kreisel, K. M., Incidence and Prevalence of Trichomonas vaginalis
Infection Among Persons Aged 15 to 59 Years: United States, 2018. Sex Transm Dis 2021, 48, (4), 232-237.
43. Kelley, C. F.; Rosenberg, E. S.; O’Hara, B. M.; Sanchez, T.; del Rio, C.; Sullivan, P. S., Prevalence of urethral Trichomonas
vaginalis in black and white men who have sex with men. Sex Transm Dis 2012, 39, (9), 739.
44. Carter-Wicker, K.; Utuama, O.; Omole, F., Can trichomoniasis cause pharyngitis? A case report. SAGE Open Med Case
Rep 2016, 4, 2050313X16682132.
45. Francis, S. C.; Kent, C. K.; Klausner, J. D.; Rauch, L.; Kohn, R.; Hardick, A.; Gaydos, C. A., Prevalence of rectal Trichomonas
vaginalis and Mycoplasma genitalium in male patients at the San Francisco STD clinic, 2005-2006. Sex Transm Dis
2008, 35, (9), 797-800.
46. Miller, M.; Liao, Y.; Gomez, A. M.; Gaydos, C. A.; D’Mellow, D., Factors associated with the prevalence and incidence
of Trichomonas vaginalis infection among African American women in New York city who use drugs. J Infect Dis
2008, 197, (4), 503-9.
47. Rogers, S. M.; Turner, C. F.; Hobbs, M.; Miller, W. C.; Tan, S.; Roman, A. M.; Eggleston, E.; Villarroel, M. A.; Ganapathi, L.;
Chromy, J. R.; Erbelding, E., Epidemiology of undiagnosed trichomoniasis in a probability sample of urban young
adults. PLoS One 2014, 9, (3), e90548.
48. Sutton, M.; Sternberg, M.; Koumans, E. H.; McQuillan, G.; Berman, S.; Markowitz, L., The prevalence of Trichomonas vagi-
nalis infection among reproductive-age women in the United States, 2001-2004. Clin Infect Dis 2007, 45, (10), 1319-26.
49. Shafir, S. C.; Sorvillo, F. J.; Smith, L., Current issues and considerations regarding trichomoniasis and human immuno-
deficiency virus in African-Americans. Clin Microbiol Rev 2009, 22, (1), 37-45, Table of Contents.
50. Cu-Uvin, S.; Ko, H.; Jamieson, D. J.; Hogan, J. W.; Schuman, P.; Anderson, J.; Klein, R. S.; Group, H. I. V. E. R. S., Prevalence,
incidence, and persistence or recurrence of trichomoniasis among human immunodeficiency virus (HIV)-positive
women and among HIV-negative women at high risk for HIV infection. Clin Infect Dis 2002, 34, (10), 1406-11.
51. Kissinger, P., Trichomonas vaginalis: a review of epidemiologic, clinical and treatment issues. BMC Infect Dis 2015, 15, 307.
52. Najafi, A.; Chaechi Nosrati, M. R.; Ghasemi, E.; Navi, Z.; Yousefi, A.; Majidiani, H.; Ghaneialvar, H.; Sayehmiri, K.; Gal-
van-Ramirez, M. L.; Fakhar, M., Is there association between Trichomonas vaginalis infection and prostate cancer
risk?: A systematic review and meta-analysis. Microb Pathog 2019, 137, 103752.
53. Lazenby, G. B.; Thompson, L.; Powell, A. M.; Soper, D. E., Unexpected High Rates of Persistent Trichomonas vaginalis
Infection in a Retrospective Cohort of Treated Pregnant Women. Sex Transm Dis 2019, 46, (1), 2-8.
54. Ijasan, O.; Okunade, K. S.; Oluwole, A. A., The prevalence and risk factors for Trichomonas vaginalis infection amongst
human immunodeficiency virus-infected pregnant women attending the antenatal clinics of a university teaching
hospital in Lagos, South-Western, Nigeria. Niger Postgrad Med J 2018, 25, (1), 21-26.
55. Muzny, C. A.; Blackburn, R. J.; Sinsky, R. J.; Austin, E. L.; Schwebke, J. R., Added benefit of nucleic acid amplification
testing for the diagnosis of Trichomonas vaginalis among men and women attending a sexually transmitted dis-
eases clinic. Clin Infect Dis 2014, 59, (6), 834-41.
56. Sutcliffe, S.; Newman, S. B.; Hardick, A.; Gaydos, C. A., Prevalence and correlates of Trichomonas vaginalis infection
among female US federal prison inmates. Sex Transm Dis 2010, 37, (9), 585-90.
57. Freeman, A. H.; Katz, K. A.; Pandori, M. W.; Rauch, L. M.; Kohn, R. P.; Liska, S.; Bernstein, K. T.; Klausner, J. D., Prevalence
and correlates of Trichomonas vaginalis among incarcerated persons assessed using a highly sensitive molecular
assay. Sex Transm Dis 2010, 37, (3), 165-8.
58. Balkus, J. E.; Richardson, B. A.; Rabe, L. K.; Taha, T. E.; Mgodi, N.; Kasaro, M. P.; Ramjee, G.; Hoffman, I. F.; Abdool Karim, S.
S., Bacterial Vaginosis and the Risk of Trichomonas vaginalis Acquisition Among HIV-1-Negative Women. Sex Transm
Dis 2014, 41, (2), 123-8.
59. Hinderfeld, A. S.; Simoes-Barbosa, A., Vaginal dysbiotic bacteria act as pathobionts of the protozoal pathogen Trich-
omonas vaginalis. Microb Pathog 2020, 138, 103820.
60. Kissinger, P.; Mena, L.; Levison, J.; Clark, R. A.; Gatski, M.; Henderson, H.; Schmidt, N.; Rosenthal, S. L.; Myers, L.; Martin,
D. H., A randomized treatment trial: single versus 7-day dose of metronidazole for the treatment of Trichomonas
vaginalis among HIV-infected women. J Acquir Immune Defic Syndr 2010, 55, (5), 565-71.
61. Kissinger, P.; Muzny, C. A.; Mena, L. A.; Lillis, R. A.; Schwebke, J. R.; Beauchamps, L.; Taylor, S. N.; Schmidt, N.; Myers, L.;
Augostini, P.; Secor, W. E.; Bradic, M.; Carlton, J. M.; Martin, D. H., Single-dose versus 7-day-dose metronidazole for
the treatment of trichomoniasis in women: an open-label, randomised controlled trial. Lancet Infect Dis 2018, 18,
(11), 1251-1259.
62. Legendre, D.; Muzny, C. A.; Kissinger, P., Pharmacokinetic and Pharmacodynamic Effects of Metronidazole May Account for
the Superior Efficacy of Multidose Therapy Among Women With Trichomoniasis. Sex Transm Dis 2019, 46, (11), 751-752.
63. Balkus, J. E.; Richardson, B. A.; Mochache, V.; Chohan, V.; Chan, J. D.; Masese, L.; Shafi, J.; Marrazzo, J.; Farquhar, C.; Mc-
Clelland, R. S., A prospective cohort study comparing the effect of single-dose 2 g metronidazole on Trichomonas
130
vaginalis infection in HIV-seropositive versus HIV-seronegative women. Sex Transm Dis 2013, 40, (6), 499-505.
64. Muzny, C. A.; Sunesara, I. R.; Martin, D. H.; Mena, L. A., Sexually transmitted infections and risk behaviors among
African American women who have sex with women: does sex with men make a difference? Sex Transm Dis 2011,
38, (12), 1118-25.
65. Simhan, H. N.; Anderson, B. L.; Krohn, M. A.; Heine, R. P.; Martinez de Tejada, B.; Landers, D. V.; Hillier, S. L., Host
immune consequences of asymptomatic Trichomonas vaginalis infection in pregnancy. Am J Obstet Gynecol 2007,
196, (1), 59 e1-5.
66. Fichorova, R. N., Impact of T. vaginalis infection on innate immune responses and reproductive outcome. J Reprod
Immunol 2009, 83, (1-2), 185-9.
67. Tanaka, Y.; Narahara, H.; Tak ai, N.; Yoshimatsu, J.; Anai, T.; Miyakawa, I., Interleukin-1beta and interleukin-8 in cervico-
vaginal fluid during pregnancy. Am J Obstet Gynecol 1998, 179, (3 Pt 1), 644-9.
68. Mann, J. R.; McDermott, S.; Barnes, T. L.; Hardin, J.; Bao, H.; Zhou, L., Trichomoniasis in pregnancy and mental retarda-
tion in children. Ann Epidemiol 2009, 19, (12), 891-9.
69. Masha, S. C.; Cools, P.; Sanders, E. J.; Vaneechoutte, M.; Crucitti, T., Trichomonas vaginalis and HIV infection acquisi-
tion: a systematic review and meta-analysis. Sex Transm Infect 2019, 95, (1), 36-42.
70. Barker, E. K.; Malekinejad, M.; Merai, R.; Lyles, C. M.; Sipe, T. A.; DeLuca, J. B.; Ridpath, A. D.; Gift, T. L.; Tailor, A.; Kahn,
J. G., Risk of Human Immunodeficiency Virus Acquisition Among High-Risk Heterosexuals With Nonviral Sexually
Transmitted Infections: A Systematic Review and Meta-Analysis. Sex Transm Dis 2022, 49, (6), 383-397.
71. Kissinger, P.; Adamski, A., Trichomoniasis and HIV interactions: a review. Sex Transm Infect 2013, 89, (6), 426-33.
72. Wang, C. C.; McClelland, R. S.; Reilly, M.; Overbaugh, J.; Emery, S. R.; Mandaliya, K.; Chohan, B.; Ndinya-Achola, J.;
Bwayo, J.; Kreiss, J. K., The effect of treatment of vaginal infections on shedding of human immunodeficiency virus
type 1. J Infect Dis 2001, 183, (7), 1017-22.
73. Kissinger, P.; Amedee, A.; Clark, R. A.; Dumestre, J.; Theall, K. P.; Myers, L.; Hagensee, M. E.; Farley, T. A.; Martin, D. H.,
Trichomonas vaginalis treatment reduces vaginal HIV-1 shedding. Sex Transm Dis 2009, 36, (1), 11-6.
74. Gottlieb, S. L.; Douglas, J. M., Jr.; Foster, M.; Schmid, D. S.; Newman, D. R.; Baron, A. E.; Bolan, G.; Iatesta, M.; Malotte, C. K.; Zenil-
man, J.; Fishbein, M.; Peterman, T. A.; Kamb, M. L.; Project, R. S. G., Incidence of herpes simplex virus type 2 infection in 5 sexu-
ally transmitted disease (STD) clinics and the effect of HIV/STD risk-reduction counseling. J Infect Dis 2004, 190, (6), 1059-67.
75. Boselli, F.; Chiossi, G.; Bortolamasi, M.; Gallinelli, A., Prevalence and determinants of genital shedding of herpes
simplex virus among women attending Italian colposcopy clinics. European journal of obstetrics, gynecology, and
reproductive biology 2005, 118, (1), 86-90.
76. Ginocchio, C. C.; Chapin, K.; Smith, J. S.; Aslanzadeh, J.; Snook, J.; Hill, C. S.; Gaydos, C. A., Prevalence of Trichomonas
vaginalis and coinfection with Chlamydia trachomatis and Neisseria gonorrhoeae in the United States as deter-
mined by the Aptima Trichomonas vaginalis nucleic acid amplification assay. J Clin Microbiol 2012, 50, (8), 2601-8.
77. Zhang, Z.; Li, Y.; Lu, H.; Li, D.; Zhang, R.; Xie, X.; Guo, L.; Hao, L.; Tian, X.; Yang, Z.; Wang, S.; Mei, X., A systematic review
of the correlation between Trichomonas vaginalis infection and infertility. Acta Trop 2022, 236, 106693.
78. Sena, A. C.; Miller, W. C.; Hobbs, M. M.; Schwebke, J. R.; Leone, P. A.; Swygard, H.; Atashili, J.; Cohen, M. S., Trichomonas
vaginalis infection in male sexual partners: implications for diagnosis, treatment, and prevention. Clin Infect Dis
2007, 44, (1), 13-22.
79. Swygard, H.; Sena, A. C.; Hobbs, M. M.; Cohen, M. S., Trichomoniasis: clinical manifestations, diagnosis and manage-
ment. Sex Transm Infect 2004, 80, (2), 91-5.
80. Landers, D. V.; Wiesenfeld, H. C.; Heine, R. P.; Krohn, M. A.; Hillier, S. L., Predictive value of the clinical diagnosis of lower
genital tract infection in women. Am J Obstet Gynecol 2004, 190, (4), 1004-10.
81. Edwards, T.; Burke, P.; Smalley, H.; Hobbs, G., Trichomonas vaginalis: Clinical relevance, pathogenicity and diagnosis.
Crit Rev Microbiol 2016, 42, (3), 406-17.
82. Wolner-Hanssen, P.; Krieger, J. N.; Stevens, C. E.; Kiviat, N. B.; Koutsky, L.; Critchlow, C.; DeRouen, T.; Hillier, S.; Holmes,
K. K., Clinical manifestations of vaginal trichomoniasis. JAMA 1989, 261, (4), 571-6.
83. Hobbs, M. M.; Sena, A. C., Modern diagnosis of Trichomonas vaginalis infection. Sex Transm Infect 2013, 89, (6), 434-8.
84. Vieira-Baptista, P.; Bornstein, J., Candidiasis, Bacterial Vaginosis, Trichomoniasis and Other Vaginal Conditions Affecting
the Vulva. In Vulvar Disease: Breaking the Myths, Bornstein, J., Ed. Springer International Publishing: Cham, 2019; pp 167-205.
85. Ohlemeyer, C. L.; Hornberger, L. L.; Lynch, D. A.; Swierkosz, E. M., Diagnosis of Trichomonas vaginalis in adolescent
females: InPouch TV culture versus wet-mount microscopy. The Journal of adolescent health : official publication of
the Society for Adolescent Medicine 1998, 22, (3), 205-8.
86. Rivers, C. A.; Muzny, C. A.; Schwebke, J. R., Diagnostic rates differ on the basis of the number of read days with the
use of the InPouch culture system for Trichomonas vaginalis screening. J Clin Microbiol 2013, 51, (11), 3875-6.
131
87. Gaydos, C. A.; Hobbs, M.; Marrazzo, J.; Schwebke, J.; Coleman, J. S.; Masek, B.; Dize, L.; Jang, D.; Li, J.; Chernesky, M.,
Rapid Diagnosis of Trichomonas vaginalis by Testing Vaginal Swabs in an Isothermal Helicase-Dependent AmpliVue
Assay. Sex Transm Dis 2016, 43, (6), 369-73.
88. Gaydos, C. A.; Schwebke, J.; Dombrowski, J.; Marrazzo, J.; Coleman, J.; Silver, B.; Barnes, M.; Crane, L.; Fine, P., Clinical
performance of the Solana(R) Point-of-Care Trichomonas Assay from clinician-collected vaginal swabs and urine
specimens from symptomatic and asymptomatic women. Expert Rev Mol Diagn 2017, 17, (3), 303-306.
89. Schwebke, J. R.; Hobbs, M. M.; Taylor, S. N.; Sena, A. C.; Catania, M. G.; Weinbaum, B. S.; Johnson, A. D.; Getman, D. K.;
Gaydos, C. A., Molecular testing for Trichomonas vaginalis in women: results from a prospective U.S. clinical trial. J
Clin Microbiol 2011, 49, (12), 4106-11.
90. Van Der Pol, B.; Williams, J. A.; Taylor, S. N.; Cammarata, C. L.; Rivers, C. A.; Body, B. A.; Nye, M.; Fuller, D.; Schwebke, J.
R.; Barnes, M.; Gaydos, C. A., Detection of Trichomonas vaginalis DNA by use of self-obtained vaginal swabs with the
BD ProbeTec Qx assay on the BD Viper system. J Clin Microbiol 2014, 52, (3), 885-9.
91. Schwebke, J. R.; Gaydos, C. A.; Davis, T.; Marrazzo, J.; Furgerson, D.; Taylor, S. N.; Smith, B.; Bachmann, L. H.; Ackerman,
R.; Spurrell, T.; Ferris, D.; Burnham, C. A.; Reno, H.; Lebed, J.; Eisenberg, D.; Kerndt, P.; Philip, S.; Jordan, J.; Quigley, N.,
Clinical Evaluation of the Cepheid Xpert TV Assay for Detection of Trichomonas vaginalis with Prospectively Collect-
ed Specimens from Men and Women. J Clin Microbiol 2018, 56, (2).
92. Van Der Pol, B., A profile of the cobas(R) TV/ MG test for the detection of Trichomonas vaginalis and Mycoplasma
genitalium. Expert Rev Mol Diagn 2020, 20, (4), 381-386.
93. Van Der Pol, B.; Torres-Chavolla, E.; Kodsi, S.; Cooper, C. K.; Davis, T. E.; Fife, K. H.; Taylor, S. N.; Augenbraun, M. H.;
Gaydos, C. A., Clinical Performance of the BD CTGCTV2 Assay for the BD MAX System for Detection of Chlamydia
trachomatis, Neisseria gonorrhoeae, and Trichomonas vaginalis Infections. Sex Transm Dis 2021, 48, (2), 134-140.
94. Morris, S. R.; Bristow, C. C.; Wierzbicki, M. R.; Sarno, M.; Asbel, L.; French, A.; Gaydos, C. A.; Hazan, L.; Mena, L.; Ma-
dhivanan, P.; Philip, S.; Schwartz, S.; Brown, C.; Styers, D.; Waymer, T.; Klausner, J. D., Performance of a single-use, rapid,
point-of-care PCR device for the detection of Neisseria gonorrhoeae, Chlamydia trachomatis, and Trichomonas
vaginalis: a cross-sectional study. Lancet Infect Dis 2021, 21, (5), 668-676.
95. Howe, K.; Kissinger, P. J., Single-Dose Compared With Multidose Metronidazole for the Treatment of Trichomoniasis
in Women: A Meta-Analysis. Sex Transm Dis 2017, 44, (1), 29-34.
96. Committee on Practice, B.-G., Vaginitis in Nonpregnant Patients: ACOG Practice Bulletin, Number 215. Obstet Gyne-
col 2020, 135, (1), e1-e17.
97. Muzny, C. A.; Schwebke, J. R.; Nyirjesy, P.; Kaufman, G.; Mena, L. A.; Lazenby, G. B.; Van Gerwen, O. T.; Graves, K. J.;
Arbuckle, J.; Carter, B. A.; McMahon, C. P.; Eder, S.; Shaw, J.; Pandey, B.; Chavoustie, S. E., Efficacy and Safety of Single
Oral Dosing of Secnidazole for Trichomoniasis in Women: Results of a Phase 3, Randomized, Double-Blind, Place-
bo-Controlled, Delayed-Treatment Study. Clin Infect Dis 2021, 73, (6), e1282-e1289.
98. Muzny, C. A.; Van Gerwen, O. T.; Legendre, D., Secnidazole: a treatment for trichomoniasis in adolescents and adults.
Expert Rev Anti Infect Ther 2022.
99. Craig-Kuhn, M. C.; Granade, C.; Muzny, C. A.; Van Der Pol, B.; Lillis, R.; Taylor, S. N.; Schmidt, N.; Martin, D. H.; Kissinger,
P., Optimal Timing for Trichomonas vaginalis Test of Cure Using Nucleic Acid Amplification Testing. Sex Transm Dis
2019, 46, (5), 312-316.
100. Schwandt, A.; Williams, C.; Beigi, R. H., Perinatal transmission of Trichomonas vaginalis: a case report. J Reprod Med
2008, 53, (1), 59-61.
101. Trintis, J.; Epie, N.; Boss, R.; Riedel, S., Neonatal Trichomonas vaginalis infection: a case report and review of literature.
Int J STD AIDS 2010, 21, (8), 606-7.
102. Carter, J. E.; Whithaus, K. C., Neonatal respiratory tract involvement by Trichomonas vaginalis: a case report and
review of the literature. Am J Trop Med Hyg 2008, 78, (1), 17-9.
103. Burtin, P.; Taddio, A.; Ariburnu, O.; Einarson, T. R.; Koren, G., Safety of metronidazole in pregnancy: a meta-analysis.
Am J Obstet Gynecol 1995, 172, (2 Pt 1), 525-9.
104. Caro-Patón, T.; Carvajal, A.; Martin de Diego, I.; Martin-Arias, L. H.; Alvarez Requejo, A.; Rodríguez Pinilla, E., Is metro-
nidazole teratogenic? A meta-analysis. Br J Clin Pharmacol 1997, 44, (2), 179-82.
105. Evaldson, G. R.; Lindgren, S.; Nord, C. E.; Rane, A. T., Tinidazole milk excretion and pharmacokinetics in lactating
women. Br J Clin Pharmacol 1985, 19, (4), 503-7.
106. Garcia-Rubio, I.; Martinez-Cocera, C.; Santos Magadan, S.; Rodriguez-Jimenez, B.; Vazquez-Cortes, S., Hypersensitivity
reactions to metronidazole. Allergol Immunopathol (Madr) 2006, 34, (2), 70-2.
107. Gendelman, S. R.; Pien, L. C.; Gutta, R. C.; Abouhassan, S. R., Modified oral metronidazole desensitization protocol.
Allergy Rhinol (Providence) 2014, 5, (2), 66-9.
132
108. Madsen, J. T.; Thormann, J.; Kerre, S.; Andersen, K. E.; Goossens, A., Allergic contact dermatitis to topical metronida-
zole - 3 cases. Contact Dermatitis 2007, 56, (6), 364-6.
109. Mishra, D.; Mobashir, M.; Zaheer, M. S., Fixed drug eruption and cross-reactivity between tinidazole and metronida-
zole. Int J Dermatol 1990, 29, (10), 740.
110. Kanwar, A. J.; Sharma, R.; Rajagopalan, M.; Kaur, S., Fixed drug eruption due to tinidazole with cross-reactivity with
metronidazole. Dermatologica 1990, 180, (4), 277.
111. Macy, E.; Romano, A.; Khan, D., Practical Management of Antibiotic Hypersensitivity in 2017. J Allergy Clin Immunol
Pract 2017, 5, (3), 577-586.
112. Van Gerwen, O. T.; Camino, A. F.; Bourla, L. N.; Legendre, D.; Muzny, C. A., Management of Trichomoniasis in the
Setting of 5-Nitroimidazole Hypersensitivity. Sex Transm Dis 2021, 48, (8), e111-e115.
113. Kurohara, M. L.; Kwong, F. K.; Lebherz, T. B.; Klaustermeyer, W. B., Metronidazole hypersensitivity and oral desensiti-
zation. J Allergy Clin Immunol 1991, 88, (2), 279-80.
114. Pearlman, M. D.; Yashar, C.; Ernst, S.; Solomon, W., An incremental dosing protocol for women with severe vaginal
trichomoniasis and adverse reaction to metronidazole. Am J Obstet Gynecol 1996, 174, (3), 934-6.
115. Muzny, C.; Barnes, A.; Mena, L., Symptomatic Trichomonas vaginalis infection in the setting of severe nitroimidazole
allergy: successful treatment with boric acid. Sex Health 2012, 9, (4), 389-91.
116. Backus, K. V.; Muzny, C. A.; Beauchamps, L. S., Trichomonas vaginalis Treated With Boric Acid in a Metronidazole
Allergic Female. Sex Transm Dis 2017, 44, (2), 120.
117. Aggarwal, A.; Shier, R. M., Recalcitrant Trichomonas vaginalis infections successfully treated with vaginal acidifica-
tion. J Obstet Gynaecol Can 2008, 30, (1), 55-58.
118. Nyirjesy, P.; S obel, J. D.; Weitz, M. V.; Leaman, D. J.; Gelone, S. P., Difficult-to-treat trichomoniasis: results with paromo-
mycin cream. Clin Infect Dis 1998, 26, (4), 986-8.
119. Helms, D. J.; Mosure, D. J.; Secor, W. E.; Workowski, K. A., Management of trichomonas vaginalis in women with
suspected metronidazole hypersensitivity. Am J Obstet Gynecol 2008, 198, (4), 370 e1-7.
120. Keating, M. A.; Nyirjesy, P., Trichomonas vaginalis Infection in a Tertiary Care Vaginitis Center. Sex Transm Dis 2015,
42, (9), 482-5.
121. Kirkcaldy, R. D.; Augostini, P.; Asbel, L. E.; Bernstein, K. T.; Kerani, R. P.; Mettenbrink, C. J.; Pathela, P.; Schwebke, J. R.; Se-
cor, W. E.; Workowski, K. A.; Davis, D.; Braxton, J.; Weinstock, H. S., Trichomonas vaginalis antimicrobial drug resistance
in 6 US cities, STD Surveillance Network, 2009-2010. Emerg Infect Dis 2012, 18, (6), 939-43.
122. Schwebke, J. R.; Barrientes, F. J., Prevalence of Trichomonas vaginalis isolates with resistance to metronidazole and
tinidazole. Antimicrob Agents Chemother 2006, 50, (12), 4209-10.
123. Bosserman, E. A.; Helms, D. J.; Mosure, D. J.; Secor, W. E.; Workowski, K. A., Utility of antimicrobial susceptibility testing
in Trichomonas vaginalis-infected women with clinical treatment failure. Sex Transm Dis 2011, 38, (10), 983-7.
124. Sobel, J. D.; Nyirjesy, P.; Brown, W., Tinidazole therapy for metronidazole-resistant vaginal trichomoniasis. Clin Infect
Dis 2001, 33, (8), 1341-6.
125. Nyirjesy, P.; Gilbert, J.; Mulcahy, L. J., Resistant trichomoniasis: successful treatment with combination therapy. S ex
Transm Dis 2011, 38, (10), 962-3.
126. Gatski, M.; Martin, D. H.; Levison, J.; Mena, L.; Clark, R. A.; Murphy, M.; Henderson, H.; Schmidt, N.; Kissinger, P., The
influence of bacterial vaginosis on the response to Trichomonas vaginalis treatment among HIV-infected women.
Sex Transm Infect 2011, 87, (3), 205-8.
127. Adamski, A.; Clark, R. A.; Mena, L.; Henderson, H.; Levison, J.; Schmidt, N.; Gebrekristos, H. T.; Martin, D. H.; Kissinger, P., The in-
fluence of ART on the treatment of Trichomonas vaginalis among HIV-infected women. Clin Infect Dis 2014, 59, (6), 883-7.
128. Kissinger, P.; Adamski, A.; Clark , R. A.; Mena, L.; Levison, J.; Martin, D. H., Does Antiretroviral Therapy Interfere With the
Treatment of Trichomonas vaginalis Among HIV+ Women? Sex Transm Dis 2013, 40, (6), 506-7.
129. Lazenby, G. B.; Unal, E. R.; Andrews, A. L.; Simpson, K., Cost-effectiveness analysis of annual Trichomonas vaginalis
screening and treatment in HIV-positive women to prevent HIV transmission. Sex Transm Dis 2014, 41, (6), 353-8.
130. Lyng, J.; Christensen, J., A double-blind study of the value of treatment with a single dose tinidazole of partners to
females with trichomoniasis. Acta Obstet Gynecol Scand 1981, 60, (2), 199-201.
131. Kissinger, P.; Schmidt, N.; Mohammed, H.; Leichliter, J. S.; Gift, T. L.; Meadors, B.; Sanders, C.; Farley, T. A., Patient-delivered
partner treatment for Trichomonas vaginalis infection: a randomized controlled trial. Sex Transm Dis 2006, 33, (7), 445-50.
132. Schwebke, J. R.; Desmond, R. A., A randomized controlled trial of partner notification methods for prevention of
trichomoniasis in women. Sex Transm Dis 2010, 37, (6), 392-6.
133
6.1
Introduction
Traditionally, lactobacilli have been considered the “good” and “protective” bacteria of the
vagina and thus excess or abnormal lactobacilli has typically not been considered a con-
cern. Cytolytic vaginosis (CV), lactobacillosis and “leptothrix” are conditions characterized
by abundant or oversized lactobacilli. It should be acknowledged that these conditions
are controversial and some experts feel that they are so poorly defined that their ex-
istence is unclear.1 Nevertheless, other experts report them in up to 5% of all the cases of
“vaginitis”.2 Lactobacilli have long been considered as a marker of vaginal health, which may
contribute to the skepticism of some authors towards these entities.
CV is characterized by an excessive number of lactobacilli and cytolysis. Lactobacillosis and
leptothrix are often used interchangeably and the definitions are evolving. The current ap-
proach is to consider lactobacillosis as an increased number of lactobacilli without cytolysis,
likely on the spectrum of CV, and to consider “leptothrix as a separate entity characterized
by elongated, serpiginous, bacteria thought to be lactobacilli, without cytolysis.3, 4
6.2
Cytolytic vaginosis
CV has been described in studies as early as 1961.5 In 1991, Cibley et al. published a pivotal
paper of this condition based on their experience in clinical practice.6 They hypothesized
the existence of CV, coined the term CV, provided clinical diagnostic criteria and suggested
treatment options. However, their paper has been critiqued for lack of rigor and CV has re-
mained a little known, understudied, and controversial condition.
6
CYTOLYTIC VAGINOSIS,
LACTOBACILLOSIS
AND LEPTOTHRIX
(alphabetical order)
Roni Kraut
Pedro Vieira-Baptista
134
Prevalence and epidemiology
Available studies suggest that the prevalence of CV in symptomatic women may be around
5%.2, 7-14 However, this remains unclear given the small number of studies, their overall low
quality, the different diagnostic techniques used, and the lack of standardized criteria.
A cross-sectional study found pregnant women with CV to have a decreased odds of group
B streptococci colonization.15
Risk factors
Evidence suggests that a cytolytic pattern may occur more commonly in pregnancy and
women <40 years; it seems to be less common in women with frequent intercourse.15-17 It is
currently unclear if prevalence varies according to geography or ethnic factors.
Complications
There is a paucity of studies that have examined the possible complications of CV and most
have significant risk of bias. Most studies focus on pregnancy5, 7, 15 and cervical dysplasia.18-20
A case-control study concluded that women with CV have an increased odds for vulvodynia.21
Signs and symptoms
CV may present without any signs or it may
include erythema, swelling, and erosions.
Symptoms include excess discharge (Figure
6.1), pruritus, burning, dysuria, pain, and
dyspareunia. The symptoms tend to be cycli-
cal, worsening after ovulation and improving
with the onset of menses. The signs and symp-
toms of CV overlap with those of vulvovaginal
candidiasis, making it difficult to differentiate
between these two conditions based on signs
and symptoms alone.16
Diagnosis
The diagnosis can be made with the use
of wet mount microscopy.3 It can also be
achieved using vaginal Gram or Pap stain.22
(Figure 6.2)
Figure 6.1 Typical discharge associated with
cytolytic vaginosis
135
Figure 6.2 Cytolytic vaginosis.
A– Wet mount microscopy (400x, phase contrast) B– Gram stain (1000x, oil immersion) C– Pap smear (conventional) (400x)
On saline microscopy the presence of abundant lactobacilli, with length variation, is not-
ed; other bacteria are typically scarce or, most often, absent; epithelial cells are fragmented
(bare nuclei and cytoplasmatic debris), and inflammation is absent. (Table 6.1) (Figure 6.2 A)
Cultures for Candida spp. are essential, but both entities may coexist.
The pH is low (often around 3.6) and the whiff test is negative. To date, there are no molecu-
lar tests commercially available for the diagnosis.
TABLE 6.1 Diagnostic criteria of cytolytic vaginosis.
* Donders criteria I or IIa or Ison-Hay criteria grade I or II23, 24
Criteria (all needed) Method
1. Abundant pleomorphic lactobacilli
2. Other bacteria scarce/absent*
3. Fragmented epithelial cells
4. Inflammation absent
Wet mount, Pap or vaginal Gram stain
The differential diagnosis primarily includes vulvovaginal candidiasis, especially non-albi-
cans candidiasis in which burning may predominate.
Treatment
The incidental diagnosis of a cytolytic pattern in asymptomatic women should not prompt
treatment.
The most commonly used treatment is sodium bicarbonate, either as irrigations or sitz baths. (Ta-
ble 6.2) Usually, the symptomatic relief is better achieved if treatment is used during the morning.
According to our clinical experience, while in some cases two weeks of treatment are enough,
most women will need to use it for several months or years, on demand. For some women, it
can be useful to record their symptoms in a calendar, so they can establish a pattern and predict
when they should resort to the prophylactic use of sodium bicarbonate. The treatment usually
does not cure CV (microscopically or clinically), but rather allows control of the symptoms.
136
Vaginal antibiotics have been suggested as second line options if sodium bicarbonate
proves to be insufficient, but data are scarce, and the effect appears to be transient.
TABLE 6.2 Treatments for cytolytic vaginosis
Treatment
First line
Sodium bicarbonate 30-40 g/L (sitz baths or irrigations) Once a day for 2 weeks
Discontinue tampon use
Until symptoms resolveDiscontinue antifungal treatment, antibiotics and probiotics
Use only water and soap to wash genital area
Alternatives Clindamycin vaginal cream (2%) Once a day 5 days
Amoxicillin 500 mg 3 times a day, orally, for 7 days
According to our clinical experience, changing the contraceptive method does not have an
impact in the presence of a cytolytic pattern or in symptom control.
If CV coexists with Candida spp., we recommend starting by adequately treating the latter,
and if symptoms persist, check if the microorganism was eliminated.
Special situations (infancy, pregnancy, postpartum/breastfeeding,
menopause, immunosuppression)
A cytolytic pattern is common during pregnancy and usually asymptomatic – it is likely that
such pattern is protective. We do not usually recommend treatment during this phase; if
treatment is required, only sitz baths should be used and never irrigations or antibiotics.
Future perspectives
CV is a little-known, under-researched condition. It has been suggested that the symptoms
and signs are physiological,25 and it is not typically considered in the differential diagnosis for
women presenting with vulvovaginal concerns.26 A gold-standard objective diagnostic technique
needs to be established and then used to further delineate this still equivocal condition.
6.3
Leptothrix
The first description of these bacteria was made in 186127 and has since been described in
Pap smear samples. The oldest clinical reference to this comes from a paper by Horowitz et
al., published in 1994 where it was referred to as lactobacillosis.28 They described women
with cyclical symptoms (irritation, burning and discharge), usually starting 7-10 days before
menses, in whom long and serpiginous anaerobic bacilli were identified. It is not yet proven
whether “leptothrix” can be a sole cause of vulvovaginal symptoms and it often is present
with other conditions.29 While usually assumed that leptothrix are lactobacilli, to date it re-
mains unresolved which species it belongs to. One theory is that these may be indeed com-
137
mon lactobacilli that due to pressure of the vaginal milieu (i.e. antibiotic or antifungal use)
acquire these characteristics.4
Prevalence and epidemiology
There are few data on the prevalence of the condition. A study from 1952 reported the presence
of “leptothrix” in 15.2% of pregnant Black women, contrasting with 0.5% in White women.27
A Russian series from 1997 reported a prevalence of these bacteria in genital discharge (of
both males and females) of 4%.30 In 2016, a similar rate was found by Meštrović et al. in
Pap smear samples.29 More recently, in a study involving 3620 women, a rate of 2.8% was
established.4
The original study by Horowitz et al. found the mean age of affected women to be of 33
years (range 24-59 years) and reported it also in postmenopausal and hysterectomized
women.28 In this study, the most commonly identified species of lactobacilli were L. acido-
philus and L. casei and most were strong hydrogen peroxide producers. However, it is unclear
if these species corresponded or not to leptothrix. In a more recent study, the mean age of
affected women was 38.8±10.65 years (range 18−76).31
Risk factors
No risk factors have been clearly identified. One study showed a higher prevalence of lepto-
thrix in women living with human immunodeficiency virus (HIV) infection (relative risk 3.0,
95% CI, 1.6–5.7), however no explanation for such was established.4
Some theories suggest that the previous use of antibiotics may be associated with the ap-
pearance of these long forms, similarly to what happens with other bacterial species.28 While
most women report previous episodes of candidiasis and antifungal treatments32, it is not
clear if those are causally associated with the presence of these bacteria or if it was an un-
confirmed diagnosis and empirical treatment.
In one observational study, in which a clear distinction between CV and lactobacillosis was
not made, it was suggested that symptoms’ worsening could be associated with the inges-
tion of dairy products.33
While the nature of the association is unknown, one study correlated the presence of lepto-
thrix with that of T. vaginalis, leading some authors to recommend excluding the presence
of the latter when the first is identified.29 More recent studies do not support the need to
exclude trichomoniasis when leptothrix is identified.4
Complications
A study presented in 2013 showed a prevalence of 13% of lactobacillosis/leptothrix in women
with vulvar pain, but it is not clear whether or not it distinguished this condition from CV.33
In one study there was no association between the presence of leptothrix and adverse out-
comes of fertility treatments or higher risk of cervical dysplasia.4
138
Signs and symptoms
Most women in whom leptothrix is identified are asymptomatic.4, 29
Horowitz et al reported the following symptoms: thick, white, curdy or creamy discharge
(83.3%), vulvar irritation (20.0%), burning (63.3%) or itching (86.7%), usually cyclical and
peaking immediately before the menses and waning once it starts.28 The vulvar exam is usu-
ally unremarkable, but there may be a discrete erythema and edema of the vulva and vag-
inal enanthema. The aspect of the cervix is unremarkable.27, 34 Given these symptoms and
signs overlap with vulvovaginal candidiasis, women often present after a lengthy duration
of symptoms (average 22.9 months, range 1-84 months) and have already been submitted
to several ineffective treatments.28 In the Vieira-Baptista et al. study, the average duration of
symptoms was 12.8±9.36 months and was significantly shorter than if women had another
explanation for their vulvovaginal symptoms.4
Diagnosis
The diagnosis is usually accomplished using wet mount microscopy, by identifying elon-
gated lactobacilli in the absence of cytolysis. These lactobacilli in some cases are very long
(60 m, range 40-75 m) serpiginous, non-motile, non-branching and sometimes appear-
ing segmented.28 Leptothrix can be found associated with different background microbiota
types, inflammation and other conditions. (Figure 6.3)
In one study, T. vaginalis was also present in 18% of cases and Candida spp. in 2%,27, 35, 36 but
more recent studies did not confirm an association with trichomoniasis. In fact, leptothrix
was more often found with a normal background microbiota (in 63.7% of cases) and associ-
ated with a higher risk of candidiasis and a lower risk of BV and CV.4
Figure 6.3 Leptothrix seen in wet mount microscopy (400x, phase contrast).
A– Leptothrix and normal background microbiota B– Leptothrix and Candida spp. blastospores
139
When seen on Pap smears, these bacteria tend to stain blue.37 The diagnosis is easier in con-
ventional smears than in liquid-based cytology.29 Using Gram stain, they are seen as Gram
positive rods.29 (Figure 6.4)
The pH has been described to be within low to
normal range (3.6-4.7).4, 28, 38
The differential diagnosis include vulvodynia32
and the presence of Actinomyces spp., which
usually branch at acute angles and are more
often found in women using intrauterine con-
traception.37
Treatment
If other changes are present (BV, Candida spp.,
trichomoniasis, cytolytic pattern) these should
be assumed to be the cause of the symptoms.
The initial studies on the topic found the in-
volved lactobacilli to be sensitive in vitro to
penicillin, ampicillin, tetracycline, clindamycin
and doxycycline and resistant to metronida-
zole, trimethoprim, gentamicin, amikacin, to-
bramycin, cefalexin, and ofloxacin.
In the Horowitz et al. report, expectant man-
agement for up to 3-4 months was not effec-
tive. The use of amoxicillin and clavulanate led
to clearance of symptoms in 86.3% of cases;
doxycycline was effective in all six cases in
which it was prescribed. There were no clinical or microscopic relapses during the 18 month
follow-up period.28 No control studies were performed.
The use of sodium bicarbonate douches may be helpful in some women, but data are scarce.4, 34
TABLE 6.3 Treatment options for leptothrix
First options Amoxicillin + clavulanate 500/125 3 times a day, per mouth,
for 7 days
Alternatives
Doxycycline 100 mg 2 times a day, per mouth, for
10 days
If allergic to penicillin
or failure of amoxicil-
lin + clavulanate
Nifuratel 200 mg and
Nifuratel 500 mg + nystatin 200000 IU
3 times a day, per mouth 7
days and once a day vagi-
nally 7 days
Scarce evidence
Sodium bicarbonate 30-40 g/L
(sitz baths or irrigations) Once a day for 2 weeks Scarce evidence of
limited results
Figure 6.4 Leptothrix seen using Gram stain
(1000x, oil immersion)
140
Special situations (infancy, pregnancy, postpartum/breastfeeding,
menopause, immunosuppression)
No data available.
6.4
Future perspectives
More rigorous epidemiological studies are needed to fully understand the role of these bac-
teria in health and disease.
To date, no 16S rRNA sequencing studies have been reported. A better understanding of the
exact species involved and its physiology could help clarify their role, the risks associated
with their presence (if any) and, if necessary, to more rationally treat.29
Recommendations
Recommendation Quality of
evidence
Strength of
recommendation
The diagnosis of cytolytic vaginosis can be made with the use of wet
mount microscopy, Gram or Pap stain. 4 C
Cultures for Candida spp. are recommended in all cases. 5 D
An increased pH excludes the diagnosis of cytolytic vaginosis. 4 C
The incidental diagnosis of a cytolytic pattern in asymptomatic women
should not prompt treatment. 5 D
If cytolytic vaginosis coexists with Candida spp. and the woman is
symptomatic, antifungals should be prescribed first. 5 D
Sodium bicarbonate (douche or sitz bath) is the recommended first
line treatment for cytolytic vaginosis. 4 D
Treatment of cytolytic vaginosis is not recommended during pregnancy. 5 D
Leptothrix should only be considered a possible cause of symptoms in
the absence of any other explanation. 4 C
There is no recommendation to exclude the presence of Trichomonas
vaginalis when leptothrix is identified. 4 C
References
1. Voytik, M.; Nyirjesy, P., Cytolytic Vaginosis: a Critical Appraisal of a Controversial Condition. Current Infectious Disease
Reports 2020, 22, (10), 26.
2. Wathne, B.; Holst, E.; Hovelius, B.; Mårdh, P. A., Vaginal discharge--comparison of clinical, laboratory and microbio-
logical findings. Acta Obstet Gynecol Scand 1994, 73, (10), 802-8.
3. Vieira-Baptista, P.; Grincevičienė, Š.; Oliveira, C.; Fonseca-Moutinho, J.; Cherey, F.; Stockdale, C. K., The International
Society for the Study of Vulvovaginal Disease Vaginal Wet Mount Microscopy Guidelines: How to Perform, Applica-
tions, and Interpretation. J Low Genit Tract Dis 2021, 25, (2), 172-180.
4. Vieira-Baptista, P.; Lima-Silva, J.; Preti, M.; Sousa, C.; Caiano, F.; Stockdale, C. K.; Bornstein, J., Vaginal Leptothrix: An
Innocent Bystander? Microorganisms 2022, 10, (8).
5. Zidovsky, J., The significance of parabasal (“postnatal”) cells in the vaginal smear in prolonged pregnancy. Acta Cytol
1961, 5, 393-398.
141
6. Cibley, L. J.; Cibley, L. J., Cytolytic vaginosis. Am J Obstet Gynecol 1991, 165, (4 Pt 2), 1245-9.
7. Akgun, I.; Yaziei Ensari, L., Cytolytic vaginosis: May cause infertility? . Virchows Arch 2012, 461, S1-S332.
8. Azevedo, S.; Lima-Silva, J.; Vieira-Baptista, P., Impact of the Sampling Site in the Result of Wet Mount Microscopy. J
Low Genit Tract Dis 2019, 23, (2), 176-181.
9. Batashki, I.; Markova, D.; Milchev, N., [Frequency of cytolytic vaginosis--examination of 1152 patients]. Akush Ginekol
(Sofiia) 2009, 48, (5), 15-6.
10. Cerikcioglu, N.; Beksac, M. S., Cytolytic vaginosis: misdiagnosed as candidal vaginitis. Infect Dis Obstet Gynecol 2004,
12, (1), 13-6.
11. Demirezen, S., Cytolytic vaginosis: examination of 2947 vaginal smears. Cent Eur J Public Health 2003, 11, (1), 23-4.
12. Fan, A. P.; Xue, F. X., [Clinical characteristics of aerobic vaginitis and its mixed infections]. Zhonghua Fu Chan Ke Za
Zhi 2010, 45, (12), 904-8.
13. Moghaddam, N.; Rajabi, P., The relationship between symptomatic vaginal candidiasis and lactobacillus flora, using
methenamine silver staining method. . RMJ 2009, 34, 82-85.
14. Raykova, V.; Baykushev, R.; Milanova, K.; Mitov, I., Prevalence of cytolytic vaginosis in symptomatic Bulgarian women
– need for microbiological study. Acta Microbiol Bulg 2018, 34, 95-99.
15. Rocchetti, T. T.; Marconi, C.; Rall, V. L.; Borges, V. T.; Corrente, J. E.; da Silva, M. G., Group B streptococci colonization in
pregnant women: risk factors and evaluation of the vaginal flora. Arch Gynecol Obstet 2011, 283, (4), 717-21.
16. Yang, S.; Zhang, Y.; Liu, Y.; Wang, J.; Chen, S.; Li, S., Clinical Significance and Characteristic Clinical Differences of
Cytolytic Vaginosis in Recurrent Vulvovaginitis. Gynecol Obstet Invest 2017, 82, (2), 137-143.
17. Giraldo, P.; Amara, l. R.; Goncalves, A.; Vicentini, R.; Martins, C.; Giraldo, H.; Fachini, A., [Influence of frequency of
vaginal intercourses and the use of doushing on vaginal microbiota] Rev Bras Ginecol Obstet 2005, 27, 257-262.
18. Nasielli, K.; Dudkiewicz, J.; Nasiell, M.; Hjerpe, A.; Silfverswärd, C., The occurrence of Bacillus vaginalis Döderlein and
cytolysis in dysplasia, carcinoma in situ, and invasive carcinoma of the uterine cervix. Acta Cytol 1972, 16, (1), 21-5.
19. Silva, C.; Almeida, E. C.; Côbo Ede, C.; Zeferino, V. F.; Murta, E. F.; Etchebehere, R. M., A retrospective study on cervical
intraepithelial lesions of low-grade and undetermined significance: evolution, associated factors and cytohistolog-
ical correlation. Sao Paulo Med J 2014, 132, (2), 92-6.
20. Vieira-Baptista, P.; Lima-Silva, J.; Tavares, S.; Beires, J.; Donders, G., Cytolytic vaginosis does not have an impact on
human papilloma virus (HPV) infection and cervical dysplasia. J Low Genit Tract Dis 2017, 21, (S27).
21. Vieira-Baptista, P.; Lima-Silva, J.; Xavier, J.; Beires, J.; Donders, G., Vaginal flora influences the risk of vulvodynia. J Low
Genit Tract Dis 2017, 21, S26.
22. Garg, K.; Khare, A.; Bansal, R.; Sharma, S.; Chaudhary, N., Effects of Different Contraceptive Methods on Cervico-Vag-
inal Cytology. J Clin Diagn Res 2017, 11, (7), Ec09-ec11.
23. Ison, C. A.; Hay, P. E., Validation of a simplified grading of Gram stained vaginal smears for use in genitourinary med-
icine clinics. Sex Transm Infect 2002, 78, (6), 413-5.
24. Donders, G. G. G.; Bellen, G.; Grinceviciene, S.; Ruban, K.; Vieira-Baptista, P., Aerobic vaginitis: no longer a stranger. Res
Microbiol 2017, 168, (9-10), 845-858.
25. Kaufman, R.; Friedrich, E.; Gardner, H., Benign diseases of the vulva and vagina. 3rd ed ed.; Year Book Medical Pub-
lishers: Chicago, 1989.
26. Workowski, K. A.; Bachmann, L. H.; Chan, P. A.; Johnston, C. M.; Muzny, C. A.; Park, I.; Reno, H.; Zenilman, J. M.; Bolan,
G. A., Sexually Transmitted Infections Treatment Guidelines, 2021. MMWR Recomm Rep 2021, 70, (4), 1-187.
27. Feo, L. G.; Dellette, B. R., Leptotrichia (Leptothrix) vaginalis. Am J Obstet Gynecol 1952, 64, (2), 382-6.
28. Horowitz, B. J.; Mårdh, P. A.; Nagy, E.; Rank, E. L., Vaginal lactobacillosis. Am J Obstet Gynecol 1994, 170, (3), 857-61.
29. Meštrović, T.; Profozić, Z., Clinical and microbiological importance of Leptothrix vaginalis on Pap smear reports.
Diagn Cytopathol 2016, 44, (1), 68-9.
30. Pliutto, A. M., [Laboratory diagnosis of bacterial vaginosis]. Klin Lab Diagn 1997, (3), 16-8.
31. Vieira-Baptista, P.; Bornstein, J., Candidiasis, Bacterial Vaginosis, Trichomoniasis and Other Vaginal Conditions Affecting
the Vulva. In Vulvar Disease: Breaking the Myths, Bornstein, J., Ed. Springer International Publishing: Cham, 2019; pp 167-205.
32. Paavonen, J., Vulvodynia--a complex syndrome of vulvar pain. Acta Obstet Gynecol Scand 1995, 74, (4), 243-7.
33. Ricci, P.; Troncoso, J., Lactobacillosis and Chronic Vulvar Pain: Looking for High-Risk Factors as Precursors in Women
Who Developed Vulvodynia. Journal of Minimally Invasive Gynecology 2013, 20, (6).
142
34. Hills, R. L., Cytolytic vaginosis and lactobacillosis. Consider these conditions with all vaginosis symptoms. Adv Nurse
Pract 2007, 15, (2), 45-8.
35. Von Maseela, T., [Leptothrix vaginalis. Morphological studies]. Fortschr Med 1976, 94, (16), 295-8.
36. McLellan, R.; Spence, M. R.; Brockman, M.; Raffel, L.; Smith, J. L., The clinical diagnosis of trichomoniasis. Obstet
Gynecol 1982, 60, (1), 30-4.
37. Fitzhugh, V. A.; Heller, D. S., Significance of a diagnosis of microorganisms on pap smear. J Low Genit Tract Dis 2008,
12, (1), 40-51.
38. Bibbo, M.; Harris, M. J., Leptothrix. Acta Cytol 1972, 16, (1), 2-4.
143
7.1
Introduction
Desquamative inflammatory vaginitis (DIV) made its debut in the medical literature in 1965,
by the hands of Gray and Barnes. In that paper, they presented their findings on 478 wom-
en complaining of vaginal discharge; among six of them the vaginas were thin, quite red-
dened, with numerous pus cells and with oval and round parabasal cells in the secretions.1
Three years later, Gardner described eight cases with similar features, among 3,000 women
with vaginitis. He summarized it by pointing the similarities between these findings and
those of atrophic vaginitis, despite the normal levels of estrogens in the affected women.
Given the lack of microbiological pattern in these women, he assumed that infection was
likely a secondary phenomenon.2
Despite the condition being known for almost 60 years, there has been no significant im-
provements in terms of understanding its etiology, diagnostic criteria or treatment, and it
still is omitted from most textbooks.
In 2002, Donders et al. described a new entity, referred to as aerobic vaginitis (AV). This
term emphasized the clear contrast with the far more common and acknowledged form of
dysbiosis: bacterial vaginosis (BV). These women, microscopically, presented with different
degrees of lactobacilli depletion, overgrowth of aerobic bacteria (mainly group B strepto-
cocci [GBS], Escherichia coli and Staphylococcus aureus), inflammation, and parabasal cells.
The authors proposed a scoring system, in which the highest scores match DIV.3 (Table 7.1)
7
AEROBIC VAGINITIS/
DESQUAMATIVE
INFLAMMATORY VAGINITIS
(alphabetical order)
Fulvio Borrela
Švitrigailė Grincevičienė
Mario Preti
Päivi Tommola
Pedro Vieira-Baptista
144
TABLE 7.1 Aerobic vaginitis score, after Donders G et al..4
LbG – lactobacillary grade; hpf – high power field; EC – epithelial cell; PBC – parabasal cell
A score <3 corresponds to “no AV”, score 3 - 4 to “light AV”, score 5 – 6 to “moderate AV” and
scores >6 to “severe AV” or DIV
Score LbG Number
of leucocytes
Proportion of toxic
leucocytes Background microbiota Proportion
of PBC
0 I or IIa ≤10/hpf None or sporadic Unremarkable or cytolysis <1%
1 IIb >10/hpf and ≤10/EC ≤50% of total leucocytes Small coliform bacteria 1-10%
2 III >10/EC >50% of total leucocytes Cocci or chains of cocci >10%
While it is not clear if these are two different entities or represent different aspects of the
same spectrum, for practical purposes, we opt to refer to it as AV/DIV.5, 6 Nevertheless, the
distinction between AV and DIV may be relevant in clinical practice, not only because of the
differences in severity and possible associated complications but mostly because treatment
regimens differ slightly between the two conditions. However, the distinction is not always
clear cut, as some overlapping in the clinical behavior exists.
Acknowledging this entity (or entities) is of uttermost importance not only for the proper
diagnosis and management of symptomatic women, but also because of its potential role in
obstetrical and non-obstetrical complications.4, 7, 8
7.2
Etiology and physiopathology
AV is characterized by moderate to severe colonization by facultative aerobic bacteria, de-
pletion of lactobacilli, and moderate to severe inflammatory reaction of the vulvovaginal
mucosa. Nevertheless an infectious etiology is unproven. It is assumed that this microbiota
shift may be secondary to a harsh milieu resulting in loss of lactobacilli species, and thus
allowing other bacteria to thrive.4
DIV (corresponding to severe AV), is characterized by colonization by aerobic facultative
bacteria, absence of lactobacilli and signs of severe inflammation of the vaginal mucosa.1, 9 It
may be postulated that it happens due to a systemic inflammatory condition that produces
vaginal inflammation resulting in abnormal vaginal microbiota, rather than the opposite.9
DIV is frequently a chronic condition, with most women reporting symptoms for more than
a year, and requiring treatment for a long period.4
Data showing similar effectiveness of vaginal steroids and 2% clindamycin in the treatment
of AV/DIV suggest that the presence of aerobic bacteria is not a primary cause but rather
the consequence of lactobacilli depletion and mucosal inflammation.9 The most commonly
reported bacteria isolated in women with AV/DIV are GBS, E. coli, S. aureus, Enterococcus fae-
calis and Klebsiella pneumoniae.10-12
Vitamin D deficiency has been postulated as a possible cause, but the correction of its level
did not lead to improvement.9 Pereira et al. hypothesized, based upon two cases, that a toxic
145
shock reaction to S. aureus in the vagina could lead to the development of the condition.13
Despite the increased proportion of parabasal epithelial cells in the vagina, lack of estrogens
has been excluded as the etiology of AV/DIV. Serum estradiol levels are usually within normal
range and isolated topical estrogens are usually insufficient for symptomatic improvement.9, 14
A genetic predisposition toward autoimmune processes has also been considered a possible
risk factor. Researchers reported an association of AV/DIV with other autoimmune condi-
tions, such as thyroiditis and Crohn’s disease.4, 9, 15
7.3
Prevalence and epidemiology
Prevalence of the condition is largely unknown, mainly due to lack of awareness and rec-
ognition of the disease by clinicians.16 Available data point to a prevalence rate of 2-25%
worldwide.4, 17 The lowest percentage was reported in South American countries (Brazil,
Chile) in which it was reported to be of only 2-3% in both pregnant and non-pregnant wom-
en.18, 19 The highest rates have been described in sub-Saharan countries (11-25%); in a study
conducted in Ethiopia, including only pregnant women, a slightly lower rate was reported
(8%).20-22 In Europe the range of disease prevalence is 8-12%, without any trend to be lower
among pregnant women.11, 23-27
The AV prevalence in pregnant women is reported to be 4.1-10.8%. 10-12 DIV has been report-
ed to be more common in perimenopausal White women.9
7.4
Risk factors
The risk factors for development of the disease are unknown. One study of AV found an
association with vaginal douching, long term antibiotic use, presence of an intrauterine de-
vice, and condom use.28 Most cases of DIV are idiopathic or primary, whereas secondary DIV
may complicate other non-genital tract autoinflammatory diseases (i.e. Crohn’s disease or
systemic lupus erythematous) or be associated with rituximab use.29, 30
7.5
Complications
Vaginal dysbiosis is acknowledged as a risk factor for several gynecological and obstetrical
complications.6-8 As in BV and trichomoniasis, the risks seem to be independent of the pres-
ence of symptoms.
AV/DIV has been associated with an increased risk for sexually transmitted infections, includ-
ing human immunodeficiency virus (HIV),20, 31 Chlamydia trachomatis32, 33 and possibly T. vagi-
nalis.3 In one study, the rate of C. trachomatis was more than three times higher in women with
AV, when compared to those with a normal vaginal microbiota (71.4 vs. 21.7%, p=0.018).32
Given that AV/DIV leads to the development of erosions and increased leukocytes in the
146
vaginal mucosa, in theory it may also increase the risk of transmission and acquisition of
herpes and human papillomavirus (HPV) infection. Some studies have shown a possible role
of AV/DIV in the development of abnormal Pap tests and cervical dysplasia.25, 34, 35 There are
no studies showing benefit in the treatment of AV/DIV to promote the clearance of the HPV
infection or regression of dysplasia.
Other possible non-obstetrical complications include infertility,20 pelvic inflammatory dis-
ease, and toxic shock syndrome.36
Obstetrical complications have been reported to be associated with AV, including abortion,
preterm labor, premature rupture of membranes (PROM), chorioamnionitis and funisitis (in-
flammation of the umbilical cord), puerperal sepsis and possibly neonatal sepsis.12, 37-40 In
one study, if AV was present in the first trimester the odds ratio (OR) of abortion (<25 weeks)
was of 5.2 (interval of confidence [IC] 95% 1.5–17.7) and that of preterm delivery (<35
weeks) was of 3.2 (IC 95% 1.2–9.1).12 In another study, the presence of severe AV in the first
trimester was correlated with a shorter cervical length at 20-24 weeks.41 A recent study con-
ducted in Vietnam showed an OR of 8.65 (IC 95% 1.41-53.16, p=0.020) of puerperal sepsis.
Bacterial colonization and infection of the lower genital tract may induce cytokines and
chemokines production, including interleukin (IL)-1β, IL-6 and IL-8, thus enhancing uterine
contractibility.42
7.6
Signs and symptoms
Many cases of AV/DIV are asympto-
matic, especially its mild forms. When
symptomatic, the most characteristic
clinical manifestation is an intense
inflammatory reaction of the vaginal
mucosa. This results in remarkable
tenderness, dyspareunia, stinging and
burning. Itching may also be present
in some cases. Vaginal and cervical
enanthema and submucosal pete-
chiae can be noted, and in the most
severe cases, the vestibule may also
be involved. The vaginal discharge is
purulent, sometimes copious, green
or yellow, and can be stained with
small amounts of blood.1, 9, 43 (Figure 7.1)
The symptoms are often long lasting and of fluctuating intensity.4 These manifestations are
strikingly different from those of the far more common causes of vaginal discharge, namely BV.
Figure 7.1 Severe aerobic vaginitis/desquamative
inflammatory vaginitis.
A– Vaginal and cervical petechiae B– Copious discharge
C– Vestibular involvement
147
7.7
Diagnosis
The diagnosis should be suspected based on the aforementioned symptoms and a compat-
ible vulvovaginal examination. The gold standard for diagnosis is wet mount microscopy
(WMM), preferably using phase contrast. (Figure 7.2)
Figure 7.2 Aerobic vaginitis/desquamative inflammatory vaginitis in wet mount microscopy (400x, phase contrast).
A– Moderate aerobic vaginitis B and C– Severe aerobic vaginitis/desquamative inflammatory vaginitis
The diagnosis can be established in the presence of:
1. reduced or absent Lactobacillus morphotypes;
2. presence of other bacteria (small rods or cocci – the latter sometimes in chains);
3. a significant amount of inflammatory cells;
4. presence of parabasal epithelial cells;
5. elevated pH and;
6. negative whiff test.4, 9
Table 7.1 shows an AV scoring sys-
tem, that can be used to diagnose
and grade its severity. The AV score
is a calculated sum of all sub-scores
(lactobacillary grade [LbG], number
of leucocytes, proportion of toxic
leucocytes, background microbiota,
and proportion of parabasal cells).
A score of less than 3 is normal, a
score 3 - 4 corresponds to “light AV”,
a score of 5 – 6 to “moderate AV” and
if higher than 6 to “severe AV.4
Gram-stain preparation is currently
not validated as a diagnostic tool
for AV/DIV, due to lack of criteria.5
(Figure 7.3)
Figure 7.3 Aerobic vaginitis/desquamative inflammatory
vaginitis aspects with Gram stain (1000x, oil immersion).
Chains of cocci seen in A and B
148
Also, it is believed that lactobacillary grades are more accurately evaluated using WMM.44
The pH is typically increased.
Routine bacterial cultures of the vaginal discharge are not recommended. These may, how-
ever, be used to rule out group A streptococci infection.45
The exclusion of the presence of T. vaginalis, using a nucleic acid amplification test is rec-
ommended, especially in the most severe cases, as the presentation of both conditions can
be very similar. In postmenopausal women, differential diagnosis from atrophic vaginitis is
not straightforward, but DIV does not respond to isolated estrogen replacement therapy.5
The distinction between AV/DIV and vaginal involvement by erosive lichen planus may be
hard to establish. However, some features of the latter can help in the differential diagnosis,
including the presence of well-demarcated erosions or glazed erythema at the vaginal intro-
itus and the involvement of other mucosal sites.46
7.8
Treatment
Recommended treatment options are shown in Table 7.2. No randomized clinical trials on
the treatment of AV/DIV exist and the few recommendations that exist are based on limited
observational studies and expert opinions.9, 16, 47, 48
The treatment regimen is guided by the microscopic findings: the presence of a disturbed
microbiota, inflammation and atrophy are treated, respectively, with topical antibiotics or
antiseptics, topical steroids, and topical estrogens.4 Usually, in severe AV/DIV, it is useful to
use a combination of all the three components at the beginning of the treatment. Both clin-
damycin and hydrocortisone have anti-inflammatory effect. Since severe AV/DIV is a chronic
condition, maintenance therapy, for a two to six months period, is recommended.16
In moderate AV conditions, when there is no suspicion of an underlying immune-inflam-
matory condition, treatment with a single course of clindamycin or dequalinium chloride
may be successful.4, 49 In cases with only slightly or moderately disturbed microbiota (lacto-
bacillary grade IIa or IIb) and without severe signs of inflammation (AV scores less than 5)
treatment with only topical antibiotics or antiseptics may be effective.
Kanamycin has good effect against Gram-negative bacilli, does not disrupt vaginal lactoba-
cilli and has also proven effective in AV treatment, used in a regimen of 100 mg vaginally for
six consecutive days.50
Oral moxifloxacin has shown some efficacy in AV treatment. Almost two thirds of the patients
treated with a single six day course of 400 mg moxifloxacin once daily, and 85% of those who
received a second course, were cured.51 Nevertheless, there is no reason to expose the woman
to a systemic antibiotic when the condition can be managed with a topical regimen.
It should be noted that metronidazole (vaginal or oral) is not a drug of choice, because bac-
teria associated with AV/DIV are not anaerobic species.
149
When a condition such as Crohn’s disease underlies DIV, the adequate treatment of the for-
mer with immunomodulators seems to effectively control the latter.4, 9, 15
In postmenopausal women, given the difficult distinction between AV/DIV and atrophic vagini-
tis, treatment with vaginal estrogens or prasterone alone may be tried.9 It can also be liberally
used in perimenopausal women, as it supports the natural vaginal lactobacilli-rich microbiota.16
Despite the theoretical benefits of the use of pro and prebiotics, data showing benefit are
scarce.52 In a randomized double-blind placebo-controlled trial, Heczko et al. demonstrated
that supplementation of standard antibiotic therapy with oral probiotics lengthened remis-
sion in patients with recurrent AV/BV, and improved clinical and microbiological parameters.53
TABLE 7.2 Recommended treatments for severe aerobic vaginitis/desquamative inflammatory
vaginitis
(a) Patients who are at risk of developing a yeast infection
(b) Peri- and postmenopausal women
Recommended treatments for
severe AV/DIV Regimen
Clindamycin
Clindamycin 2% cream
5 g vaginally daily at bedtime for 2-4
weeks; consider maintenance thera-
py twice a week for 2-6 months16, 47
Clindamycin 100 mg suppository
2 suppositories vaginally daily at
bedtime for 2-4 weeks; consider
maintenance therapy twice a week
for 2-6 months16, 48
Corticosteroids
Hydrocortisone 300 – 500 mg
Vaginally daily at bedtime for 2-4
weeks; consider maintenance thera-
py twice a week for 2-6 months48
Cortisone acetate suppository 25 mg
Vaginally daily at bedtime for 2-4
weeks; consider maintenance thera-
py twice a week for 2-6 months47
Ancillary treatments for DIV
Fluconazole(a) Fluconazole 150 mg Orally once weekly suppression for
2-6 months
Estradiol or estriol(b) Estradiol or estriol cream or suppository Vaginally twice a week for 2-6 months
Recommended treatments for
moderate AV
Dequalinium chloride Dequalinium chloride 10 mg
suppository 10 mg daily at bedtime for 6 days43, 49
Clindamycin Clindamycin 2% cream 5 g vaginally daily at bedtime for 7 days4
150
7.9
Special situations (pregnancy, postpartum/breastfeeding)
AV, as previously referred, is associated with adverse pregnancy outcomes, such as miscar-
riage, preterm delivery, PROM and stillbirth, intra-amniotic aerobic infection and chorioam-
niotitis.10 Nevertheless, there is no recommendation for systematic screening in pregnancy.
The authors opt to treat the condition when diagnosed and recommend, despite the ab-
sence of good quality data, that it should be screened in women with prior adverse obstet-
rical outcomes possibly associated with AV/DIV.
GBS, E. coli and S. aureus are often associated with AV/DIV and also with negative obstetrical
outcomes. Nevertheless, not all women colonized by these bacteria have AV/DIV criteria.
It is estimated that 7 to 25% of pregnant women between 35 and 37 weeks of gestation are pos-
itive for GBS.54-56 Universal screening of GBS is recommended, as it is the first cause of neonatal
mortality and morbidity worldwide57 and ascending vaginal infection can lead to chorioamnio-
nitis, PROM and endometritis,58-60 resulting in neonatal sepsis and stillbirth.
E. coli causing AV seems to be a separate strain from those isolated from the gut, bladder or
other sites of infection and thus specific strains may cause maternal disease.61 E. coli is as-
sociated with adverse pregnancy outcomes and may cause frequent infections in pregnant
women, mainly of the urinary tract and vagina, especially in the third trimester.
S. aureus is able to secrete exotoxins capable of inducing a cascade upregulating proinflam-
matory genes transcription, and is reported to be present in 4-22% of pregnant women.62, 63
It is a leading cause responsible of late-onset sepsis in newborn64 and a major pathogen in
pediatric intensive care units.65
Despite limited data, it is not clear if there is an advantage in screening for AV/DIV and other
bacteria beyond GBS in the 3rd trimester of pregnancy.17
Clindamycin is a broad spectrum antibiotic and its use in pregnant women is reported to
lower the incidence of premature delivery.66-70 It is considered a category B drug according
to the FDA Pregnancy and Lactation Labeling Rule.69 Data on its use in all trimesters is reas-
suring.71, 72 The route of administration is vaginal, either as 2% cream or 100 mg supposito-
ries. Clindamycin use during breastfeeding is unlikely to cause newborn side effects and the
va ginal route of administration is preferred.73
Moxifloxacin and kanamycin both have shown efficacy in AV/DIV treatment in non-pregnant
women. As potential hazards to the fetus, use of these compounds should be avoided dur-
ing pregnancy.74, 50, 75
The use of oral or vaginal probiotics can be considered and has limited efficacy in improving condi-
tions of the vaginal microbiota, but no clear impact on pregnancy outcomes have emerged.53, 76-78
151
7.10
Future perspectives
AV/DIV remains a poorly understood condition. More information is clearly needed in order
to allow better management of women with vaginitis, but also to reduce the associated
complications.
The complete understanding of the etiology of the condition would allow the rational devel-
opment of adequate and effective treatments. Animal models, based on bacteria inoculation
have been attempted, but it still remains unproven if AV/DIV is purely an infectious condition.79
The importance of screening and treatment of AV/DIV during pregnancy is an area that ur-
gently needs to be studied.
Improvement is needed in terms of definition of the condition and consequent develop-
ment of diagnostic tools. The development and validation of criteria for the diagnosis using
Gram stain may be helpful in increasing the accuracy of diagnosis.80 While some attempts
have been made to develop molecular tests, these still need further advancement and val-
idation. PCR-based methods targeted to detect bacteria that commonly associate with AV/
DIV may be of some usefulness in the future, especially in settings where microscopy is not
available.16, 24, 47 Artificial intelligence is likely to be a game-changer in this field.80 Meanwhile,
more education and training in the practice of WMM by clinicians is needed.
Recommendations
Recommendation Quality of
evidence
Strength of
recommendation
There is no recommendation to treat asymptomatic aerobic vaginitis/
desquamative inflammatory vaginitis to improve HPV clearance. 5 D
The gold standard for diagnosis of aerobic vaginitis/desquamative
inflammatory vaginitis is wet mount microscopy. 3b C
The “AV score” can be used for the grading of aerobic vaginitis. 4 C
Routine bacterial cultures of the vaginal discharge are not recommended. 5 D
In severe cases of suspected severe aerobic vaginitis/desquamative
inflammatory vaginitis the presence of T. vaginalis should be excluded
using a molecular test.
5 D
The treatment regimen is guided by the microscopic findings. 4 C
A combination of topical antibiotics or antiseptics, topical steroids, and
topical estrogens is usually recommended. 4 C
In moderate forms of aerobic vaginitis, a single course of topical clinda-
mycin or dequalinium chloride can be attempted. 5 D
In severe forms of aerobic vaginitis, maintenance therapy, for a two to six
months period, is recommended. 5 D
152
When there is an underlying condition for desquamative inflamma-
tory vaginitis (i.e. Crohn’s disease or rituximab treatment) it should be
controlled first.
4 C
In postmenopausal women with suspected aerobic vaginitis/desquama-
tive inflammatory vaginitis, treatment with topical estrogens should be
attempted initially.
5 D
There is no recommendation to use pre or probiotics. 4 C
There is no recommendation to screen for aerobic vaginitis/desquama-
tive inflammatory vaginitis during pregnancy. 5 D
References
1. Gray, L. A.; Barnes, M. L., VAGINITIS IN WOMEN, DIAGNOSIS AND TREATMENT. Am J Obstet Gynecol 1965, 92, 125-36.
2. Gardner, H. L., Desquamative inflammatory vaginitis: a newly defined entity. Am J Obstet Gynecol 1968, 102, (8), 1102-5.
3. Donders, G. G.; Vereecken, A.; Bosmans, E.; Dekeersmaecker, A.; Salembier, G.; Spitz, B., Definition of a type of abnor-
mal vaginal flora that is distinct from bacterial vaginosis: aerobic vaginitis. Bjog 2002, 109, (1), 34-43.
4. Donders, G. G. G.; Bellen, G.; Grinceviciene, S.; Ruban, K.; Vieira-Baptista, P., Aerobic vaginitis: no longer a stranger. Res
Microbiol 2017, 168, (9-10), 845-858.
5. Vieira-Baptista, P.; Grincevičienė, Š.; Oliveira, C.; Fonseca-Moutinho, J.; Cherey, F.; Stockdale, C. K., The International
Society for the Study of Vulvovaginal Disease Vaginal Wet Mount Microscopy Guidelines: How to Perform, Applica-
tions, and Interpretation. J Low Genit Tract Dis 2021, 25, (2), 172-180.
6. Lev-Sagie, A.; De Seta, F.; Verstraelen, H.; Ventolini, G.; Lonnee-Hoffmann, R.; Vieira-Baptista, P., The Vaginal Microbi-
ome: II. Vaginal Dysbiotic Conditions. J Low Genit Tract Dis 2022, 26, (1), 79-84.
7. De Seta, F.; Lonnee-Hoffmann, R.; Campisciano, G.; Comar, M.; Verstraelen, H.; Vieira-Baptista, P.; Ventolini, G.; Lev-Sagie, A.,
The Vaginal Microbiome: III. The Vaginal Microbiome in Various Urogenital Disorders. J Low Genit Tract Dis 2022, 26, (1), 85-92.
8. Ventolini, G.; Vieira-Baptista, P.; De Seta, F.; Verstraelen, H.; Lonnee-Hoffmann, R.; Lev-Sagie, A., The Vaginal Microbiome:
IV. The Role of Vaginal Microbiome in Reproduction and in Gynecologic Cancers. J Low Genit Tract Dis 2022, 26, (1), 93-98.
9. Reichman, O.; Sobel, J., Desquamative inflammatory vaginitis. Best Pract Res Clin Obstet Gynaecol 2014, 28, (7), 1042-50.
10. Ma, X.; Wu, M.; Wang, C.; Li, H.; Fan, A.; Wang, Y.; Han, C.; Xue, F., The pathogenesis of prevalent aerobic bacteria in
aerobic vaginitis and adverse pregnancy outcomes: a narrative review. Reprod Health 2022, 19, (1), 21.
11. Zodzika, J.; Rezeberga, D.; Jermakova, I.; Vasina, O.; Vedmedovska, N.; Donders, G., Factors related to elevated vaginal
pH in the first trimester of pregnancy. Acta Obstet Gynecol Scand 2011, 90, (1), 41-6.
12. Donders, G. G.; Van Calsteren, K.; Bellen, G.; Reybrouck, R.; Van den Bosch, T.; Riphagen, I.; Van Lierde, S., Predictive
value for preterm birth of abnormal vaginal flora, bacterial vaginosis and aerobic vaginitis during the first trimester
of pregnancy. Bjog 2009, 116, (10), 1315-24.
13. Pereira, N.; Edlind, T. D.; Schlievert, P. M.; Nyirjesy, P., Vaginal toxic shock reaction triggering desquamative inflamma-
tory vaginitis. J Low Genit Tract Dis 2013, 17, (1), 88-91.
14. Zaino, R. J.; Nucci, M. R.; Kurman, R. J., Diseases of the Vagina. In Blaustein’s Pathology of the Female Genital Tract,
Kurman, R. J.; Hedrick Ellenson, L.; Ronnett, B. M., Eds. Springer US: New York, NY, 2018; pp 1-63.
15. Shukla, A.; Surapaneni, S.; Sobel, J. D., Desquamative Inflammatory Vaginitis as an Extraintestinal Manifestation of
Crohn’s Disease. Current Infectious Disease Reports 2020, 22, (9), 24.
16. Paavonen, J.; Brunham, R. C., Bacterial Vaginosis and Desquamative Inflammatory Vaginitis. N Engl J Med 2018, 379,
(23), 2246-2254.
17. Nguyen, A. T. C.; Le Nguyen, N. T.; Hoang, T. T. A.; Nguyen, T. T.; Tran, T. T. Q.; Tran, D. N. T.; Nguyen, A. T. K.; Tran, L. M.; Nguyen,
D. H. C.; Le, T. M.; Ho, B. D.; Rööp, T.; Kõljalg, S.; Štšepetova, J.; Van Le, A.; Salumets, A.; Mändar, R., Aerobic vaginitis in the third
trimester and its impact on pregnancy outcomes. BMC Pregnancy Childbirth 2022, 22, (1), 432.
18. Gondo, D. C.; Duarte, M. T.; da Silva, M. G.; de Lima Parada, C. M., Abnormal vaginal flora in low-risk pregnant women
cared for by a public health service: prevalence and association with symptoms and findings from gynecological
exams. Rev Lat Am Enfermagem 2010, 18, (5), 919-27.
19. Villaseca, R.; Ovalle, A.; Amaya, F.; Labra, B.; Escalona, N.; Lizana, P.; Montoya, M. J.; Lillo, E.; Martínez, M. A., [Vaginal
infections in a Family Health Clinic in the Metropolitan Region, Chile]. Rev Chilena Infectol 2015, 32, (1), 30-6.
20. Donders, G. G.; Gonzaga, A.; Marconi, C.; Donders, F.; Michiels, T.; Eggermont, N.; Bellen, G.; Lule, J.; Byamughisa, J., In-
creased vaginal pH in Ugandan women: what does it indicate? Eur J Clin Microbiol Infect Dis 2016, 35, (8), 1297-303.
153
21. Vieira-Baptista, P.; Grinceviciene, S.; Bellen, G.; Sousa, C.; Saldanha, C.; Broeck, D. V.; Bogers, J. P.; Donders, G., Genital Tract
Infections in an Isolated Community: 100 Women of the Príncipe Island. Infect Dis Obstet Gynecol 2017, 2017, 3058569.
22. Yalew, G. T.; Muthupandian, S.; Hagos, K.; Negash, L.; Venkatraman, G.; Hagos, Y. M.; Meles, H. N.; Weldehaweriat, H.
H.; Al-Dahmoshi, H. O. M.; Saki, M., Prevalence of bacterial vaginosis and aerobic vaginitis and their associated risk
factors among pregnant women from northern Ethiopia: A cross-sectional study. PLoS One 2022, 17, (2), e0262692.
23. Tibaldi, C.; Cappello, N.; Latino, M. A.; Polarolo, G.; Masuelli, G.; Cavallo, F.; Benedetto, C., Maternal risk factors for
abnormal vaginal flora during pregnancy. Int J Gynaecol Obstet 2016, 133, (1), 89-93.
24. Rumyantseva, T. A.; Bellen, G.; Savochkina, Y. A.; Guschin, A. E.; Donders, G. G., Diagnosis of aerobic vaginitis by
quantitative real-time PCR. Arch Gynecol Obstet 2016, 294, (1), 109-14.
25. Vieira-Baptista, P.; Lima-Silva, J.; Pinto, C.; Saldanha, C.; Beires, J.; Martinez-de-Oliveira, J.; Donders, G., Bacterial vaginosis, aer-
obic vaginitis, vaginal inflammation and major Pap smear abnormalities. Eur J Clin Microbiol Infect Dis 2016, 35, (4), 657-64.
26. Tomusiak, A.; Heczko, P. B.; Janeczko, J.; Adamski, P.; Pilarczyk-Zurek, M.; Strus, M., Bacterial infections of the lower genital
tract in fertile and infertile women from the southeastern Poland. Ginekol Pol 2013, 84, (5), 352-8.
27. Donders, G. G., The prevalence of bacterial vaginosis and aerobic vaginitis in young Finish women. Apmis 2011, 119,
(3), 224-5; author reply 226.
28. Geng, N.; Wu, W.; Fan, A.; Han, C.; Wang, C.; Wang, Y.; Xue, F., Analysis of the Risk Factors for Aerobic Vaginitis: A
Case-Control Study. Gynecol Obstet Invest 2015.
29. Vempati, Y. S.; Sobel, J. D., Desquamative Inflammatory Vaginitis as an Expression of Systemic Lupus Erythematosus.
J Low Genit Tract Dis 2022, 26, (4), 345-346.
30. Yockey, L.; Dowst, S.; Zonozi, R.; Huizenga, N.; Murphy, P.; Laliberte, K.; Rosenthal, J.; Niles, J. L.; Mitchell, C. M., Inflammatory
vaginitis in women on long-term rituximab treatment for autoimmune disorders. BMC Womens Health 2021, 21, (1), 285.
31. Mascellino, M. T.; Iona, E.; Iegri, F.; Catania, S.; Trinchieri, V.; Oliva, P.; Amenta, L.; Revérberi, L.; Sorice, F., Evaluation of
vaginal microflora in patients infected with HIV. Microbiologica 1991, 14, (4), 343-9.
32. Marconi, C.; Donders, G. G.; Martin, L. F.; Ramos, B. R.; Duarte, M. T.; Parada, C. M.; Tristão, A. R.; Silva, M. G., Chlamydial
infection in a high risk population: association with vaginal flora patterns. Arch Gynecol Obstet 2012, 285, (4), 1013-8.
33. Donders, G.; De Wet, H. G.; Hooft, P.; Desmyter, J., Lactobacilli in Papanicolaou smears, genital infections, and preg-
nancy. Am J Perinatol 1993, 10, (5), 358-61.
34. Jahic, M.; Mulavdic, M.; Hadzimehmedovic, A.; Jahic, E., Association between aerobic vaginitis, bacterial vaginosis
and squamous intraepithelial lesion of low grade. Med Arch 2013, 67, (2), 94-6.
35. Plisko, O.; Zodzika, J.; Jermakova, I.; Pcolkina, K.; Prusakevica, A.; Liepniece-Karele, I.; Donders, G. G. G.; Rezeberga, D.,
Aerobic Vaginitis-Underestimated Risk Factor for Cervical Intraepithelial Neoplasia. Diagnostics (Basel) 2021, 11, (1).
36. MacPhee, R. A.; Miller, W. L.; Gloor, G. B.; McCormick, J. K.; Hammond, J. A.; Burton, J. P.; Reid, G., Influence of the
vaginal microbiota on toxic shock syndrome toxin 1 production by Staphylococcus aureus. Appl Environ Microbiol
2013, 79, (6), 1835-42.
37. Donders, G.; Bellen, G.; Rezeberga, D., Aerobic vaginitis in pregnancy. Bjog 2011, 118, (10), 1163-70.
38. Rezeberga, D.; Lazdane, G.; Kroica, J.; Sokolova, L.; Donders, G. G., Placental histological inflammation and repro-
ductive tract infections in a low risk pregnant population in Latvia. Acta Obstet Gynecol Scand 2008, 87, (3), 360-5.
39. Vedmedovska, N.; Rezeberga, D.; Teibe, U.; Polukarova, S.; Donders, G. G., Fetal growth restriction in Latvia. Int J
Gynaecol Obstet 2010, 111, (2), 185-6.
40. Yang, S.; Zhang, Y.; Liu, Y.; Wang, J.; Chen, S.; Li, S., Clinical Significance and Characteristic Clinical Differences of
Cytolytic Vaginosis in Recurrent Vulvovaginitis. Gynecol Obstet Invest 2017, 82, (2), 137-143.
41. Donders, G. G.; Van Calsteren, C.; Bellen, G.; Reybrouck, R.; Van den Bosch, T.; Riphagen, I.; Van Lierde, S., Association
between abnormal vaginal flora and cervical length as risk factors for preterm birth. Ultrasound Obstet Gynecol 2010.
42. Cauci, S.; Culhane, J. F.; Di Santolo, M.; McCollum, K., Among pregnant women with bacterial vaginosis, the hydro-
lytic enzymes sialidase and prolidase are positively associated with interleukin-1beta. Am J Obstet Gynecol 2008,
198, (1), 132.e1-7.
43. Nyirjesy, P.; Peyton, C.; Weitz, M. V.; Mathew, L.; Culhane, J. F., Causes of chronic vaginitis: analysis of a prospective
database of affected women. Obstet Gynecol 2006, 108, (5), 1185-91.
44. Donders, G. G.; Vereecken, A.; Dekeersmaecker, A.; Van Bulck, B.; Spitz, B., Wet mount microscopy reflects functional
vaginal lactobacillary flora better than Gram stain. J Clin Pathol 2000, 53, (4), 308-13.
45. Donders, G.; Greenhouse, P.; Donders, F.; Engel, U.; Paavonen, J.; Mendling, W., Genital Tract GAS Infection ISIDOG
Guidelines. J Clin Med 2021, 10, (9).
46. Simpson, R. C.; Thomas, K. S.; Leighton, P.; Murphy, R., Diagnostic criteria for erosive lichen planus affecting the vulva:
an international electronic-Delphi consensus exercise. Br J Dermatol 2013, 169, (2), 337-43.
154
47. Sobel, J. D.; Reichman, O.; Misra, D.; Yoo, W., Prognosis and treatment of desquamative inflammatory vaginitis. Obstet
Gynecol 2011, 117, (4), 850-855.
48. Sobel, J. D., Desquamative inflammatory vaginitis: a new subgroup of purulent vaginitis responsive to topical 2%
clindamycin therapy. Am J Obstet Gynecol 1994, 171, (5), 1215-20.
49. Mendling, W.; Weissenbacher, E. R.; Gerber, S.; Prasauskas, V.; Grob, P., Use of locally delivered dequalinium chloride
in the treatment of vaginal infections: a review. Arch Gynecol Obstet 2016, 293, (3), 469-84.
50. Tempera, G.; Furneri, P. M., Management of aerobic vaginitis. Gynecol Obstet Invest 2010, 70, (4), 244-9.
51. Wang, C.; Han, C.; Geng, N.; Fan, A.; Wang, Y.; Yue, Y.; Zhang, H.; Xue, F., Efficacy of oral moxifloxacin for aerobic vagi-
nitis. Eur J Clin Microbiol Infect Dis 2016, 35, (1), 95-101.
52. Borges, S.; Silva, J.; Teixeira, P., The role of lactobacilli and probiotics in maintaining vaginal health. Arch Gynecol
Obstet 2014, 289, (3), 479-89.
53. Heczko, P. B.; Tomusiak, A.; Adamski, P.; Jakimiuk, A. J.; Stefański, G.; Mikołajczyk-Cichońska, A.; Suda-Szczurek, M.;
Strus, M., Supplementation of standard antibiotic therapy with oral probiotics for bacterial vaginosis and aerobic
vaginitis: a randomised, double-blind, placebo-controlled trial. BMC Womens Health 2015, 15, 115.
54. Rocchetti, T. T.; Marconi, C.; Rall, V. L.; Borges, V. T.; Corrente, J. E.; da Silva, M. G., Group B streptococci colonization in
pregnant women: risk factors and evaluation of the vaginal flora. Arch Gynecol Obstet 2011, 283, (4), 717-21.
55. Zhou, X.; Brotman, R. M.; Gajer, P.; Abdo, Z.; Schüette, U.; Ma, S.; Ravel, J.; Forney, L. J., Recent advances in understand-
ing the microbiology of the female reproductive tract and the causes of premature birth. Infect Dis Obstet Gynecol
2010, 2010, 737425.
56. Romero, R.; Chaiworapongsa, T.; Espinoza, J., Micronutrients and intrauterine infection, preterm birth and the fetal
inflammatory response syndrome. J Nutr 2003, 133, (5 Suppl 2), 1668s-1673s.
57. Seale, A. C.; Bianchi-Jassir, F.; Russell, N. J.; Kohli-Lynch, M.; Tann, C. J.; Hall, J.; Madrid, L.; Blencowe, H.; Cousens, S.;
Baker, C. J.; Bartlett, L.; Cutland, C.; Gravett, M. G.; Heath, P. T.; Ip, M.; Le Doare, K.; Madhi, S. A.; Rubens, C. E.; Saha, S.
K.; Schrag, S. J.; Sobanjo-Ter Meulen, A.; Vekemans, J.; Lawn, J. E., Estimates of the Burden of Group B Streptococcal
Disease Worldwide for Pregnant Women, Stillbirths, and Children. Clin Infect Dis 2017, 65, (suppl_2), S200-s219.
58. Tsolia, M.; Psoma, M.; Gavrili, S.; Petrochilou, V.; Michalas, S.; Legakis, N.; Karpathios, T., Group B streptococcus colonization
of Greek pregnant women and neonates: prevalence, risk factors and serotypes. Clin Microbiol Infect 2003, 9, (8), 832-8.
59. Vornhagen, J.; Armistead, B.; Santana-Ufret, V.; Gendrin, C.; Merillat, S.; Coleman, M.; Quach, P.; Boldenow, E.; Alishetti,
V.; Leonhard-Melief, C.; Ngo, L. Y.; Whidbey, C.; Doran, K. S.; Curtis, C.; Waldorf, K. M. A.; Nance, E.; Rajagopal, L., Group
B streptococcus exploits vaginal epithelial exfoliation for ascending infection. J Clin Invest 2018, 128, (5), 1985-1999.
60. Shabayek, S.; Spellerberg, B., Group B Streptococcal Colonization, Molecular Characteristics, and Epidemiology.
Front Microbiol 2018, 9, 437.
61. Lobos, O.; Padilla, C., Phenotypic characterization and genomic DNA polymorphisms of Escherichia coli strains iso-
lated as the sole micro-organism from vaginal infections. Microbiology (Reading) 2009, 155, (Pt 3), 825-830.
62. Top, K. A.; Buet, A.; Whittier, S.; Ratner, A. J.; Saiman, L., Predictors of Staphylococcus aureus Rectovaginal Coloniza-
tion in Pregnant Women and Risk for Maternal and Neonatal Infections. J Pediatric Infect Dis Soc 2012, 1, (1), 7-15.
63. Bourgeois-Nicolaos, N.; Lucet, J. C.; Daubié, C.; Benchaba, F.; Rajguru, M.; Ruimy, R.; Andremont, A.; Armand-Lefèvre,
L., Maternal vaginal colonisation by Staphylococcus aureus and newborn acquisition at delivery. Paediatr Perinat
Epidemiol 2010, 24, (5), 488-91.
64. Grass, B.; Leone, A., Severe complications in preterm infant with late-onset Staphylococcus aureus sepsis. Swiss Soc.
Neonatol 2013.
65. Lazenby, G. B.; Soper, D. E.; Beardsley, W.; Salgado, C. D., Methicillin-resistant Staphylococcus aureus colonization
among women admitted for preterm delivery. Am J Obstet Gynecol 2012, 206, (4), 329.e1-5.
66. Larsson, P. G.; Fåhraeus, L.; Carlsson, B.; Jakobsson, T.; Forsum, U., Late miscarriage and preterm birth after treatment
with clindamycin: a randomised consent design study according to Zelen. Bjog 2006, 113, (6), 629-37.
67. Subramaniam, A.; Abramovici, A.; Andrews, W. W.; Tita, A. T., Antimicrobials for preterm birth prevention: an over-
view. Infect Dis Obstet Gynecol 2012, 2012, 157159.
68. Schmitz, T.; Sentilhes, L.; Lorthe, E.; Gallot, D.; Madar, H.; Doret-Dion, M.; Beucher, G.; Charlier, C.; Cazanave, C.;
Delorme, P.; Garabédian, C.; Azria, E.; Tessier, V.; Sénat, M. V.; Kayem, G., Preterm premature rupture of the mem-
branes: Guidelines for clinical practice from the French College of Gynaecologists and Obstetricians (CNGOF). Eur J
Obstet Gynecol Reprod Biol 2019, 236, 1-6.
155
69. Ugwumadu, A.; Manyonda, I.; Reid, F.; Hay, P., Effect of early oral clindamycin on late miscarriage and preterm de-
livery in asymptomatic women with abnormal vaginal flora and bacterial vaginosis: a randomised controlled trial.
Lancet 2003, 361, (9362), 983-8.
70. Kiss, H.; Petricevic, L.; Husslein, P., Prospective randomised controlled trial of an infection screening programme to
reduce the rate of preterm delivery. Bmj 2004, 329, (7462), 371.
71. Workowski, K. A.; Bachmann, L. H.; Chan, P. A.; Johnston, C. M.; Muzny, C. A.; Park, I.; Reno, H.; Zenilman, J. M.; Bolan,
G. A., Sexually Transmitted Infections Treatment Guidelines, 2021. MMWR Recomm Rep 2021, 70, (4), 1-187.
72. Lamont, R. F.; Nhan-Chang, C. L.; Sobel, J. D.; Workowski, K.; Conde-Agudelo, A.; Romero, R., Treatment of abnormal
vaginal flora in early pregnancy with clindamycin for the prevention of spontaneous preterm birth: a systematic
review and metaanalysis. Am J Obstet Gynecol 2011, 205, (3), 177-90.
73. Drugs and Lactation Database (LactMed) [Internet]. Bethesda (MD): National Library of Medicine (US); 2006-. Clin-
damycin. [Updated 2021 Feb 15]. Available from: https://www.ncbi.nlm.nih.gov/books/NBK501208/. 2006.
74. Von Keutz, E.; Rühl-Fehlert, C.; Drommer, W.; Rosenbruch, M., Effects of ciprofloxacin on joint cartilage in immature
dogs immediately after dosing and after a 5-month treatment-free period. Arch Toxicol 2004, 78, (7), 418-24.
75. Wang, Z.; Liou, L., Auditory effect of kanamycin given to newborn guinea pigs whose mothers received kanamycin
during pregnancy. Ann Otol Rhinol Laryngol 1994, 103, (12), 983-5.
76. Russo, R.; Edu, A.; De Seta, F., Study on the effects of an oral lactobacilli and lactoferrin complex in women with
intermediate vaginal microbiota. Arch Gynecol Obstet 2018, 298, (1), 139-145.
77. Samuel, T. M.; Sakwinska, O.; Makinen, K.; Burdge, G. C.; Godfrey, K. M.; Silva-Zolezzi, I., Preterm Birth: A Narrative
Review of the Current Evidence on Nutritional and Bioactive Solutions for Risk Reduction. Nutrients 2019, 11, (8).
78. Othman, M.; Neilson, J. P.; Alfirevic, Z., Probiotics for preventing preterm labour. Cochrane Database Syst Rev 2007,
(1), Cd005941.
79. Fatahi Dehpahni, M.; Chehri, K.; Azadbakht, M., Therapeutic effects of silver nanoparticle and L-carnitine on aerobic
vaginitis in mice: an experimental study. Bioimpacts 2022, 12, (1), 33-42.
80. Dong, M.; Wang, C.; Li, H.; Yan, Y.; Ma, X.; Li, H.; Li, X.; Wang, H.; Zhang, Y.; Qi, W.; Meng, K.; Tian, W.; Wang, Y.; Fan, A.;
Han, C.; Donders, G. G. G.; Xue, F., Aerobic Vaginitis Diagnosis Criteria Combining Gram Stain with Clinical Features:
An Establishment and Prospective Validation Study. Diagnostics (Basel) 2022, 12, (1).
156
157
8.1
Introduction
The reduction of estrogen production associated with menopause leads to genital and sys-
temic changes. One of the most common and more uncomfortable consequences of meno-
pause is vulvovaginal atrophy (VVA).1 Several other terms are used to refer to this condition,
including atrophic vaginitis, urogenital atrophy, urogenital syndrome, and genitourinary
syndrome of menopause (GSM).2 The term atrophic vaginitis may be used when inflamma-
tion is present, along with atrophy.3 While hot flashes usually subside with time, VVA often
persists and might worsen if left untreated.4
8.2
Etiology and physiopathology
The vaginal wall has estrogen, progesterone, and androgen receptors. During the reproduc-
tive years, the female genital tract maintains its trophism under the stimulation of estrogens
and progesterone. The estrogen receptor density is higher in the vagina and lower in the
external genitalia. Progesterone receptors are found in the vagina and the transitional epi-
thelium of the vulvovaginal junction. Androgens also play a significant role in lower genital
tract trophism. The density of androgen receptors is low in the vagina and higher in the ex-
ternal genitalia.5-11 With the decline of ovarian function after menopause, the entire genital
tract becomes atrophic.
The vaginal microbiome (VMB) varies throughout a woman’s life. Levels of sex hormones,
glycogen content in the vaginal epithelium, menstrual cycle, vaginal pH, intercourse, and
immune responses influence these changes. Lactobacillus spp. dominance in the vaginal
niche is generally driven by the availability of glycogen, which accumulates in an estro-
gen-dependent manner in the cervicovaginal environment.12-14 The VMB has been primarily
studied in reproductive-age women. While 20 species of lactobacilli have been found in the
vagina, it is usually dominated by a single species, more often L. crispatus or L. iners.14, 15
8
VULVOVAGINAL ATROPHY
(alphabetical order)
Susana Aidé
Švitrigailė Grincevičienė
Faustino R. Pérez-Lopez
Isabel do Val
158
During the reproductive years, the “normal” pH is usually lower than 4.5 in White and Asian
women, and slightly higher in Black and Hispanic women.16 Lactobacilli in the vagina play
an important protective role, counteracting the overgrowth of other microorganisms which
may compete for nutrients and tissue adherence. This function is accomplished by modu-
lating the local immune system, reducing the vaginal pH, producing organic acids (mainly
lactic acid), and antimicrobial substances, such as bacteriocins. The glycogen content of the
vaginal epithelium alters with the estrogen levels and, in general, high estradiol levels favor
a lactobacilli-dominant environment.14, 15
Circulating estrogen decreases drastically in menopause, leading to a reduction of Lactoba-
cillus spp. dominance and a concomitant increase in the diversity of species. Despite some
contradictory findings, some studies show that moderate to severe atrophy and dryness can
be associated with community state type IV-A (diversity group, not dominated by Fannyhes-
sea [Atopobium] spp. and/or Gardnerella spp.), while the state type IV-B (corresponding to
bacterial vaginosis) is less symptomatic.14, 15
Hypoestrogenism induces a decline in the vaginal epithelium glycogen level, which is the
substrate for Lactobacillus spp. The resulting depletion of lactobacilli leads to a pH increase,
which is typical of VVA. Nevertheless, some menopausal women still have a strong presence
of lactobacilli in their VMB.15 The connection between the vaginal microbiota and estrogens
demonstrates the importance of its use to prevent or treat VVA. In postmenopausal women
with VVA, vaginal or oral low dose estrogen therapy effectively increases the level of Lac-
tobacillus spp., decreases Gardnerella spp. and vaginal pH, and also leads to a significant
improvement in the Vaginal Maturation Index ( VMI).17 The transition from a lactobacilli dom-
inated to a non-dominated VMB is neither abrupt nor time predictable.
Hypoestrogenism affects the normal structure and function of the genital tissues, largely
contributing to the loss of mucosal elasticity, and inducing the fusion and hyalinization of
collagen fibers, and the fragmentation of elastin fibers. Estrogen receptor (ER)-α is present in
the vaginal tissues of both pre and postmenopausal women, whereas ER-β appears to have
no or low expression in postmenopausal vaginal tissue.
There is a decrease in the hydration of the vaginal mucosa in the dermal layer, with a re-
duction of mucopolysaccharides and intercellular hyaluronic acid, which generates a thin
stratified epithelium with only the basal and parabasal layers.18
8.3
Prevalence and epidemiology
The self-reported prevalence of symptoms of VVA ranges from 4% in the early postmeno-
pausal years to 50% among late postmenopausal women (>10 years of menopause).19, 20
In addition to menopause, VVA can be a physiological finding during breastfeeding due to
the transient but significant hypoestrogenism seen during that period. Non-physiological
conditions such as immunological disorders, premature ovarian failure, oophorectomy, radi-
otherapy, and chemotherapy can also cause VVA. Additionally, some endocrine treatments,
such as tamoxifen, aromatase inhibitors, progestins, and gonadotropin-releasing hormone
analogs can induce symptoms of VVA.1, 21
159
8.4
Complications
The symptoms of VVA impact the quality-of-life, sexual function, social or mental health (anxi-
ety and/or depressive symptoms, isolation, etc.). Loss of estrogen predisposes to urinary symp-
toms such as urgency, dysuria, and nocturia, in addition to recurrent urinary tract infections.
Estrogens play an important role in urinary continence through several mechanisms, including
its effect on the vessels of the periurethral region, on the striated and smooth muscles and on
the pelvic connective tissue, thus, hypoestrogenism can potentiate stress urinary incontinence.1
There may be progressive loss of elasticity, thinning of the vaginal walls, shortening of the
vaginal barrel and disappearance of the mucosal rugae. Consequently, the mucosa may be-
come friable and easily damaged, leading to petechiae, dyspareunia and bleeding upon
contact, creating more sexual difficulties.1 (Figure 8.1)
Figure 8.1 A and B– Colposcopic aspect of the vagina of a postmenopausal woman. Loss of vaginal rugae,
petechiae, and easy bleeding.
Sexual dysfunction can be potentiated by other conditions that are prevalent during the
postmenopausal years, such as: depressive symptoms, trauma, decreased mobility, previous
hysterectomy, hot flashes, sleep disorders, use of multiple drugs, overweight, and chronic
diseases (including metabolic syndrome).1
8.5
Signs and symptoms
The most frequent symptoms are vaginal dryness, burning, pain, itching, and vulvar irrita-
tion. Upon clinical examination, signs of vaginal inflammation with hyperemia in addition
to yellowish discharge may be present. These may be associated with sexual discomfort,
including dyspareunia or post-coital bleeding.20 As the urethra and the bladder trigone are
estrogen-dependent tissues, its deficiency in postmenopausal women can contribute to uri-
nary incontinence, urgency, and recurrent urinary tract infections.1
160
8.6
Diagnosis
The diagnosis of VVA is based on symptoms, complemented with clinical examination. The
provider must rule out possible clinical conditions that are part of the differential diagnosis,
such as aerobic vaginitis/desquamative inflammatory vaginitis, trichomoniasis, and derma-
tosis (lichen sclerosus, erosive lichen planus, lichen simplex chronicus, etc.).1 (Table 8.1)
TABLE 8.1 Differential diagnosis of genital diseases or conditions during the identification of
atrophic vaginitis.
Adapted from Pérez-López et al.1
Disease or condition Clinical characteristics
Vaginal atrophy Associated with hypoestrogenism states; thin and fragile vaginal
epithelium, but inflammation absent.
Atrophic vaginitis Term used when inflammation is present, along with atrophy.
Desquamative inflammatory vaginitis A syndrome that is frequently unrecognized, characterized by vaginal
enanthema, pethechiae and purulent discharge (see chapter 7).
Trichomoniasis A sexually transmitted infection caused by the protozoan Trichomon-
as vaginalis (see chapter 5).
Erosive lichen planus Inflamed painful red plaques or erosions that can affect the skin, nails,
and mucous membranes, including the genital area.
Clinical criteria of VVA are: vaginal dryness, itching or irritation, and dyspareunia; the vulvar
examination may show atrophy of the labia minora, pubic hair scarcity, reduction of the vol-
ume of the labia majora, retraction of the vestibule and the presence of an urethral caruncle.
The vagina is usually pale, dry and smooth, with loss of rugae. (Figure 8.1)
In some cases, it may however, be shiny and a purulent discharge may be present. In the
presence of signs of inflammation, it may be classified as atrophic vaginitis. VMI (Figure 8.2)
is not usually required in clinical practice but can be a simple way of documenting clinical
findings and their course.1
Figure 8.2 Flowchart for the clinical assessment in suspected vaginal atrophy/atrophic vaginitis.
Adapted from Pérez-López et al. 2021.1
VMI– Vaginal maturation index
161
Wet mount microscopy allows immediate assessment of the hormonal status of the vagi-
na.3 Vaginal atrophy is characterized by is an increase in parabasal cells and a decrease in
superficial cells.3 Sometimes, abundant leukocytes and the presence of bacteria other than
Lactobacillus morphotypes can be found, resembling desquamative inflammatory vaginitis.
(Figure 8.3 and 8.4)
Figure 8.3 Wet mount microscopy (400x, phase contrast).
A– Vaginal atrophy B– Atrophic vaginitis
Figure 8.4 Gram stain (1000x, oil immersion), vaginal atrophy.
A– Vaginal atrophy B– Atrophic vaginitis
162
A vaginal pH >5.0 in the absence of other causes, such as infection or semen, is considered
an indicator of vaginal atrophy.3
8.7
Treatment
VVA can be treated with hormonal and non-hormonal therapies. Non-hormonal treatment
recommendations include vaginal lubricants and moisturizers, and continued sexual activi-
ty should be encouraged.22 In this section both hormonal and non-hormonal therapies will
be discussed.10 (Table 8.2)
TABLE 8.2 Recommendations for the management of vulvovaginal atrophy. Adapted from Pérez-
López et al. 2021.10
DHEA dehydroepiandrosterone, VMI vaginal maturation index, VVA vulvovaginal atrophy
Treatment Recommendation
1. Low dose and
ultralow dose vaginal
estrogens
Estradiol, conjugated equine estrogens, estriol, and promestriene are effective for
VVA, and without risk of endometrial or systemic effects.
2. Vaginal prasterone
Intravaginal prasterone reduces vaginal pH, improves VMI, and decreases dyspareu-
nia. Circulating levels of DHEA and its metabolites (testosterone and estradiol) remain
in the postmenopausal range in up to 52 weeks of use.
3. Systemic estrogens Should not be used for the sole purpose of treating vaginal atrophy; it is an option to con-
sider in women who also have vasomotor symptoms. Not always effective in treating VVA.
4. Vaginal
testosterone
Topical testosterone reduces vaginal pH and improves VMI and the number of lacto-
bacilli. Longer and larger studies are needed to assess safety and efficacy.
5. Lubricants and
moisturizers
Lubricants and moisturizers are appropriate for those women that cannot use or do
not want to receive hormone treatments.
6. Vaginal LASER
CO and erbium laser treatments have been reported in women with VVA, although there
is no clear evidence of the benefits as compared to hormone treatments. Currently, the
ISSVD does not endorse the use of these technologies out of the setting of clinical trials.
7. Radiofrequency
Intravaginal microablative radiofrequency has been suggested as a possible alterna-
tive treatment for VVA, but data are scarce. Currently, the ISSVD does not endorse the
use of these technologies out of the setting of clinical trials.
Vaginal lubricants and moisturizers
While less effective than hormonal treatments, some women and healthcare providers
prefer non-hormonal therapy as the first therapeutic approach to relieve the symptoms of
VVA.23 Non-hormonal approaches are particularly beneficial in women with contraindica-
tions to the use of hormones, or for those who prefer not to use them.24
Lubricants can be used before intercourse to reduce friction and discomfort during pene-
tration in sexual activity. They can be water-based, silicone, mineral oil, or herbal products
applied to the vagina and vulva and/or to the partner’s genitals. However, these products
are not effective in the treatment of the underlying causes of VVA.25
Moisturizers adhere to the vaginal mucosa, promoting rehydration and mimicking normal
163
lubrication. These products improve the integrity, elasticity, and flexibility of the tissue. They
must be used regularly (from daily to every three days). Moisturizers contain water and other
substances such as hyaluronic acid or polycarbophil.23, 26 Hyaluronic acid is a polymer found in
cartilage and other soft tissues in the body. In randomized clinical trials (RCTs) comparing hyalu-
ronic acid to placebo or vaginal estrogens, all were associated with a decrease in the severity of
dryness and dyspareunia (probably because the placebo had a lubricant effect). To date, there
is no evidence that products with hyaluronic acid have a greater benefit than nonhyaluronic
acid moisturizers.5 Studies with the use of moisturizers show improvement in vaginal dryness
and sexual function, as well as an improvement in the vaginal epithelium maturation. Despite
some mild irritation associated with its use, no serious adverse events have been reported.26, 27
Estrogen (systemic and vaginal) and selective estrogen
receptor modulator therapy
Vaginal estrogens are effective for the management of VVA. Several low-dose formulations
are available: creams, pessaries, tablets, and vaginal rings. Available active ingredient op-
tions include promestriene, estradiol, conjugated estrogens, and estriol.19, 21 The absorption
is variable depending on the degree of VVA, but plasma estrogen levels do not exceed the
normal postmenopausal range.2, 19, 28 Topical vaginal estrogens should be started with a
nightly application for two to three weeks and are later reduced to two to three times a
week, depending on the degree of atrophy. Women should be warned about a potential
burning sensation during the first weeks and that maximum effect may take up to eight
weeks to be reached.
Promestriene (3-propyl 17β -methyl diether estradiol) is a synthetic estrogen that is used
vaginally in a 1% cream formulation, which appears to have intramucosal effects only and
has been tested on women with gynecologic cancer. However, despite the promising re-
sults, larger and longer studies in relation to long-term safety are lacking.29, 30
The ultra-low-dose concentration of estriol formulations (vaginal gel containing 50 g/gram
of estriol or 30 g associated with L. crispatus, formulated in vaginal pills) significantly im-
prove both the VMI and pH when compared to placebo after 12 weeks.31, 32 The same findings
were confirmed in a double-blind randomized clinical trial comparing the use of 200 g and
30 g estriol pessaries: at 12 weeks, VMI and pH similarly and significantly improved. Ad-
verse events were rare and similar among all the groups.33
A Cochrane systematic review evaluated randomized controlled trials comparing vaginal
estrogens vs. placebo over 12 weeks for the treatment of VVA. The authors concluded that
there were no substantial differences in the effects of the different options. However, en-
dometrial thickness was increased in women who received estrogen cream compared to
those who wore rings — likely due to exposure to a higher dose in the former. There were
no differences in this aspect between users of pills or creams.34
Biehl et al. published a systematic review of 53 RCTs reporting on the efficacy and safety of
different vaginal estrogens used for GSM. Compared to placebo, all vaginal estrogens, re-
gardless of dosages and formulations, were superior in objective and subjective outcomes.
164
They also showed superiority over lubricants and moisturizers for improving objective, but
not subjective, clinical outcomes. Doses as low as 4 g have been shown to be effective. In
a review of studies of one year of treatment with vaginal estrogen, the complication rate
was overall low: vulvovaginal mycosis (0.73%), vaginal bleeding (0.75%), endometrial hy-
perplasia (0.06%), and there was one case of endometrial cancer (out of more than 4,500
women).35 Another systematic review of 20 RCTs on the use of vaginal estrogen alone for
12 to 52 weeks in postmenopausal women showed that the rate of endometrial cancer and
hyperplasia was 0.03 and 0.4%, respectively.36 Finally, treatment with vaginal estrogen in
women not exposed to menopausal systemic hormone therapy for more than 18 years has
shown that the risk of cardiovascular disease, cancer, and hip fracture is similar to that of
non-vaginal estrogen users.37
Systemic estrogen therapies are also available for patients with vasomotor symptoms. How-
ever, risks and benefits should be discussed. Systemic estrogen therapy should be used to-
gether with progestogens for women with an intact uterus, or alone after hysterectomy.38
This option can be tried in women suffering from vaginal atrophy and concomitant vaso-
motor symptoms.39 However, for some women, systemic hormone therapy is insufficient
and further require local therapy. The Women’s Health Initiative found that 74% of patients
reported improvement after one year of systemic hormone therapy.40 The fact that up to 1/4
of women using systemic hormonal therapy continue to experience symptoms of urogen-
ital atrophy is sufficient reason to justify not recommending systemic hormonal therapy in
women with vaginal symptoms only; many women initially require a combination of sys-
temic and local estrogen therapy, especially when it is used at low doses.41, 42
Ospemifene is an oral tissue-selective estrogen receptor modulator (SERM).43 It has antag-
onistic-antiestrogenic effects on the breast.44 Since the ER-beta is significantly reduced in
postmenopausal women, ospemifene seems to act on the ER-alpha.6 It does not increase
the risk of endometrial hyperplasia or thrombosis but improves bone density.44 Studies
show that ospemifene improves the VMI, vaginal pH, and decreases vaginal dryness, as well
as dyspareunia.44 Being an oral drug, it avoids local discomfort related to excipients of vagi-
nal drug delivery systems and can be considered in women with a history of breast cancer.45
Vaginal androgen (testosterone) therapy
Intravaginal testosterone has been studied in short-term interventions (4-12 weeks). Sys-
temic absorption of a single intravaginal dose of 2 mg in a double-blind, placebo-controlled
study in premenopausal women resulted in supraphysiological serum testosterone levels,
while there were no changes in estradiol.46 In a randomized study of women between 40-70
years of age comparing vaginal treatment with conjugated estrogen, testosterone, or place-
bo (glycerin lubricant), applied three times a week for 12 weeks, it was shown that hormone
treatments reduced the pH to <5 and increased the VMI as well as the number of lactobacilli. In
addition, there was no significant difference in serum hormone levels between hormone treat-
ments and placebo; there was also no difference in endometrial thickness among the groups.47
However, longer and better studies to evaluate safety and efficacy are needed before the
use of vaginal testosterone can be recommended.10, 48
165
Vaginal dehydroepiandrosterone (prasterone) therapy
Dehydroepiandrosterone (DHEA; prasterone) is converted to estradiol and testosterone in
the vaginal epithelium. It is an alternative to estrogens, which is administered vaginally at a
daily dose of 6.5 mg with no reported risks of cancer, although there are no long-term stud-
ies. Compared to placebo, vaginal prasterone for 12 weeks was associated with improve-
ment in dyspareunia, pH, and vaginal maturation, and there were no endometrial chang-
es.49 Vaginal dryness and discharge, vaginal epithelium thickness, and color improved while
circulating steroid levels remained within the normal range for postmenopausal women.50
Despite the paucity of data, it can be considered in women with an history of breast cancer.51
Vaginal LASER
Physical methods such as LASER in non-ablative, ablative, and microablative forms have
been used for skin “rejuvenation” on the face, neck, and body. Fractional LASER is also used
in the vaginal mucosa, allegedly promoting neocollagenesis and neoelastogenesis.52-54
LASER purportedly induces morphological changes in vaginal tissue, leading to relief from
vaginal dryness and dyspareunia.24 The application of microablative fractional LASER has
generated controversial opinions due to being based on poor studies and because of the
use outside of the released or approved intended uses.55
The two main types of LASERS available are the microablative fractional carbon dioxide
(CO) LASER and the non-ablative vaginal erbium:YAG LASER. Regarding the CO LASER, it
is hypothesized that the thermal energy deposited in the vaginal wall stimulates neovascu-
larization, promotes collagen synthesis, and improves natural lubrication and leads to a sig-
nificant improvement in vaginal health.50, 52 Cruz et al. compared three arms: fractional CO
LASER, topical estriol, and CO LASER with estriol for 20 weeks. The combined LASER and es-
trogen treatment showed the most significant change in the Vaginal Health Index (VHI), and
both the LASER treatment alone arm and the combined treatment demonstrated significant
improvement in dyspareunia, burning, and dryness when compared to the estrogen group.
Importantly, in the LASER treatment alone arm there was an increase in pain. However, this
study had some limitations, including that it was designed to detect differences only in the
VHI and not in the other parameters.56
In 2021, a RCT comparing the effect of fractional CO LASER versus sham treatment on vaginal
symptom severity was conducted. Of the 85 randomized participants (mean age, 57 years),
78 (91.7%) completed the 12-month follow-up. From baseline to 12 months, there was no
significant difference between the CO LASER and the sham treatment groups concerning
symptom´s severity, quality of life score, VHI or histology. There were 16 adverse events in the
LASER group and 17 in the sham group, including vaginal pain/discomfort, spotting, discharge,
and lower urinary tract symptoms. No severe adverse events were reported in either group.57
A recent study, on an ewe model, showed that CO LASER effect in histological terms was
similar to that of sham treatment, contrarily to what was noticed in the estrogen arm.58 The
same group, in a well-designed RCT, showed that there was similar improvement in terms of
the most bothersome symptom in the LASER and placebo arm, highlighting the significant
placebo effect, as well as that of mechanical manipulation.59
166
The recently available data, combined with the high placebo effect expected in treatments
to improve sexual function, sustains the recommendation issued by the International Soci-
ety for the Study of Vulvovaginal Disease (ISSVD) in 2019, that vaginal LASERs should not be
used out of the setting of clinical trials.55, 60
Vaginal radiofrequency
Radiofrequency is performed by cutting and/or coagulating biological tissues, using a high-
frequency alternating current, which instantly raises the cell temperature up to 100°C, lead-
ing to the expansion and rupture of the cell membrane. Observational studies have shown
an apparent change in pH, increase in lactobacilli, VMI and VHI. These studies suffered from
several limitations including the small number of enrolled participants, the lack of a control
arm, and the short follow-up.53, 61
The available data are insufficient to demonstrate efficacy and safety as an alternative to
hormonal treatments.55, 62 Similar to LASERs, there is lack of studies including objectives
and standardized measurable results, as well as follow-up of short- and long-term adverse
effects.63 Currently, the ISSVD does not endorse the use of these technologies outside the
clinical trial setting.55
Pelvic floor rehabilitation
Pelvic floor physiotherapy with muscle training significantly reduces VVA in postmeno-
pausal women. Mercier et al. showed that a 12-week program, oriented and monitored by
physical therapists increases vaginal wall lubrication, thickens the vaginal epithelial surface,
and improves the vaginal mucosa color.64 Pelvic floor rehabilitation has also been used com-
bined with intravaginal estriol for six months and compared to an estriol arm only.65 This
approach has also been examined along with the addition of L. acidophilus and showed that
triple therapy (L. acidophilus, estriol and pelvic floor rehabilitation) was effective and could
be considered as first-line treatment for symptoms of urogenital aging in postmenopausal
women.66
8.8
Special situations (postpartum/breastfeeding, breast cancer)
In transient postpartum/breastfeeding situations due to the increase in prolactin, and the
consequent blockage of the hypothalamic-pituitary-ovarian hormonal axis, it is not uncom-
mon for women to experience transient hypoestrogenism and VVA. The issue should be ad-
dressed with women who, if symptomatic, may opt to be treated with a similar approach to
postmenopausal women.
Women diagnosed with breast cancer may experience early menopause or worsening
symptoms if already post-menopausal, due to chemotherapy, radiotherapy, and/or endo-
crine treatments. In breast cancer survivors, estrogens are usually avoided as they may pose
a theoretical risk of cancer recurrence, possible interference with tamoxifen or aromatase
167
inhibitors, or fear of a patient lawsuit against the physician.67 Moisturizers and lubricants are
first-line therapy. The data on the safety of the use of vaginal estrogens in women treated
with aromatase inhibitors is contradictory.68, 69 Due to tamoxifen’s receptor-blocking action,
the use of vaginal estrogen therapies may be safer than in women treated with aromatase
inhibitors.10
Some studies have addressed the safety of using ultra-low doses of vaginal estriol in women
with a history of breast cancer, showing that despite an initial transient elevation, the sys-
temic levels remain within the normal post-menopausal range.70 A meta-analysis reported
the safety of vaginal estrogen application in women with breast cancer receiving aromatase
inhibitors.71 There were no changes in serum LH and estradiol levels, while FSH almost dou-
bled compared to baseline levels. Therefore, it can be assumed that vaginal estrogens are
not significantly absorbed, which is indirect evidence of safety. Of note, the efficacy of vag-
inal estrogens in women receiving aromatase inhibitors is not confirmed in all studies.72 In
another study it was shown that the estradiol ring (7.5 g/d) was effective when compared
to testosterone.73 Caution should be exercised when prescribing hormone treatments in pa-
tients with hormone-dependent cancer, as transient elevations of estradiol have been re-
ported in women with breast cancer on aromatase inhibitors who received vaginal estradiol
or testosterone.74, 75
Prasterone has been studied as a treatment for GSM in cancer survivors; the limited data
available have demonstrated improved vaginal symptoms at 12 weeks.74
In general, longer and larger studies are needed to assess safety and efficacy of vaginal hor-
mone treatments in women with a history of breast cancer.
8.9
Future perspectives
The therapeutic management of VVA should follow a sequential order, taking into consider-
ation the woman’s age, preferences, symptoms, and general health status, as well as previ-
ous treatments. Systemic hormone therapy should only be used to treat VVA in women with
other menopausal symptoms and without contraindications. Lifestyle, comorbidities, and
chronic diseases can also influence the choice of treatment. Vaginal options that produce
benefits for VVA include lubricants and moisturizers, estrogens (estradiol, estriol, prome-
striene), or prasterone. Although LASER and radiofrequency procedures are currently used,
the ISSVD does not currently endorse their use outside the clinical trial setting, due to the
lack of evidence on safety and efficacy.55
There are significant limitations in publications on VVA and related issues, including hetero-
geneity of outcomes, with the available evidence being based on short-term interventions
and small samples. Another relevant issue is that the ages of the population studied correspond
to young postmenopausal women, and VVA is a progressive phenomenon requiring specific infor-
mation related to treatments in women over 65 years of age. In ad dition , sexu al nee ds and prac-
tices change with age, and partner capacities should also be considered in future studies.
168
Recommendations
Recommendation Quality of
evidence
Strength of
recommendation
The diagnosis of vaginal atrophy is clinical. 5 D
Wet mount microscopy can be used to confirm the diagnosis of
vaginal atrophy. 3b C
A pH>5 in the absence of semen, infection or use of vaginal medi-
cation is suggestive of vaginal atrophy. 2b B
Vaginal lubricants and moisturizers are particularly beneficial in
women with contraindications to the use of hormones, or for those
who prefer not to use them.
2a B
Vaginal lubricants and moisturizers are the first line choice for
vaginal atrophy in women with breast cancer. 5 D
Topical vaginal estrogens should be started with a nightly applica-
tion for 2 to 3 weeks and are later reduced to 2 to 3 times a week. 2b B
Systemic estrogens can be used in women without contraindica-
tion and who have vaginal atrophy and vasomotor symptoms. 2a B
Ultralow dose vaginal estriol may be safe in women with breast
cancer who are taking tamoxifen. 3a C
Ospemifen may be an option in women who prefer an oral option. 5 D
Ospemifen can be considered in women with a history of breast
cancer. 4 C
The data on vaginal testosterone is too limited to allow recom-
mending it to treat vaginal atrophy. 4 C
Prasterone can be used to treat vaginal atrophy. 2a B
The available data do not allow recommending LASER for the
treatment of vaginal atrophy. 2a B
The available data do not allow recommending radiofrequency for
the treatment of vaginal atrophy. 2b B
Pelvic floor physiotherapy with muscle training can be recom-
mended for vaginal atrophy. 4 C
References
1. Pérez-López, F. R.; Vieira-Baptista, P.; Phillips, N.; Cohen-Sacher, B.; Fialho, S.; Stockdale, C. K., Clinical manifestations
and evaluation of postmenopausal vulvovaginal atrophy. Gynecol Endocrinol 2021, 37, (8), 740-745.
2. Portman, D. J.; Gass, M. L., Genitourinary syndrome of menopause: new terminology for vulvovaginal atrophy from
the International Society for the Study of Women’s Sexual Health and the North American Menopause Society.
Menopause 2014, 21, (10), 1063-8.
3. Vieira-Baptista, P.; Grincevičienė, Š.; Oliveira, C.; Fonseca-Moutinho, J.; Cherey, F.; Stockdale, C. K., The International
Society for the Study of Vulvovaginal Disease Vaginal Wet Mount Microscopy Guidelines: How to Perform, Applica-
tions, and Interpretation. J Low Genit Tract Dis 2021, 25, (2), 172-180.
4. Palma, F.; Volpe, A.; Villa, P.; Cagnacci, A., Vaginal atrophy of women in postmenopause. Results from a multicentric
observational study: The AGATA study. Maturitas 2016, 83, 40-4.
169
5. The 2020 genitourinary syndrome of menopause position statement of The North American Menopause Society.
Menopause 2020, 27, (9), 976-992.
6. Gebhart, J. B.; Rickard, D. J.; Barrett, T. J.; Lesnick, T. G.; Webb, M. J.; Podratz, K. C.; Spelsberg, T. C., Expression of es-
trogen receptor isoforms alpha and beta messenger RNA in vaginal tissue of premenopausal and postmenopausal
women. Am J Obstet Gynecol 2001, 185, (6), 1325-30; discussion 1330-1.
7. Labrie, F.; Archer, D. F.; Martel, C.; Vaillancourt, M.; Montesino, M., Combined data of intravaginal prasterone against
vulvovaginal atrophy of menopause. Menopause 2017, 24, (11), 1246-1256.
8. Traish, A. M.; Vignozzi, L.; Simon, J. A.; Goldstein, I.; Kim, N. N., Role of Androgens in Female Genitourinary Tissue Struc-
ture and Function: Implications in the Genitourinary Syndrome of Menopause. Sex Med Rev 2018, 6, (4), 558-571.
9. Maseroli, E.; Vignozzi, L., Testosterone and Vaginal Function. Sex Med Rev 2020, 8, (3), 379-392.
10. Pérez-López, F. R.; Phillips, N.; Vieira-Baptista, P.; Cohen-Sacher, B.; Fialho, S.; Stockdale, C. K., Management of post-
menopausal vulvovaginal atrophy: recommendations of the International Society for the Study of Vulvovaginal
Disease. Gynecol Endocrinol 2021, 37, (8), 746-752.
11. Bertin, J.; Dury, A. Y.; Ouellet, J.; Pelletier, G.; Labrie, F., Localization of the androgen-synthesizing enzymes, androgen
receptor, and sex steroids in the vagina: possible implications for the treatment of postmenopausal sexual dysfunc-
tion. J Sex Med 2014, 11, (8), 1949-61.
12. Verstraelen, H.; Vervaet, C.; Remon, J. P., Rationale and Safety Assessment of a Novel Intravaginal Drug-Delivery
System with Sustained DL-Lactic Acid Release, Intended for Long-Term Protection of the Vaginal Microbiome. PLoS
One 2016, 11, (4), e0153441.
13. Mirmonsef, P.; Hotton, A. L.; Gilbert, D.; Gioia, C. J.; Maric, D.; Hope, T. J.; Landay, A. L.; Spear, G. T., Glycogen Levels in Undi-
luted Genital Fluid and Their Relationship to Vaginal pH, Estrogen, and Progesterone. PLoS One 2016, 11, (4), e0153553.
14. Verstraelen, H.; Vieira-Baptista, P.; De Seta, F.; Ventolini, G.; Lonnee-Hoffmann, R.; Lev-Sagie, A., The Vaginal Micro-
biome: I. Research Development, Lexicon, Defining “Normal” and the Dynamics Throughout Women’s Lives. J Low
Genit Tract Dis 2022, 26, (1), 73-78.
15. Brotman, R. M.; Shardell, M. D.; Gajer, P.; Fadrosh, D.; Chang, K.; Silver, M. I.; Viscidi, R. P.; Burke, A. E.; Ravel, J.; Gravitt, P.
E., Association between the vaginal microbiota, menopause status, and signs of vulvovaginal atrophy. Menopause
2014, 21, (5), 450-8.
16. Ravel, J.; Gajer, P.; Abdo, Z.; Schneider, G. M.; Koenig, S. S.; McCulle, S. L.; Karlebach, S.; Gorle, R.; Russell, J.; Tacket, C.
O.; Brotman, R. M.; Davis, C. C.; Ault, K.; Peralta, L.; Forney, L. J., Vaginal microbiome of reproductive-age women. Proc
Natl Acad Sci U S A 2011, 108 Suppl 1, (Suppl 1), 4680-7.
17. Shen, J.; Song, N.; Williams, C. J.; Brown, C. J.; Yan, Z.; Xu, C.; Forney, L. J., Effects of low dose estrogen therapy on the
vaginal microbiomes of women with atrophic vaginitis. Sci Rep 2016, 6, 24380.
18. Archer, D. F., Efficacy and tolerability of local estrogen therapy for urogenital atrophy. Menopause 2010, 17, (1), 194-203.
19. Rees, M.; Pérez-López, F. R.; Ceasu, I.; Depypere, H.; Erel, T.; Lambrinoudaki, I.; Schenck-Gustafsson, K.; Simoncini, T.;
van der Schouw, Y.; Tremollieres, F., EMAS clinical guide: low-dose vaginal estrogens for postmenopausal vaginal
atrophy. Maturitas 2012, 73, (2), 171-4.
20. Sherrard, J.; Wilson, J.; Donders, G.; Mendling, W.; Jensen, J. S., 2018 European (IUSTI/WHO) International Union
against sexually transmitted infections (IUSTI) World Health Organisation (WHO) guideline on the management of
vaginal discharge. Int J STD AIDS 2018, 29, (13), 1258-1272.
21. Mac Bride, M. B.; Rhodes, D. J.; Shuster, L. T., Vulvovaginal atrophy. Mayo Clin Proc 2010, 85, (1), 87-94.
22. Paladine, H. L.; Desai, U. A., Vaginitis: Diagnosis and Treatment. Am Fam Physician 2018, 97, (5), 321-329.
23. Edwards, D.; Panay, N., Treating vulvovaginal atrophy/genitourinary syndrome of menopause: how important is
vaginal lubricant and moisturizer composition? Climacteric 2016, 19, (2), 151-61.
24. Kamilos, M. F.; Borrelli, C. L., New therapeutic option in genitourinary syndrome of menopause: pilot study using
microablative fractional radiofrequency. Einstein (Sao Paulo) 2017, 15, (4), 445-451.
25. Cunha, A. R.; Machado, R. M.; Palmeira-de-Oliveira, A.; Martinez-de-Oliveira, J.; das Neves, J.; Palmeira-de-Oliveira, R.,
Characterization of commercially available vaginal lubricants: a safety perspective. Pharmaceutics 2014, 6, (3), 530-42.
26. Chen, J.; Geng, L.; Song, X.; Li, H.; Giordan, N.; Liao, Q., Evaluation of the efficacy and safety of hyaluronic acid vaginal
gel to ease vaginal dryness: a multicenter, randomized, controlled, open-label, parallel-group, clinical trial. J Sex Med
2013, 10, (6), 1575-84.
27. Vale, F.; Rezende, C.; Raciclan, A.; Bretas, T.; Geber, S., Efficacy and safety of a non-hormonal intravaginal moisturizer
for the treatment of vaginal dryness in postmenopausal women with sexual dysfunction. Eur J Obstet Gynecol Re-
prod Biol 2019, 234, 92-95.
28. Phillips, N. A.; Bachmann, G. A., Genitourinary syndrome of menopause: Common problem, effective treatments.
Cleve Clin J Med 2018, 85, (5), 390-398.
170
29. Del Pup, L., Management of vaginal dryness and dyspareunia in estrogen sensitive cancer patients. Gynecol Endo-
crinol 2012, 28, (9), 740-5.
30. Del Pup, L.; Postruznik, D.; Corona, G., Effect of one-month treatment with vaginal promestriene on serum estrone
sulfate levels in cancer patients: a pilot study. Maturitas 2012, 72, (1), 93-4.
31. Cano, A.; Estévez, J.; Usandizaga, R.; Gallo, J. L.; Guinot, M.; Delgado, J. L.; Castellanos, E.; Moral, E.; Nieto, C.; del Prado,
J. M.; Ferrer, J., The therapeutic effect of a new ultra low concentration estriol gel formulation (0.005% estriol vaginal
gel) on symptoms and signs of postmenopausal vaginal atrophy: results from a pivotal phase III study. Menopause
2012, 19, (10), 1130-9.
32. Mueck, A. O.; Ruan, X.; Prasauskas, V.; Grob, P.; Or tmann, O., Treatment of vaginal atrophy with estriol and lactobacilli
combination: a clinical review. Climacteric 2018, 21, (2), 140-147.
33. Griesser, H.; Skonietzki, S.; Fischer, T.; Fielder, K.; Suesskind, M., Low dose estriol pessaries for the treatment of vag-
inal atrophy: a double-blind placebo-controlled trial investigating the efficacy of pessaries containing 0.2mg and
0.03mg estriol. Maturitas 2012, 71, (4), 360-8.
34. Lethaby, A.; Ayeleke, R. O.; Roberts, H., Local oestrogen for vaginal atrophy in postmenopausal women. Cochrane
Database Syst Rev 2016, 2016, (8), Cd001500.
35. Biehl, C.; Plotsker, O.; Mirkin, S., A systematic review of the efficacy and safety of vaginal estrogen products for the
treatment of genitourinary syndrome of menopause. Menopause 2019, 26, (4), 431-453.
36. Constantine, G. D.; Bruyniks, N.; Princic, N.; Huse, D.; Palmer, L.; Lenhart, G.; Blumentals, W. A.; Nappi, R. E., Incidence of
genitourinary conditions in women with a diagnosis of vulvar/vaginal atrophy. Curr Med Res Opin 2014, 30, (1), 143-8.
37. Bumphenkiatikul, T.; Panyakhamlerd, K.; Chatsuwan, T.; Ariyasriwatana, C.; Suwan, A.; Taweepolcharoen, C.; Taechakra-
ichana, N., Effects of vaginal administration of conjugated estrogens tablet on sexual function in postmenopausal wom-
en with sexual dysfunction: a double-blind, randomized, placebo-controlled trial. BMC Womens Health 2020, 20, (1), 173.
38. NICE NICE: Menopause, Diagnosis and Management – from Guideline to Practice Guideline Summary. https://
thebms.org.uk/wp-content/uploads/2019/04/09-BMS-TfC-NICE-Menopause-Diagnosis-and-Management-from-
Guideline-to-Practice-Guideline-Summary-01-April2019.pdf (11/07/2022),
39. The 2017 hormone therapy position statement of The North American Menopause Society. Menopause 2018, 25,
(11), 1362-1387.
40. Barnabei, V. M.; Cochrane, B. B.; Aragaki, A. K.; Nygaard, I.; Williams, R. S.; McGovern, P. G.; Young, R. L.; Wells, E. C.;
O’Sullivan, M. J.; Chen, B.; Schenken, R.; Johnson, S. R., Menopausal symptoms and treatment-related effects of
estrogen and progestin in the Women’s Health Initiative. Obstet Gynecol 2005, 105, (5 Pt 1), 1063-73.
41. Palacios, S.; Mejía, A.; Neyro, J. L., Treatment of the genitourinary syndrome of menopause. Climacteric 2015, 18
Suppl 1, 23-9.
42. Sturdee, D. W.; Panay, N., Recommendations for the management of postmenopausal vaginal atrophy. Climacteric
2010, 13, (6), 509-22.
43. Del Pup, L., Ospemifene: a safe treatment of vaginal atrophy. Eur Rev M ed Pharmacol Sci 2016, 20, (18), 3934-3944.
44. Berga, S. L., Profile of ospemifene in the breast. Reprod Sci 2013, 20, (10), 1130-6.
45. Lilue, M.; Palacios, S.; Del Carmen Pingarrón Santofimia, M., Experience with ospemifene in patients with vulvar and
vaginal atrophy and a history of breast cancer: case studies. Drugs Context 2020, 9.
46. Apperloo, M.; Midden, M.; van der Stege, J.; Wouda, J.; Hoek, A.; Weijmar Schultz, W., Vaginal application of testos-
terone: A study on pharmacokinetics and the sexual response in healthy volunteers. J Sex Med 2006, 3, (3), 541-9.
47. Fernandes, T.; Pedro, A. O.; Baccaro, L. F.; Costa-Paiva, L. H., Hormonal, metabolic, and endometrial safety of testos-
terone vaginal cream versus estrogens for the treatment of vulvovaginal atrophy in postmenopausal women: a
randomized, placebo-controlled study. Menopause 2018, 25, (6), 641-647.
48. Simon, J. A.; Goldstein, I.; Kim, N. N.; Davis, S. R.; Kellogg-Spadt, S.; Lowenstein, L.; Pinkerton, J. V.; Stuenkel, C. A.;
Traish, A. M.; Archer, D. F.; Bachmann, G.; Goldstein, A. T.; Nappi, R. E.; Vignozzi, L., The role of androgens in the
treatment of genitourinary syndrome of menopause (GSM): International Society for the Study of Women’s Sexual
Health (ISSWSH) expert consensus panel review. Menopause 2018, 25, (7), 837-847.
49. Archer, D. F.; Labrie, F.; Bouchard, C.; Portman, D. J.; Koltun, W.; Cusan, L.; Labrie, C.; Côté, I.; Lavoie, L.; Martel, C.;
Balser, J., Treatment of pain at sexual activity (dyspareunia) with intravaginal dehydroepiandrosterone (prasterone).
Menopause 2015, 22, (9), 950-63.
50. Labrie, F.; Archer, D. F.; Koltun, W.; Vachon, A.; Young, D.; Frenette, L.; Portman, D.; Montesino, M.; Côté, I.; Parent, J.;
Lavoie, L.; AB, B. S.; Martel, C.; Vaillancourt, M.; Balser, J.; Moyneur, É., Efficacy of intravaginal dehydroepiandrosterone
(DHEA) on moderate to severe dyspareunia and vaginal dryness, symptoms of vulvovaginal atrophy, and of the
genitourinary syndrome of menopause. Menopause 2018, 25, (11), 1339-1353.
171
51. Sussman, T. A.; Kruse, M. L.; Thacker, H. L.; Abraham, J., Managing Genitourinary Syndrome of Menopause in Breast
Cancer Survivors Receiving Endocrine Therapy. J Oncol Pract 2019, 15, (7), 363-370.
52. Bretas, T. L. B.; Issa, M. C. A.; Fialho, S.; Villar, E. A. G.; Velarde, L. G. C.; Pérez-López, F. R., Vaginal collagen I and III chang-
es after carbon dioxide laser application in postmenopausal women with the genitourinary syndrome: a pilot study.
Climacteric 2022, 25, (2), 186-194.
53. Sarmento, A. C.; Fernandes, F. S.; Marconi, C.; Giraldo, P. C.; Eleutério-Júnior, J.; Crispim, J. C.; Gonçalves, A. K., Impact
of microablative fractional radiofrequency on the vaginal health, microbiota, and cellularity of postmenopausal
women. Clinics (Sao Paulo) 2020, 75, e1750.
54. Salvatore, S.; Pitsouni, E.; Grigoriadis, T.; Zacharakis, D.; Pantaleo, G.; Candiani, M.; Athanasiou, S., CO(2) laser and
the genitourinary syndrome of menopause: a randomized sham-controlled trial. Climacteric 2021, 24, (2), 187-193.
55. Preti, M.; Vieira-Baptista, P.; Digesu, G. A.; Bretschneider, C. E.; Damaser, M.; Demirkesen, O.; Heller, D. S.; Mangir, N.;
Marchitelli, C.; Mourad, S.; Moyal-Barracco, M.; Peremateu, S.; Tailor, V.; Tarcan, T.; De, E. J. B.; Stockdale, C. K., The Clin-
ical Role of LASER for Vulvar and Vaginal Treatments in Gynecology and Female Urology: An ICS/ISSVD Best Practice
Consensus Document. J Low Genit Tract Dis 2019, 23, (2), 151-160.
56. Cruz, V. L.; Steiner, M. L.; Pompei, L. M.; Strufaldi, R.; Fonseca, F. L. A.; Santiago, L. H. S.; Wajsfeld, T.; Fernandes, C. E., Ran-
domized, double-blind, placebo-controlled clinical trial for evaluating the efficacy of fractional CO2 laser compared
with topical estriol in the treatment of vaginal atrophy in postmenopausal women. Menopause 2018, 25, (1), 21-28.
57. Li, F. G.; Maheux-Lacroix, S.; Deans, R.; Nesbitt-Hawes, E.; Budden, A.; Nguyen, K.; Lim, C. Y.; Song, S.; McCormack, L.; Ly-
ons, S. D.; Segelov, E.; Abbott, J. A., Effect of Fractional Carbon Dioxide Laser vs Sham Treatment on Symptom Severity
in Women With Postmenopausal Vaginal Symptoms: A Randomized Clinical Trial. JAMA 2021, 326, (14), 1381-1389.
58. Mackova, K.; Mazzer, A. M.; Mori Da Cunha, M.; Hajkova Hympanova, L.; Urbankova, I.; Kastelein, A. W.; Vodegel, E.;
Vander Linden, K.; Fehervary, H.; Guler, Z.; Roovers, J. P.; Krofta, L.; Verhaeghe, J.; Deprest, J., Vaginal Er:YAG laser applica-
tion in the menopausal ewe model: a randomised estrogen and sham-controlled trial. Bjog 2021, 128, (6), 1087-1096.
59. Page, A. S.; Verbakel, J. Y.; Verhaeghe, J.; Latul, Y. P.; Housmans, S.; Deprest, J., Laser versus sham for genitourinary
syndrome of menopause: A randomised controlled trial. Bjog 2022.
60. Pérez-López, F. R.; Varikasuvu, S. R., Vulvovaginal atrophy management with a laser: the placebo effect or the condi-
tioning Pavlov reflex. Climacteric 2022, 25, (4), 323-326.
61. Juhász, M. L. W.; Kor ta, D. Z.; Mesinkovska, N. A., Vaginal Rejuvenation: A Retrospective Review of Lasers and Radiof-
requency Devices. Dermatol Surg 2021, 47, (4), 489-494.
62. Pitsouni, E.; Grigoriadis, T.; Douskos, A.; Kyriakidou, M.; Falagas, M. E.; Athanasiou, S., Efficacy of vaginal therapies
alternative to vaginal estrogens on sexual function and orgasm of menopausal women: A systematic review and
meta-analysis of randomized controlled trials. Eur J Obstet Gynecol Reprod Biol 2018, 229, 45-56.
63. Vieira-Baptista, P., Better studies needed on LASER use in urinary incontinence. Bjog 2020, 127, (11), 1347.
64. Mercier, J.; Morin, M.; Zaki, D.; Reichetzer, B.; Lemieux, M. C.; Khalifé, S.; Dumoulin, C., Pelvic floor muscle training as
a treatment for genitourinary syndrome of menopause: A single-arm feasibility study. Maturitas 2019, 125, 57-62.
65. Capobianco, G.; Donolo, E.; Borghero, G.; Dessole, F.; Cherchi, P. L.; Dessole, S., Effects of intravaginal estriol and pelvic
floor rehabilitation on urogenital aging in postmenopausal women. Arch Gynecol Obstet 2012, 285, (2), 397-403.
66. Capobianco, G.; Wenger, J. M.; Meloni, G. B.; Dessole, M.; Cherchi, P. L.; Dessole, S., Triple therapy with Lactobacilli
acidophili, estriol plus pelvic floor rehabilitation for symptoms of urogenital aging in postmenopausal women. Arch
Gynecol Obstet 2014, 289, (3), 601-8.
67. ACOG Committee Opinion No. 659: The Use of Vaginal Estrogen in Women With a History of Estrogen-Dependent
Breast Cancer. Obstet Gynecol 2016, 127, (3), e93-e96.
68. Cold, S.; Cold, F.; Jensen, M. B.; Cronin-Fenton, D.; Christiansen, P.; Ejlertsen, B., Systemic or Vaginal Hormone Therapy
After Early Breast Cancer: A Danish Observational Cohort Study. J Natl Cancer Inst 2022, 114, (10), 1347-1354.
69. Pavlović, R. T.; Janković, S. M.; Milovanović, J. R.; Stefanović, S. M.; Folić, M. M.; Milovanović, O. Z.; Mamillapalli, C.;
Milosavljević, M. N., The Safety of Local Hormonal Treatment for Vulvovaginal Atrophy in Women With Estrogen Re-
ceptor-positive Breast Cancer Who Are on Adjuvant Aromatase Inhibitor Therapy: Meta-analysis. Clin Breast Cancer
2019, 19, (6), e731-e740.
70. Sánchez-Rovira, P.; Hirschberg, A. L.; Gil-Gil, M.; Bermejo-De Las Heras, B.; Nieto-Magro, C., A Phase II Prospective,
Randomized, Double-Blind, Placebo-Controlled and Multicenter Clinical Trial to Assess the Safety of 0.005% Estriol
Vaginal Gel in Hormone Receptor-Positive Postmenopausal Women with Early Stage Breast Cancer in Treatment
with Aromatase Inhibitor in the Adjuvant Setting. Oncologist 2020, 25, (12), e1846-1854.
71. Hirschberg, A. L.; Sánchez-Rovira, P.; Presa-Lorite, J.; Campos-Delgado, M.; Gil-Gil, M.; Lidbrink, E.; Suárez-Almarza, J.;
Nieto-Magro, C., Efficacy and safety of ultra-low dose 0.005% estriol vaginal gel for the treatment of vulvovaginal
atrophy in postmenopausal women with early breast cancer treated with nonsteroidal aromatase inhibitors: a phase
II, randomized, double-blind, placebo-controlled trial. Menopause 2020, 27, (5), 526-534.
172
72. Jain, A. L.; Jamy, O.; Mullins, J.; Usman, R. M.; Hare, F.; Valasareddy, P.; Chaudhry, A.; Ryder, J.; Smith, J. R.; Miller, E.;
Ranganath, H.; Schwartzberg, L.; Stepanski, E.; Walker, M.; Gatwood, J.; Vidal, G. A., Usefulness of patient-reported
outcomes to assess the effectiveness of topical hormonal therapy for gynecologic symptoms after antihormonal
treatment for breast cancer. Proc (Bayl Univ Med Cent) 2020, 33, (3), 331-335.
73. Melisko, M. E.; Goldman, M. E.; Hwang, J.; De Luca, A.; Fang, S.; Esserman, L. J.; Chien, A. J.; Park, J. W.; Rugo, H. S.,
Vaginal Testosterone Cream vs Estradiol Vaginal Ring for Vaginal Dryness or Decreased Libido in Women Receiving
Aromatase Inhibitors for Early-Stage Breast Cancer: A Randomized Clinical Trial. JAMA Oncol 2017, 3, (3), 313-319.
74. Barton, D. L.; Sloan, J. A.; Shuster, L. T.; Gill, P.; Griffin, P.; Flynn, K.; Terstriep, S. A.; Rana, F. N.; Dockter, T.; Atherton, P. J.;
Tsai, M.; Sturtz, K.; Lafky, J. M.; Riepl, M.; Thielen, J.; Loprinzi, C. L., Evaluating the efficacy of vaginal dehydroepian-
dosterone for vaginal symptoms in postmenopausal cancer survivors: NCCTG N10C1 (Alliance). Support Care Cancer
2018, 26, (2), 643-650.
75. Reeder-Hayes, K.; Muss, H. B., Vaginal Estrogens and Aromatase Inhibitors: How Safe Is Safe Enough? JAMA Oncol
2017, 3, (3), 305-306.
173
9.1
Introduction
Vulvovaginitis is the most common gynecological problem in prepubertal females. Many
providers classify vaginitis under the general heading of vulvovaginitis and include causes
of vulvitis. Most causes of vaginitis are different from causes of vulvitis, which are primar-
ily skin diseases. Vulvitis is not uncommon in children and can be due to dermatological
conditions such as dermatitis which can produce scaling and weeping, psoriasis and lichen
sclerosus (LS) which can be mistaken for discharge.
This section addresses vaginitis, which is uncommon in the prepubertal female and, except
for infectious causes, there is little research and data. Although most parents and medical
personnel think of vaginal discharge and vaginitis as being of infectious origin, especially
yeast, these are nearly non-existent in the healthy child, and etiologies other than infection
should be considered.
9.2
The vagina in the prepubertal child
The vestibule
Erythema of the vestibule is common in children and, in the absence of symptoms, it is of
no significance. If there is an associated discharge, vaginitis should be considered. Incon-
tinence in children, especially in babies, is far from being uncommon, causing urinary and
stool leakage to be mistaken for vaginal discharge. Pooling of urine and stool in the vesti-
bule may result in dysuria and irritation. This may be the result of abnormal voiding posture.1
The hymen
With very rare exceptions, all girls are born with a hymen, the shape and appearance of
which is highly variable.
9
VAGINITIS IN CHILDREN
(alphabetical order)
Libby Edwards
Gayle Fischer
Caroline Oliveira
Henrique Soares
174
At birth, as a result of exposure to maternal oestrogen, the hymen is thickened, returning to
a thin translucent membrane over the next two years of life.
In some cases the hymen may be imperforate, requiring surgical treatment at puberty.2
The vagina and vaginal discharge
Data on the normal vagina in children are limited. However, as in adults, vaginal ridges and
columns are normal variants. Peri-urethral and peri-hymenal fibrous bands may form a
pocket on either side of the urethra or hymen.
The linea vestibularis is a normal variant described as an avascular area of the posterior ves-
tibule which appears pale and may be confused with scarring.
In utero, the vaginal epithelium of the fetus is stimulated by maternal hormones that cross
the placenta into the fetal circulation. After delivery, these hormone levels fall rapidly, and
a thick, greyish-white, mucoid discharge from the neonates vagina can be observed. The
discharge usually resolves in 10 days. In some baby girls, the discharge from the vagina is
blood-tinged or even grossly bloody. This is a physiological endometrial response to the
drop of maternal estrogens after birth.3
The normal length of the vagina in a newborn is 4 cm with a long cervix which is larger than
the uterine corpus. In childhood, the vaginal length increases to about 8 cm. In the neonate,
a vaginal vault smear shows polygonal epithelial cells and in a prepubertal child the epithe-
lial cells are round (parabasal cells).4-8 (Figure 9.1)
Figure 9.1 Wet mount microscopy (200x) from a prepubertal girls’ vagina.
A– Exclusive presence of parabasal cells; lactobacilli absent B– Presence of inflammation in a case of bacterial vaginitis
Lactobacilli are typically absent from the prepubertal vagina, and the pH is usually higher
than in adult women.
175
The lack of estrogens in prepubertal girls results in remarkable differences compared to the
normal vaginal microbiota of post pubertal women and, by extension, to the organisms
that produce vaginitis in children.9 Normal bacteria in the prepuberal vagina include enteric
microorganisms such as Diphtheroids spp., Peptococcus spp., Bacteroides spp., Proteus spp.,
and Escherichia coli, as well as those of respiratory origins such as group A Streptococcus and
other streptococci, Haemophilus influenzae, and Klebsiella spp..10 Other microorganisms such
as Staphylococcus aureus and S. epidermidis, Ureaplasma urealyticum, Gardnerella spp., Lacto-
bacillus spp., and Candida spp. are sometimes found in asymptomatic children.10 Generally,
data regarding both normal and pathogenic vaginal organisms in prepubertal children have
not been categorized according to age or Tanner stage. Estrogen effects begin to occur at
about seven years of age, so lumping all ages together is less than ideal.11
In the prepubertal child there is minimal vaginal discharge. However, an inoffensive, thin
milky or greenish discharge which is not associated with symptoms or clinically obvious
inflammation is common and harmless.
As puberty approaches, up to three years prior to menarche, a milky discharge is normal, the
pH becomes lower and lactobacilli and Gardnerella spp. appear.11-15
9.3
How to conduct a vaginal examination in a child
There exists literature with recommendations for conducting a genital exam in a child and
all stress the importance of creating a non-threatening environment, and distraction tech-
niques to gain the child’s and parents’ trust. Each clinician will have their own method of
facilitating this. Small children may be most comfortable remaining on their carer’s lap, but
options of either doing this or lying on the exam table should be offered.
Prepubertal children have often been told not to let anyone look at their genitals, so the exception
of a medical examination needs to be explained by their parent before and during the consultation.
Adolescents are often highly embarrassed by any sort of genital examination, so preserving
their modesty and asking any male relatives to leave the room is recommended.
The most often recommended position is prone knee to chest and visualisation of the vesti-
bule can be facilitated by gently pulling the labia majora out and up.
The supine position with the legs in a frog-leg position is recommended to visualise the
vaginal opening, however if the child is not relaxed, contraction of the perineal and gluteal
muscles may render the examination difficult.
In some cases, if it is necessary to examine the upper vagina in a child (i.e. if a tumour or
foreign body is suspected) an examination under general anesthesia is recommended.1, 16, 17
176
9.4
Testing for infection
If a streptococcal or H. influenzae infection is suspected, a saline moistened swab can be
used to take a sample from the introitus. Although attempting to take a vaginal swab is
unpleasant for a small child, in these cases it is necessary.
If there is a thick green discharge and/or sexual abuse is suspected, a vaginal sample is nec-
essary to rule out Neisseria gonorrhoeae, Chlamydia trachomatis and Trichomonas vaginalis.
The sample may be taken using a thin catheter and a syringe rather than inserting a swab. A
urine polymerase chain reaction test can be considered rather than a vaginal sample.
9.5
Vaginal discharge
The presence of discharge in the absence of symptoms is not usually a cause of concern and
should not prompt further investigation.
Although candidiasis and bacterial vaginosis (BV) are the most common causes of acute
vaginal discharge or pruritus in an adult, these are rare causes of vulvovaginal symptoms in
children. In fact, in some cases, what may be perceived as discharge may have other causes
(i.e. an ectopic ureter or a lymphatic malformation).11-15
TABLE 9.1 Etiologies of vaginal discharge in children
Physiologic
Foreign Body
Infection: group A Streptococcus, H. influenza
Lymphatic malformation
Ectopic ureter
Fistula
Bacterial vaginosis (very rare)
Lichen planus (very rare)
Vaginal discharge due to infection
Bacterial vaginitis
Introduction
Bacterial vaginitis (not to be confused with BV) is an uncommon cause of vulvovaginal
symptoms in the prepubertal child. However, the thin, fragile, un-estrogenized vagina of
prepubertal girls, the proximity of the vagina to the perianal skin colonized with enteric
organisms, and hygiene habits of small children provide a fertile environment for bacterial
infection of the vagina compared to that of well estrogenized adolescent and adult women.
177
Prevalence
The prevalence of prepubertal bacterial vaginitis is unknown in part because many microor-
ganisms colonizing the vagina of young girls are only occasionally pathogens, with asymp-
tomatic girls occasionally exhibiting positive cultures for these bacteria. These issues pre-
vent clear diagnoses, and therefore the few available data on prevalence are of poor quality.
Etiology and pathophysiology
The most common causes of symptomatic vaginitis in childhood are S. pyogenes, group B
Streptococcus (S. agalactiae), S. aureus, H. influenzae, E. coli, Enterococcus faecalis, Klebsiella
pneumoniae, and Shigella spp..
Risk factors
Risk factors for bacterial vaginitis are believed to include poor perineal hygiene, wiping back
to front, the presence of foreign bodies, and sexual abuse.
Signs and symptoms
The most frequent symptoms of
bacterial vaginitis include vulvar
itching, pain, odor, and vulvar
dysuria. Redness of the introitus
and, often, the modified mu-
cous membranes of the vulva,
as well as a yellow or even green
vaginal discharge, are the most
common signs. (Figure 9.2)
Sometimes the discharge may
also be described as yellow/
brown staining of the under-
wear. Vaginal bleeding occurs
in a minority of children.
Diagnosis
There is considerable overlap between colonizers and pathogens, so a diagnosis of bacterial
vaginitis by culture alone is not recommended. Cultures should be performed in the setting
of clinical inflammation and symptoms, while keeping in mind that the estrogen deficient
introitus is often normally strikingly erythematous.
Culture results showing a pure growth of a potential pathogen rather than mixed flora are
more likely to represent true infection, and a positive response to treatment confirms the diag-
nosis. Generally, most bacterial vaginitis are associated with respiratory pathogens.10, 18 S. py-
ogenes is by far the most common cause of childhood bacterial vaginitis, and often accom-
panies or follows streptococcal pharyngitis. H. influenza previously was another relevant cause,
Figure 9.2 Vulvar and perianal redness in a young girl with bacterial
vaginitis (S. pyogenes)
178
but the increasing number of children vaccinated against this microorganism has likely resulted
in a decreased prevalence of H. influenza bacterial vaginitis. K. pneumoniae and S. aureus, can
also cause bacterial vaginitis.10, 18, 19 Enteric organisms, especially E. coli and Enterococcus spp., are
found more often in cultures from children with vaginitis symptoms compared to controls, but
their role can be difficult to establish.19, 20 Likewise, Shigella spp. and Yersinia spp. are rarely found.
Rarely, testing yields N. gonorrhoeae as a cause for purulent vaginitis. These children are
likely to have been sexually abused; any concern for sexual abuse should prompt molecu-
lar studies for gonorrhea, chlamydia and trichomonas as causes for the vaginitis; certainly,
there should be screening for other sexually transmitted diseases and appropriate referrals
for these girls. However, half or more girls with symptoms and signs of vaginitis show no
recognized pathogens on culture and are labeled “nonspecific vulvovaginitis”.21, 22 Alterna-
tive diagnoses should be considered in these children. For example, dermatoses such as LS,
eczema, and irritant contact dermatitis can produce itching, pain, and superficial exudate
that can mimic vaginitis. Urinary tract infections can also be considered.
Treatment
The management of symptomatic bacterial vaginitis consists of oral antibiotics chosen
based on the culture results, as well as counseling regarding local care. The use of a mild
emollient such as petroleum jelly often provides some comfort, especially if used liberally
as a protective barrier before urination. Wiping back to front, particularly in children with
enteric organisms on culture, irritants on delicate skin (including medicated creams), and
excessive cleaning that leads to inflamed skin may play a role in the development of bacte-
rial vaginitis in some children, so these are not recommended.
Bacterial vaginitis can be recurrent. In addition to ensuring that local care has been addressed
and followed, these children should be evaluated for complicating factors, such as vaginal foreign
bodies or accompanying skin disease such as LS, irritant contact dermatitis, or eczema that in-
creases the risk of secondary infection. When enteric organisms are found recurrently and per-
ineal hygiene has been addressed, the rare event of an enteric fistula should be considered.
Pinworms/threadworms
Introduction
Although often recognized as a common cause of perianal symptoms in young children,
pinworms can also migrate from the intestines into the vagina to lay eggs. The infection
occurs when eggs are ingested via contaminated fingers. Scratching the area transfers eggs
to under the fingernails. Then, ingested eggs perpetuate the cycle.23
Etiology and pathophysiology
Pinworms/threadworms (Enterobius vermicularis) are common intestinal parasites.
179
Prevalence
Pinworms are common, particularly in overcrowded living conditions with poor personal
hygiene.24
Risk Factors
Risk factors include crowded conditions, poor hygiene, and warm, tropical climates.
Signs and symptoms
These worms resemble short strings of white thread and cause inflammation of the fragile
poorly estrogenized vaginal mucosa of young girls. Usually, children exhibit itching, irrita-
tion and, often redness and dermatitis of the perianal skin as well as abdominal pain at times
and sleep disturbance. When vaginitis occurs, there is redness of the introitus and a purulent
vaginal discharge.23
Diagnosis
The diagnosis is made by direct visualization of the worms at the anal verge, usually at night,
or by identification of worms microscopically; tape is pressed against the anal verge when the
child first wakes in the morning, and the tape is then affixed to a glass slide for microscopy.25
Treatment
Treatment consists of mebendazole, pyrantel pamoate, or albendazole. Any of these drugs
is given in one dose, which is then repeated two weeks later.26 It is, however, important
to know that these drugs do not inactivate the parasite’s eggs. A low potency topical cor-
ticosteroid ointment (i.e. desonide 0.05% or hydrocortisone 2.5%) applied to the areas of
inflammation can hasten resolution of symptoms of itching and pain. All members of the
household should be treated, and careful patient education regarding means of transmis-
sion and handwashing is important.25, 27
Candidiasis
Topical antifungal therapy is a common empirical management strategy for the treatment
of any vulvovaginal discharge, itching or irritation. However, once a child is out of diapers,
any relief is almost always from the emollient properties of the vehicle, rather than from the
antifungal activity of the medication.19, 28
Yeasts are a rare cause of vaginitis in prepubertal children, despite positive culture of C. al-
bicans from genital samples (not necessarily vaginal) reported in up to 5% of asymptomatic
children.10 The unestrogenized vagina is a hostile environment for the growth of yeasts, and
unless a child is immunosuppressed, obese, diapered, diabetic, and rarely following anti-
biotics, the likelihood of candidiasis is exceptional. Not infrequently, in older, immediately
prepubertal girls, florid vulvovaginal candidiasis may present prior to onset of menses due
to the physiological increase of estrogen. Menses usually follow in weeks or months.
When candidiasis is suspected, its presence should be confirmed by culture, microscopy, or
molecular tests, and underlying predisposing factors sought.
180
In the rare case of vaginal candidiasis in a child, oral fluconazole is indicated, with topical
nystatin ointment as a less irritating medication for vulvar involvement.
Bacterial vaginosis
BV is generally a disease of post pubertal women. Despite the absence of prevalence data in
children, BV is believed to be rare.
Recent data reports vaginal/introital cultures of prepubertal girls that yield bacteria asso-
ciated with this condition. Almost 14% of asymptomatic girls were found to have vaginal
Gardnerella spp., one of the bacteria involved in the development of BV.10 Although con-
troversial, some limited data have shown these organisms to be more common in sexually
abused girls. In case of a confirmed diagnosis of BV in a child, a history of sexual abuse
should be investigated.10 However the presence of Gardnerella spp. alone is not synonymous
with BV. As with BV, lack of lactobacilli and elevated pH are normal in children due to lack of
estrogen and are not useful for the diagnosis in this population.
If treatment is needed, oral metronidazole or clindamycin are recommended, since intravag-
inal therapy is inappropriate.
Other infections producing vaginitis
There are several systemic infections that sometimes produce a purulent vaginitis. These
include varicella, which is extremely rare in countries that vaccinate against it. Vaginal ero-
sions from short-lived mucosal vesicles produce inflammation, which predispose to S. pyo-
genes infection, another common cause of bacterial vaginitis in children.29 Therefore, vagini-
TABLE 9.2 Infectious causes of vaginal discharge and management
Diagnosis Management
S. pyogenes, S. agalactiae
Penicillin 250 mg 2-3 times a day
Amoxicillin 50 mg/Kg daily (maximum 1 g)
Cephalexin 20 mg/Kg 2 times a day (maximum 500 mg per dose)
Clindamycin 7 mg/Kg/dose, 3 times a day (maximum dose 300 mg)
Duration of treatment for vaginitis has not been studied in children
S. aureus, H. influenzae, E.coli
and all other causes of bacterial
vaginitis
By sensitivities on culture
Bacterial vaginosis
Clindamycin 5-7 mg/Kg twice daily for 7 days (maximum daily dose 300 mg)
Metronidazole 15-25 mg/Kg/day in three divided doses, maximum 2 g for
7 days
Pinworms
Mebendazole 100 mg once, repeated in 3 weeks for children over 2 years
Pyrantel pamoate1 mg/Kg in one dose, maximum 1 g; repeat dose in 2 weeks
Albendazole 400 mg as a single dose; repeat dose in 2 weeks
Candidiasis
Confirm with microscopy, culture or molecular tests
Fluconazole 12 mg/Kg in single dose, may repeat in 3 days
Nystatin ointment 3 times a day for vulvar involvement
Systemic infections producing
vaginitis Identification and management of systemic infection
181
tis associated with varicella should undergo bacterial culture and antibiotic treatment when
bacterial pathogens are identified. Measles, upper respiratory infections, and gastrointesti-
nal infections are sometimes associated with vaginitis.
Vaginal foreign bodies
Introduction
A classic cause of persistent discharge in girls and prepubertal vaginal bleeding is a foreign
body in the vagina. Bleeding in such a situation might be accompanied by pelvic pain and a
foul-smelling discharge.30
Etiology and pathophysiology
Although the range of objects that find their way into the vagina of a prepubertal child
can be remarkable, toilet paper is the most common finding. Toys, household items such as
safety pins and pen caps occur. The most damaging foreign bodies are batteries, which can
produce ulcers, scarring, become embedded into the vaginal walls, and cause fistulae.31, 32
Prevalence
The frequency of vaginal foreign bodies is unknown.
Signs and Symptoms
Vaginal bleeding is a more common presenting sign than a vaginal discharge, and in a re-
cent study, foreign bodies were by far the most common cause of vaginal bleeding in 158
prepubertal girls.33 In addition, rectal bleeding with a vaginal foreign body sometimes oc-
curs.34 When vaginal foreign bodies present as a purulent discharge, this is often mistaken
for a primary bacterial vaginitis. However, despite initial improvement following antibiotic
treatment, the discharge recurs after therapy. Other signs of an inflammatory vaginitis may
be present, including erythema of the introitus as a result of either secondary infection, or
simply an irritant contact dermatitis from the purulent vaginal secretions. The unidentified
foreign body may lead to urinary tract infection or dermatosis and, in serious cases, to per-
foration into the peritoneal cavity or fistula formation.30
Diagnosis
The diagnosis is made by identification of the foreign body. The vagina can sometimes be
visualized without instrumentation in a knee-chest position, and the foreign body seen. Oc-
casionally, especially when the foreign body is toilet paper, it is both identified and treated
by rinsing the vagina with saline.
Ultrasound can be used to identify a foreign body in some instances, although items such as
toilet paper may not be evident.35, 36 Metal and some other dense objects may be seen on a
plain radiograph. The vagina can also be examined, usually under general anesthesia, using
a hysteroscope inserted into the vagina.
Rarely, foreign bodies present for prolonged periods have been associated with pelvic ab-
182
scess, vesicovaginal fistulae, rectovaginal fistulae, and vaginal stenosis. The symptoms in
these cases can include dysuria, urinary incontinence, and pelvic/abdominal pain.37
Treatment
The treatment consists of removing the foreign body. Occasionally, the vagina can be rinsed
with a soft urinary catheter, and the foreign body washed out with the fluid. This is especially
likely when the foreign body is toilet paper. When the retained object is a battery, emergent
removal is indicated.
Otherwise, the foreign body is removed under general anesthesia/conscious sedation.38
9.6
Dermatoses and dermatitis which involve the vagina
Fixed drug eruption, toxic epidermal necrolysis, and erythema multiforme are severe cuta-
neous drug reactions (SCAR) which involve skin and mucosa. These conditions may include
a severe erosive vaginitis which can be followed by vaginal synechiae.
Lichen planus is very rare in children but can occasionally cause vaginitis in prepubertal
children. Involvement of the vulva is usually also present.
In practice, vaginitis with discharge seldom occurs in isolation as the associated discharge
usually causes skin irritation and inflammation of the vulva.
Chemical irritants such as bubble bath, soaps and chlorine in swimming pools can cause
inflammation of the vestibule which can result in weeping and simulate a discharge.4, 39
Recommendations
Recommendation Quality of
evidence
Strength of
recommendation
Erythema of the vestibule is common in children and, in the absence of
symptoms, it is of no significance and does not require investigation or
treatment.
5 D
The presence of an inoffensive, thin milky or greenish discharge,
not associated with symptoms or clinically obvious inflammation is
common and harmless in prepubertal children and does not require
investigation or treatment.
5 D
If it is necessary to examine the upper vagina in a child, an examina-
tion under general anesthesia is recommended. 5 D
If a streptococcal or Haemophilus influenzae infection is suspected, a
saline moistened swab can be used to take a sample from the introitus. 5 D
If there is a thick green discharge and/or sexual abuse is suspected, a
vaginal swab is necessary to rule out sexually transmitted infections.
If a vaginal sample is needed, it may be taken using a thin catheter and
a syringe rather than inserting a swab. 5 D
183
A diagnosis of bacterial vaginitis by culture alone is not recommended. 5 D
The diagnosis of a sexually transmitted infection must prompt ade-
quate referral due to suspected sexual abuse. 3b B
The management of symptomatic bacterial vaginitis consists of oral an-
tibiotics chosen on the basis of the culture results, as well as counseling
regarding local care.
4 C
The treatment of pinworms consists of all members of the household
with mebendazole, pyrantel pamoate, or albendazole (one dose, which
is then repeated two weeks later).
3a B
In the rare case of vaginal candidiasis in a child, oral fluconazole is indicated. 5 D
In a case of a confirmed diagnosis of bacterial vaginosis in a child, a
history of sexual abuse should be investigated. 5 D
References
1. McCann, J.; Wells, R.; Simon, M.; Voris, J., Genital findings in prepubertal girls selected for nonabuse: a descriptive
study. Pediatrics 1990, 86, (3), 428-39.
2. Abdelrahman, H. M.; Feloney, M. P., Imperforate Hymen. In StatPearls, StatPearls Publishing Copyright © 2022, Stat-
Pearls Publishing LLC.: Treasure Island (FL), 2022.
3. Wróblewska-Seniuk, K.; Jarząbek-Bielecka, G.; Kędzia, W., Gynecological Problems in Newborns and Infants. J Clin
Med 2021, 10, (5).
4. Berenson, A. B., The prepubertal genital exam: what is normal and abnormal. Curr Opin Obstet Gynecol 1994, 6, (6), 526-30.
5. Pillai, M., Genital findings in prepubertal girls: what can be concluded from an examination? J Pediatr Adolesc Gy-
necol 2008, 21, (4), 177-85.
6. Ayson, N.; Starling, S., Normal Examination Findings and Variants. In Handbook of Interpersonal Violence and Abuse
Across the Lifespan: A project of the National Partnership to End Interpersonal Violence Across the Lifespan (NPEIV),
Geffner, R.; White, J. W.; Hamberger, L. K.; Rosenbaum, A.; Vaughan-Eden, V.; Vieth, V. I., Eds. Springer International
Publishing: Cham, 2020; pp 1-14.
7. Goff, C. W.; Burke, K. R.; Rickenback, C.; Buebendorf, D. P., Vaginal opening measurement in prepubertal girls. Am J
Dis Child 1989, 143, (11), 1366-8.
8. Elstein, M., Vaginal cytology of the newborn. J Obstet Gynaecol Br Commonw 1963, 70, 1050-5.
9. Chen, X.; Lu, Y.; Chen, T.; Li, R., The Female Vaginal Microbiome in Health and Bacterial Vaginosis. Front Cell Infect
Microbiol 2021, 11, 631972.
10. Neyazi, S., Prepubertal vulvovaginitis. Journal of Nature and Science of Medicine 2019, 2, (1), 14-22.
11. Xiaoming, W.; Jing, L.; Yuchen, P.; Huili, L.; Miao, Z.; Jing, S., Characteristics of the vaginal microbiomes in prepubertal
girls with and without vulvovaginitis. Eur J Clin Microbiol Infect Dis 2021, 40, (6), 1253-1261.
12. Sugar, N. F.; Graham, E. A., Common gynecologic problems in prepubertal girls. Pediatr Rev 2006, 27, (6), 213-23.
13. Hayes, L.; Creighton, S. M., Prepubertal vaginal discharge. The Obstetrician & Gynaecologist 2007, 9, (3), 159-163.
14. Hickey, R. J.; Zhou, X.; Settles, M. L.; Erb, J.; Malone, K.; Hansmann, M. A.; Shew, M. L.; Van Der Pol, B.; Fortenberry, J.
D.; Forney, L. J., Vaginal microbiota of adolescent girls prior to the onset of menarche resemble those of reproduc-
tive-age women. mBio 2015, 6, (2).
15. Matytsina, L. A.; Greydanus, D. E.; Gurkin, Y. A., Vaginal microbiocoenosis and cytology of prepubertal and adolescent
girls: their role in health and disease. World J Pediatr 2010, 6, (1), 32-7.
16. Jacobs, A. M.; Alderman, E. M., Gynecologic examination of the prepubertal girl. Pediatr Rev 2014, 35, (3), 97-104.
17. Physicians, T. R. A. C. o. Genital examinations in girls and young women: a clinical practice guideline. https://ranzcog.edu.
au/wp-content/uploads/2022/05/Genital-Examinations-in-Girls-and-Young-Women-A-Clinical-Practice-Guideline.pdf
18. Kim, H. C.; Lee, M. H.; Hong, S. G., Pediatric Vulvovaginitis: A Study of Clinical and Microbiologic features and the
Efficacy of Perineal Hygienic Care. Korean J Obstet Gynecol 1999, 42, (12), 2821-2828.
19. Loveless, M.; Myint, O., Vulvovaginitis- presentation of more common problems in pediatric and adolescent gyne-
cology. Best Pract Res Clin Obstet Gynaecol 2018, 48, 14-27.
20. Gorbachinsky, I.; Sherertz, R.; Russell, G.; Krane, L. S.; Hodges, S. J., Altered perineal microbiome is associated with
vulvovaginitis and urinary tract infection in preadolescent girls. Ther Adv Urol 2014, 6, (6), 224-9.
184
21. Jarienė, K.; Drejerienė, E.; Jaras, A.; Kabašinskienė, A.; Čelkienė, I.; Urbonavičienė, N., Clinical and Microbiological
Findings of Vulvovaginitis in Prepubertal Girls. J Pediatr Adolesc Gynecol 2019, 32, (6), 574-578.
22. Baka, S.; Demeridou, S.; Kaparos, G.; Tsoutsouras, K.; Touloumakos, S.; Dagre, M.; Meretaki, S.; Chasiakou, A.; Koumaki,
V.; Tsakris, A., Microbiological findings in prepubertal and pubertal girls with vulvovaginitis. Eur J Pediatr 2022, 181,
(12), 4149-4155.
23. Serban, E. D., Perianal infectious dermatitis: An underdiagnosed, unremitting and stubborn condition. World J Clin
Pediatr 2018, 7, (4), 89-104.
24. Rivero, M. R.; De Angelo, C.; Feliziani, C.; Liang, S.; Tiranti, K.; Salas, M. M.; Salomon, O. D., Enterobiasis and its risk
factors in urban, rural and indigenous children of subtropical Argentina. Parasitology 2022, 149, (3), 396-406.
25. Bharti, B.; Bharti, S.; Khurana, S., Worm Infestation: Diagnosis, Treatment and Prevention. Indian J Pediatr 2018, 85,
(11), 1017-1024.
26. Centers for Disease Control and Prevention, Parasites - enterobiasis, https://www.cdc.gov/parasites/pinworm/treat-
ment.html
27. Weatherhead, J. E.; Hotez, P. J., Worm Infections in Children. Pediatr Rev 2015, 36, (8), 341-52; quiz 353-4.
28. Fischer, G. O., Vulval disease in pre-pubertal girls. Australas J Dermatol 2001, 42, (4), 225-34; quiz, 235-6.
29. Hasin, O.; Hazan, G.; Rokney, A.; Dayan, R.; Sagi, O.; Ben-Shimol, S.; Greenberg, D.; Danino, D., Invasive Group A Strep-
tococcus Infection in Children in Southern Israel Before and After the Introduction of Varicella Vaccine. J Pediatric
Infect Dis Soc 2020, 9, (2), 236-239.
30. Dwiggins, M.; Gomez-Lobo, V., Current review of prepubertal vaginal bleeding. Curr Opin Obstet Gynecol 2017, 29,
(5), 322-327.
31. Nakib, G.; Calcaterra, V.; Pelizzo, G., Longstanding Presence of a Vaginal Foreign Body (Battery): Severe Stenosis in a
13-Year-Old Girl. J Pediatr Adolesc Gynecol 2017, 30, (1), e15-e18.
32. Yanoh, K.; Yonemura, Y., Severe vaginal ulcerations secondary to insertion of an alkaline battery. J Trauma 2005, 58,
(2), 410-2.
33. Zhang, J.; Zhang, B.; Su, Y.; Guo, S.; Liu, C.; Bai, J.; Xie, X., Prepubertal Vaginal Bleeding: An Inpatient Series from a
Single Center in Fujian China. J Pediatr Adolesc Gynecol 2020, 33, (2), 120-124.
34. Shiryazdi, S. M.; Heiranizadeh, N.; Soltani, H. R., Rectorrhagia and vaginal discharge caused by a vaginal foreign
body--a case report and review of literature. J Pediatr Adolesc Gynecol 2013, 26, (3), e73-5.
35. Gross, I. T.; Riera, A., Vaginal Foreign Bodies: The Potential Role of Point-of-Care-Ultrasound in the Pediatric Emergen-
cy Department. Pediatr Emerg Care 2017, 33, (11), 756-759.
36. Yang, X.; Sun, L.; Ye, J.; Li, X.; Tao, R., Ultrasonography in Detection of Vaginal Foreign Bodies in Girls: ARetrospective
Study. J Pediatr Adolesc Gynecol 2017, 30, (6), 620-625.
37. Ekinci, S.; Karnak, İ.; Tanyel, F. C.; Çiftçi, A., Prepubertal vaginal discharge: Vaginoscopy to rule out foreign body. Turk
J Pediatr 2016, 58, (2), 168-171.
38. Ma, W.; Sun, Y. F.; Liu, J. H.; He, D. W.; Lin, T.; Wei, G. H., Vaginal foreign bodies in children: a single-center retrospective
10-year analysis. Pediatr Surg Int 2022, 38, (4), 637-641.
39. Berenson, A. B.; Heger, A. H.; Hayes, J. M.; Bailey, R. K.; Emans, S. J., Appearance of the hymen in prepubertal girls.
Pediatrics 1992, 89, (3), 387-94.
185
10.1
Introduction
According to consensus guidelines from the International Scientific Association for Probi-
otics and Prebiotics (ISAPP), a probiotic is a “live microorganisms that, when administered
in adequate amounts, confer a health benefit on the host”.1 This definition excludes micro-
bial transplants and live cultures in food. The same organization has provided a definition
for prebiotic, which states: “a substrate that is selectively utilized by host microorganisms
conferring a health benefit”.2 Finally, a synbiotic is defined as “a mixture comprising live
microorganisms and substrate(s) selectively utilized by host microorganisms that confers
a health benefit on the host”. The ISAPP further divides synbiotics into synergistic and com-
plementary. Per the publication: “A ‘synergistic synbiotic’ is a synbiotic in which the substrate
is designed to be selectively utilized by the co-administered microorganism(s). A ‘comple-
mentary synbiotic’ is a synbiotic composed of a probiotic combined with a prebiotic, which
is designed to target autochthonous microorganisms”.3
Before discussing data for the use of the products in specific conditions, we must first ad-
dress the available data (or lack thereof) on whether there is a difference in oral vs. vaginal
administration. We presume that for a product to be effective for vaginal health it must ac-
tually reach the vagina. Vaginal administration directly applies the product to the target
site. In the few studies that have looked for probiotic strains in the vagina and gut after oral
administration, the probiotic strains were able to be cultivated from the vagina in 8-75% of
women during administration, but colonization decreased after cessation of use.4-8 Howev-
er, only in one of the two studies which included a placebo arm, the probiotic was found in
45% of participants.8 In the only study to use advanced, molecular detection methods, the
probiotic strains were rarely detected in either vaginal or fecal samples.9
We should also acknowledge that probiotics, prebiotics and synbiotics are broad terms, en-
compassing a wide variety of products, and that there is significant heterogeneity in the
literature with regard to dose, dosing frequency, duration and specific microbial strains,
making a comprehensive and consistent summary of the data challenging.
10
PROBIOTICS, PREBIOTICS
AND SYNBIOTICS
FOR VAGINITIS
(alphabetical order)
Colin MacNeill
Caroline Mitchell
Francesco de Seta
186
10.2
Bacterial vaginosis
Probiotics
The desire to use probiotics in the management of bacterial vaginosis (BV) is fueled by the
high rate of BV recurrence. Both providers and patients are looking desperately for some-
thing better. Some studies look at probiotics as an alternative to antibiotic therapy, while
most studies evaluate the impact of post-antibiotic use of probiotics to prevent BV recur-
rence. As noted above, there is significant heterogeneity in the studies’ methodologies.
When evaluating the available data, it is important to consider the dimensions a study would
need to have to confidently prove that a therapy is either not effective, or is non-inferior to
a standard regimen. When considering recurrent BV at one or six months, and two different
magnitudes of reduction in that rate, the smallest randomized trial that would have enrolled
a sufficient number of participants to detect a 50% reduction in BV recurrence at six months
would be of at least 84 women. (Table 10.1) A study can still find a significant difference be-
tween arms when underpowered, but if no significant differences are seen between arms in a
smaller trial the conclusion should not be that the arms are statistically equivalent – but that the
study is underpowered and the answer remains unknown. Additionally, follow-up duration var-
ies between studies. For clinicians and patients, long-term absence of recurrence is the primary
clinical goal, thus we will focus on studies which had at least one month or more of follow up.
TABLE 10.1 Estimates of necessary sample size for a trial to be able to detect a difference
between the referent and desired bacterial vaginosis incidence.
Condition Referent Desired Sample size
Recurrent BV 30% at 1 month 20% at 1 month 293 per arm
15% at 1 month 120 per arm
Recurrent BV 60% at 6 months 40% at 6 months 97 per arm
30% at 6 months 42 per arm
Considering only randomized studies over 85 participants, with more than one month of
follow up, we are left with 10 studies: four evaluating an oral probiotic10-13 and six that used
a vaginal formulation.14-19 Three of the four studies on oral probiotics first treated partici-
pants with metronidazole, and all treated women with a probiotic for 30-120 days. Of the
four studies, one did not have clear specifications for the primary analysis, which suggests
significant potential for bias.12 Of the remaining three, two showed a statistically significant
reduction in BV recurrence in the probiotic arm: one using L. rhamnosus GR-1 and L. reuteri
RC-1410, and one that used a combination of L. crispatus LMG S-29995, L. brevis, and L. aci-
dophilus in proportion of 60%, 20%, and 20%, respectively.11 The study which did not show
benefit also used L. rhamnosus GR-1 and L. reuteri RC-14, but in a liquid drink form.13
Of the six studies meeting our selection criteria evaluating vaginal probiotics, all but one first
187
treated participants with antibiotics and then delivered the probiotic between seven days to 11
weeks, though some included intermittent repeat dosing. Three studies showed a significant
reduction in BV recurrence (one using L. rhamnosus, L. acidophilus, S. thermophilus19; one using
L. gasseri and L. rhamnosus17; and one using L. crispatus15), and three did not (products included
an L. casei and L. fermentum impregnated tampon16; L. acidophilus alone14; or L. casei alone18).
In 2022 Liu et al. included eighteen studies in a review – however, two of the included stud-
ies reported the use of fluconazole, suggesting not all studies were for BV treatment.20 In
comparison with isolated antibiotics, antibiotics plus probiotics significantly decreased the
recurrence rate of BV and increased the cure/remission rate of BV at 1-3 months and overall
analysis. Compared with placebo, probiotics decreased the recurrence rate of BV (at 1-3
months and overall analysis) and increased the cure/remission rate of BV (at 1-3 months). In
comparison with short-term probiotics treatment (<1 month), long-term probiotics treat-
ment (1-3 months) yielded superior beneficial outcomes and efficacy in the treatment of BV.
Besides, probiotics were indeed evidently more effective than placebo, and antibiotic plus
probiotics produced better results than isolated antibiotics.
In 2021 Tidbury et al., including 33 studies for a systematic review, focused the metanalysis
on two major categories: treatment and prevention of BV.21 The authors considered as main
outcomes efficacy of treatment, cure of BV, BV recurrence rate, improvement of vaginal mi-
crobiota and/ or clinical signs and symptoms. The treatment group was categorized based
on the type of intervention (oral lactobacilli, vaginal lactobacilli, lactobacilli and estriol, lac-
tobacilli supplementary to antibiotics, lactobacilli and estriol supplementary to antibiotics,
lactic acid, and sucrose). The prevention group was based on whether the intervention was
administered directly after standard antibiotic treatment (prevention of persistence) or on
currently healthy women with a history of recurring BV (prevention of recurrence).
In the same year Munoz-Barreno et al. reported a total of 57 randomized controlled trials (RCTs),
comparing the effectiveness of BV treatments with different doses of antibiotics and/or pro-
biotics through oral and local administration.22 The highest P-scores (a scores that estimates
the effect sizes of pairwise treatment comparisons) in clinical cure rate were obtained by: (1)
a combined therapy of local probiotic treatment, vaginal and oral antibiotic (5-nitroimidazole
and clindamycin, respectively) (P-score = 0.92); (2) a combined therapy of oral administration
of 5-nitroimidazole and probiotic treatment (P-score = 0.82); and (3) a combined therapy of
local administration of 5-nitroimidazole and oral probiotic treatment (P-score = 0.68). Finally,
combined therapies suggested a reduction of the optimal concentration of antibiotics, and
double phase treatments of antibiotics indicated an increment of clinical cure rates in BV.22
In 2020 Jeng et al. tried to clarify the efficacy of probiotics in the treatment of common vaginal
infections in non-pregnant females including vulvovaginal candidiasis (VVC), BV and mixed
infection (BV plus VVC). In conclusion, the authors stressed the concept that probiotics as
a supplement to conventional pharmacological treatments are effective in the short term
for the treatment of common vaginal infections in non-pregnant adult females.23 However,
high-quality evidence for the effectiveness of probiotics alone in recurrent or curative vagi-
nal infections is limited.
188
In 2019, two studies reported that probiotic regimes are safe and may exhibit a short and
long-term beneficial effect on BV treatment but there is currently no strong evidence that
probiotic monotherapy is more effective than traditional antibiotics.24, 25
Probiotics for bacterial vaginosis in pregnancy
The failure of antibiotic treatment in pregnant women with BV to reduce preterm birth risk
has led investigators to postulate that the underlying causative biologic disorder may be the
absence of Lactobacillus spp.. Early trials randomizing patients to oral probiotic or placebo
were faulted for treating with an inappropriate strain or number of species of lactobacilli, an
inadequate probiotic dose, duration or delivery route, or whether the administered probiot-
ic was identified in the vaginal microbiota.
Husain et al.26 sought to determine whether a daily oral probiotic containing L. rhamnosus
and L. reuteri (each at 2.5 x 109 colony forming units [CFU] per dose) would colonize the vagi-
na and reduce the incidence of BV. They randomized 304 women from East London between
9-14 weeks gestational age to either probiotic or placebo from entry until delivery. The pri-
mary outcome was the rate of BV at 18-20 weeks by Nugent score. At 18-20 weeks, BV was
present in 15% of the probiotic group and in 9% of the placebo group, which was not statistically
significant. They concluded that the oral probiotic used in the study did not reduce the incidence
of BV in pregnant women. The trial was underpowered to detect a change in the preterm birth risk.
Yang et al.27 also investigated whether abnormal Nugent scores in pregnancy could be nor-
malized by a probiotic approach. They randomized 86 asymptomatic subjects under 17
weeks of gestation to twice daily oral L. rhamnosus GR-1 and L. reuteri RC-14 (each at 2.5 x
109 CFU per dose) or placebo for 12 weeks and assessed the vaginal microbiota, cytokines
and chemokines at 26 and 35 weeks. There was no significant reduction in Nugent score,
Shannon diversity index or in cytokines at 28 or 35 weeks in either arm.
Prebiotics and synbiotics
In one randomized trial of 100 women with BV, Hakimi et al. reported that concurrent use
of a daily prebiotic vaginal gel containing 2% red clover extract, 10% inulin and 10% fruc-
to-oligosaccharides improved the efficacy of oral metronidazole for BV treatment compared
to a placebo gel (76 vs. 30%, p=0.012 cure by Amsel criteria and Nugent score at 10 days).28
Randomization to an oral synbiotic formulation containing L. acidophilus, L. rhamnosus and
lactoferrin (a glycoprotein found in cervical mucus), was associated with a lower rate of BV
recurrence at six months in people with recurrent BV compared to placebo (29 vs. 58%,
p<0.05).29, 30 Several studies have been conducted with a vaginal synbiotic vaginal formula-
tion of L. rhamnosus and lactose, though many of these studies were unblinded.31
Nasioudis et al. proposed that restoration of Lactobacillus spp. dominance requires the res-
toration of innate immune factors such as lactoferrin that target anaerobic bacteria and the
availability of nutrients favoring proliferation of lactobacilli.32 Based on the ability of lactofer-
rin to sequester iron required by anaerobic bacteria, Miranda et al. reviewed data prospec-
tively collected from all consecutive patients with a history of preterm birth who screened
189
positive for BV before 13 weeks and administered vaginal lactoferrin 300 mg daily for 21
days.33 The primary outcome was preterm birth (<37 weeks) in those administered lacto-
ferrin compared to similar patients who did not receive lactoferrin. They found that those
who were administered lactoferrin had a significantly lower preterm birth rate (25 vs. 44.6%,
p=0.02). No adverse events were reported. Because the study was not randomized and lacks
microbiologic outcomes, we do not at this time recommend vaginal lactoferrin treatment.34
Clinical recommendations
In general, results are often not comparable between studies due to differences in species,
strains, dose, and route of administration. Additionally, upon critical review there is a high
risk of bias in many papers. Although there is not enough evidence yet for these alternatives
to be a part of formal treatment recommendations, there may be some significant benefit
for some people, without reported significant adverse effects. This makes them an attrac-
tive, despite unproven option for patients with refractory, recurrent BV and may be consid-
ered in a clinical setting. Several variables (such as Lactobacillus species and concentration,
formulation, administration route, time and phases of treatment) should be considered.
Unresolved controversies include whether probiotics more successfully decrease the recur-
rence rate of BV in short (one month) vs. long-term period (three months) and if they have to be
used after standard antibiotic treatment (prevention of persistence) or can be taken by current-
ly healthy women with a history of recurring BV (prevention of recurrence), and whether repeat
maintenance dosing is necessary. Also, it must be taken into account that probiotic formula-
tions are usually expensive and, if effective, likely to have to be used for long periods of time.
Based on Husain26, Yang27, and Mirandas33 studies we recommend against probiotic and
synbiotic treatment of BV in pregnancy until subsequent studies are conducted that are
powered to detect a reduction of the end points of preterm birth and chorioamnionitis.
10.3
Vulvovaginal candidiasis
Probiotics
The idea that promoting vaginal lactobacilli colonization for prevention or treatment of
yeast is supported by in vitro laboratory data showing that many Lactobacillus species inhib-
it the growth of Candida spp., alter the expression of Candida spp. virulence factors or inhibit
the hyphal transformation which is thought to increase the likelihood of symptoms.35, 36 In
mouse models, vaginal application of lactobacilli decreases fungal burden.37, 38 However, in
humans, several large epidemiologic studies have shown no association between a Lacto-
bacillus-dominant microbiota and a lower risk for VVC. In fact, more often there is a higher
prevalence of VVC in women with high proportions of vaginal lactobacilli.39-43
There are few well executed randomized clinical trials of sufficient size on which to base rec-
ommendations for use of probiotics for prevention of VVC. There are only two randomized
trials with a sample size over 100. One is a randomized, open label trial of a vaginal product
190
including L. acidophilus, L. rhamnosus, Steptococcus thermophilus, and L. delbrueckii subsp.
bulgaricus used daily for 10 days after an antifungal. Of the 416 premenopausal women
included, 5% of the intervention group (antifungal followed by probiotic) and 37% of the
antifungal only arm had a positive culture for Candida spp. 30-45 days after antifungal treat-
ment.44 The second trial randomized 278 women about to receive antibiotics for a non-gy-
necologic infection in a multifactorial design with four arms comparing oral and vaginal
probiotics (oral: L. rhamnosus, B. longum; vaginal: L. rhamnosus, L. delbrueckii, L. acidophilus,
S. thermophilus). There was no preventive effect of either formulation vs. placebo (odds ratio
for oral formulation 1.06 [0.58-1.94] and vaginal formulation 1.38 [0.75-2.54]).45
One smaller randomized trial (N= 48) of an oral product with L. acidophilus and L. rhamnosus
demonstrated a significantly lower rate of symptomatic candidiasis after three months of
maintenance dosing.46 A slightly larger study (N = 95) of a vaginally delivered product con-
taining L. gasseri, L. fermentum, L. rhamnosus and P. acidilacti, used after an initial antifungal
treatment, did not demonstrate any reduction in symptomatic VVC recurrence one month
after treatment.47 A more comprehensive review of data can be found in a number of recent
reviews on this subject. 48-50
Prebiotics and synbiotics
A small study of 48 women with acute, culture-positive vaginal candidiasis and a history of re-
current VVC randomized participants to vaginal clotrimazole and a concurrent oral synbiotic
containing L. acidophilus, L. rhamnosus and lactoferrin or placebo. Participants continued main-
tenance dosing with the study product for 10 days a month for six months. Three months after
antifungal treatment, the synbiotic group had lower rates of recurrence (8.3 vs. 66.7%, p<0.01).46
Clinical recommendations
Given the lack of high-quality data to support the efficacy of probiotics, and the epidemi-
ologic evidence showing in vivo correlations between vaginal lactobacilli and VVC, we do
not recommend use of either oral or vaginal probiotics for treatment or prevention of VVC.
10.4
Aerobic vaginitis/desquamative inflammatory vaginitis
Probiotics
One study tested the inclusion of a probiotic in their approach to aerobic vaginitis (AV) ther-
apy. Heczko et al. screened women with a history of recurrent BV, and treated them with oral
metronidazole and 10 days of oral probiotics containing L. gasseri, L. plantarum, L. fermen-
tum (prOVag, IBSS, Poland).51 At follow-up those with aerobes (AV) or resistant Gardnerella
spp. by culture received targeted antibiotics and were randomized to 10 days of the oral
probiotic mix or placebo each month for three months, with clinical and culture testing each
month preformed one week after completion of the 10-day probiotic course. The authors
state that time to AV relapse was up to 76% (p<0.05) longer in the probiotic group. Heczko’s
191
study is challenging to interpret: inflammatory markers were not used to define AV cases
nor outcomes, and outcomes measured (recurrence of symptoms or positive culture at a
follow-up visits) do not distinguish BV from AV.
Prebiotics and synbiotics
While not strictly AV, insights into probiotic approach to correct dysbiosis and thus pre-
vent AV can be gained from studies of intermediate microbiota (IM). Women with IM have
Nugent scores from 4 – 6, are largely devoid of lactobacilli but do not have BV; they are often
found with itchy irritation and discharge, and many will go on to have AV. Patients with IM
are at risk for the same sequelae as those with AV. Russo et al. randomized 40 patients with
itching, irritation, discharge and IM to the oral synbiotic product containing L. acidophilus,
L. rhamnosus and 50 mg of bovine lactoferrin once daily for 15 days. Assessment at the end
of treatment found significantly less itching and discharge (p<0.001), synbiotic intragroup
normalization of Nugent score (p=0.0004) and reduction in Nugent score in the synbiotic
arm vs. placebo (p=0.0110). Quantitative real time polymerase chain reaction demonstrated
that lactobacilli were significantly increased after 15 days of the synbiotic, however data on
durability of this finding are not available.29
Clinical recommendations
There are few data upon which to base clinical recommendations, so we would approach
this similar to BV. For patients who have refractory symptoms and have failed standard ther-
apy, there may be biologic plausibility that a probiotic or synbiotic containing lactoferrin
could be helpful.
10.5
Trichomoniasis
Probiotics
Trichomonas vaginalis is the most common pathogenic protozoan in humans in industrial-
ized countries. 5-nitroimidazole treatment is the only effective treatment, however recurrent
infections are common, on some occasions due to re-infection, on others due to antibiotic
resistance. Metronidazole resistance occurs in up to 10% of cases of vaginal trichomonas.
T. vaginalis co-infection with BV is a frequent occurrence. In this setting metronidazole effec-
tiveness in treating trichomoniasis may be reduced and explained in part by the decreased
redox potential found in BV subjects.52, 53 In vitro evidence suggests that addition of a probi-
otic to metronidazole therapy increases the cure rate of metronidazole: Sgibnev et al. found
that co-culture of opportunistic bacteria with human derived lactobacilli, or L. rhamnosus
LCR35 supernatants containing hydrogen peroxide, lactic acid and surfactants, increased
the antibiotic sensitivity of opportunistic bacteria.54 The authors postulate that vaginal lacto-
bacilli administration could improve T. vaginalis’ in vivo sensitivity to metronidazole.
In a well-conducted clinical trial, Sgibnev et al. randomized 90 patients with T. vaginalis and
192
BV who failed prior therapy to receive metronidazole 500 mg twice a day and either vaginal
L. rhamnosus CR35 or vaginal placebo twice for seven days, then to continue the vaginal
probiotic twice daily for seven more days.55 Symptoms, pH, redox potential, Nugent score
and presence of T. vaginalis were assessed before the start of therapy and on the 4th, 8th
and 15th day of therapy. The authors report that women in the treatment arm found symp-
toms improved significantly, confirmed by vaginal examination, and a significant decrease
in T. vaginalis positive culture rate (6.8 vs. 47.6%) at the completion of metronidazole. As pH
decreased and redox potential increased more intensely in the probiotic arm, the authors
ascribe the improved cure rate to an increase in metronidazole effectiveness secondary to
physiochemical changes induced by this probiotic in the presence of BV.
Prebiotics and synbiotics
No clinical trials have been conducted to test efficacy of numerous food, marine and medic-
inal extracts that display strong anti-trichomonal activity in vitro. If in vitro activity translates
to in vivo and clinical activity these extracts could provide novel strategies to combat resist-
ant trichomonads.56
Clinical recommendations
When T. vaginalis is refractory to therapy, and when BV is present, ancillary probiotic therapy
may be considered.
10.6
Conclusion
It is biologically plausible that probiotics, prebiotics and synbiotics could improve treatment
and prevention of BV, especially in cases that are refractory to standard antibiotic therapy.
However, there is no consensus on the appropriate species, dose, formulation, delivery route
or duration of treatment. Given the out-of-pocket expense for many of these products, we
encourage caution in recommending their use, and follow up to assess efficacy.
Recommendations
Recommendation Quality of
evidence
Strength of
recommendation
Monotherapy with probiotics for bacterial vaginosis is not recommended. 2a B
There is no recommendation to prophylactically use probiotics or prebi-
otics in pregnancy. 3b B
There is no benefit of using probiotics in women with vulvovaginal
candidiasis. 2a B
A probiotic or synbiotic containing lactoferrin can be tried in refractory
cases of aerobic vaginitis/desquamative inflammatory vaginitis. 4 C
In refractory trichomoniasis, associated with bacterial vaginosis, probiot-
ics can be added to the treatment schemes. 4 C
193
References
1. Hill, C.; Guarner, F.; Reid, G.; Gibson, G. R.; Merenstein, D. J.; Pot, B.; Morelli, L.; Canani, R. B.; Flint, H. J.; Salminen, S.;
Calder, P. C.; Sanders, M. E., Expert consensus document. The International Scientific Association for Probiotics and
Prebiotics consensus statement on the scope and appropriate use of the term probiotic. Nat Rev Gastroenterol
Hepatol 2014, 11, (8), 506-14.
2. Gibson, G. R.; Hutkins, R.; Sanders, M. E.; Prescott, S. L.; Reimer, R. A.; Salminen, S. J.; Scott, K.; Stanton, C.; Swanson, K.
S.; Cani, P. D.; Verbeke, K.; Reid, G., Expert consensus document: The International Scientific Association for Probiotics
and Prebiotics (ISAPP) consensus statement on the definition and scope of prebiotics. Nat Rev Gastroenterol Hepatol
2017, 14, (8), 491-502.
3. Swanson, K. S.; Gibson, G. R.; Hutkins, R.; Reimer, R. A.; Reid, G.; Verbeke, K.; Scott, K. P.; Holscher, H. D.; Azad, M. B.;
Delzenne, N. M.; Sanders, M. E., The International Scientific Association for Probiotics and Prebiotics (ISAPP) con-
sensus statement on the definition and scope of synbiotics. Nat Rev Gastroenterol Hepatol 2020, 17, (11), 687-701.
4. Koirala, R.; Gargari, G.; Arioli, S.; Taverniti, V.; Fiore, W.; Grossi, E.; Anelli, G. M.; Cetin, I.; Guglielmetti, S., Effect of oral
consumption of capsules containing Lactobacillus paracasei LPC-S01 on the vaginal microbiota of healthy adult
women: a randomized, placebo-controlled, double-blind crossover study. FEMS Microbiol Ecol 2020, 96, (6).
5. Strus, M.; Chmielarczyk, A.; Kochan, P.; Adamski, P.; Chelmicki, Z.; Chelmicki, A.; Palucha, A.; Heczko, P. B., Studies on
the effects of probiotic Lactobacillus mixture given orally on vaginal and rectal colonization and on parameters
of vaginal health in women with intermediate vaginal flora. Eur J Obstet Gynecol Reprod Biol 2012, 163, (2), 210-5.
6. Bohbot, J. M.; Cardot, J. M., Vaginal impact of the oral administration of total freeze-dried culture of LCR 35 in healthy
women. Infect Dis Obstet Gynecol 2012, 2012, 503648.
7. Houng, H. S.; Noon, K. F.; Ou, J. T.; Baron, L. S., Expression of Vi antigen in Escherichia coli K-12: characterization of ViaB
from Citrobacter freundii and identity of ViaA with RcsB. J Bacteriol 1992, 174, (18), 5910-5.
8. De Alberti, D.; Russo, R.; Terruzzi, F.; Nobile, V.; Ouwehand, A. C., Lactobacilli vaginal colonisation after oral consumption
of Respecta((R)) complex: a randomised controlled pilot study. Archives of gynecology and obstetrics 2015, 292, (4), 861-7.
9. Chen, C.; Hao, L.; Zhang, Z.; Tian, L.; Zhang, X.; Zhu, J.; Jie, Z.; Tong, X.; Xiao, L.; Zhang, T.; Jin, X.; Xu, X.; Yang, H.; Wang,
J.; Kristiansen, K.; Jia, H., Cervicovaginal microbiome dynamics after taking oral probiotics. J Genet Genomics 2021,
48, (8), 716-726.
10. Anukam, K. C.; Osazuwa, E.; Osemene, G. I.; Ehigiagbe, F.; Bruce, A. W.; Reid, G., Clinical study comparing probiotic
LactobacillusGR-1 and RC-14 with metronidazole vaginal gel to treat symptomatic bacterial vaginosis. Microbes
Infect 2006, 8, (12-13), 2772-6.
11. Reznichenko, H.; Henyk, N.; Maliuk, V.; Khyzhnyak, T.; Tynna, Y.; Filipiuk, I.; Veresniuk, N.; Zubrytska, L.; Quintens, J.;
Richir, K.; Gerasymov, S., Oral Intake of Lactobacilli Can Be Helpful in Symptomatic Bacterial Vaginosis: A Rand-
omized Clinical Study. J Low Genit Tract Dis 2020, 24, (3), 284-289.
12. Vujic, G.; Jajac Knez, A.; Despot Stefanovic, V.; Kuzmic Vrbanovic, V., Efficacy of orally applied probiotic capsules for
bacterial vaginosis and other vaginal infections: a double-blind, randomized, placebo-controlled study. Eur J Obstet
Gynecol Reprod Biol 2013, 168, (1), 75-9.
13. Zhang, Y.; Lyu, J.; Ge, L.; Huang, L.; Peng, Z.; Liang, Y.; Zhang, X.; Fan, S., Probiotic Lacticaseibacillus rhamnosus GR-1
and Limosilactobacillus reuteri RC-14 as an Adjunctive Treatment for Bacterial Vaginosis Do Not Increase the Cure
Rate in a Chinese Cohort: A Prospective, Parallel-Group, Randomized, Controlled Study. Front Cell Infect Microbiol
2021, 11, 669901.
14. Bradshaw, C. S.; Pirotta, M.; De Guingand, D.; Hocking, J. S.; Morton, A. N.; Garland, S. M.; Fehler, G.; Morrow, A.; Walker,
S.; Vodstrcil, L. A.; Fairley, C. K., Efficacy of oral metronidazole with vaginal clindamycin or vaginal probiotic for bacte-
rial vaginosis: randomised placebo-controlled double-blind trial. PLoS One 2012, 7, (4), e34540.
15. Cohen, C. R.; Wierzbicki, M. R.; French, A. L.; Morris, S.; Newmann, S.; Reno, H.; Green, L.; Miller, S.; Powell, J.; Parks, T.; Hemmer-
ling, A., Randomized Trial of Lactin-V to Prevent Recurrence of Bacterial Vaginosis. N Engl J Med 2020, 382, (20), 1906-1915.
16. Eriksson, K.; Carlsson, B.; Forsum, U.; Larsson, P. G., A double -blind treatment study of bacterial vaginosis with normal
vaginal lactobacilli after an open treatment with vaginal clindamycin ovules. Acta Derm Venereol 2005, 85, (1), 42-6.
17. Larsson, P. G.; Stray-Pedersen, B.; Ryttig, K. R.; Larsen, S., Human lactobacilli as supplementation of clindamycin to
patients with bacterial vaginosis reduce the recurrence rate; a 6-month, double-blind, randomized, placebo-con-
trolled study. BMC Womens Health 2008, 8, 3.
18. Petricevic, L.; Witt, A., The role of Lactobacillus casei rhamnosus Lcr35 in restoring the normal vaginal flora after
antibiotic treatment of bacterial vaginosis. BJOG 2008, 115, (11), 1369-74.
19. Ya, W.; Reifer, C.; Miller, L. E., Efficacy of vaginal probiotic capsules for recurrent bacterial vaginosis: a double-blind,
randomized, placebo-controlled study. Am J Obstet Gynecol 2010, 203, (2), 120 e1-6.
20. Liu, H. F.; Yi, N., A systematic review and meta-analysis on the efficacy of probiotics for bacterial vaginosis. Eur Rev
Med Pharmacol Sci 2022, 26, (1), 90-98.
194
21. Tidbury, F. D.; Langhart, A.; Weidlinger, S.; Stute, P., Non-antibiotic treatment of bacterial vaginosis-a systematic
review. Arch Gynecol Obstet 2021, 303, (1), 37-45.
22. Munoz-Barreno, A.; Cabezas-Mera, F.; Tejera, E.; Machado, A., Comparative Effectiveness of Treatments for Bacterial
Vaginosis: A Network Meta-Analysis. Antibiotics (Basel) 2021, 10, (8).
23. Jeng, H. S.; Yan, T. R.; Chen, J. Y., Treating vaginitis with probiotics in non-pregnant females: A systematic review and
meta-analysis. Exp Ther Med 2020, 20, (4), 3749-3765.
24. Wang, Z.; He, Y.; Zheng, Y., Probiotics for the Treatment of Bacterial Vaginosis: A Meta-Analysis. Int J Environ Res Public
Health 2019, 16, (20).
25. Chen, X.; Lu, Y.; Chen, T.; Li, R., The Female Vaginal Microbiome in Health and Bacterial Vaginosis. Front Cell Infect
Microbiol 2021, 11, 631972.
26. Husain, S.; Allotey, J.; Drymoussi, Z.; Wilks, M.; Fernandez-Felix, B. M.; Whiley, A.; Dodds, J.; Thangaratinam, S.; Mc-
Court, C.; Prosdocimi, E. M.; Wade, W. G.; de Tejada, B. M.; Zamora, J.; Khan, K.; Millar, M., Effects of oral probiotic
supplements on vaginal microbiota during pregnancy: a randomised, double-blind, placebo-controlled trial with
microbiome analysis. Bjog 2020, 127, (2), 275-284.
27. Yang, S.; Reid, G.; Challis, J. R. G.; Gloor, G. B.; Asztalos, E.; Money, D.; Seney, S.; Bocking, A. D., Effect of Oral Probiotic
Lactobacillus rhamnosus GR-1 and Lactobacillus reuteri RC-14 on the Vaginal Microbiota, Cytokines and Chemok-
ines in Pregnant Women. Nutrients 2020, 12, (2).
28. Hakimi, S.; Farhan, F.; Farshbaf-Khalili, A.; Dehghan, P.; Javadzadeh, Y.; Abbasalizadeh, S.; Khalvati, B., The effect of
prebiotic vaginal gel with adjuvant oral metronidazole tablets on treatment and recurrence of bacterial vaginosis: a
triple-blind randomized controlled study. Arch Gynecol Obstet 2018, 297, (1), 109-116.
29. Russo, R.; Edu, A.; De Seta, F., Study on the effects of an oral lactobacilli and lactoferrin complex in women with
intermediate vaginal microbiota. Arch Gynecol Obstet 2018, 298, (1), 139-145.
30. Russo, R.; Karadja, E.; De Seta, F., Evidence-based mixture containing Lactobacillus strains and lactoferrin to prevent recur-
rent bacterial vaginosis: a double blind, placebo controlled, randomised clinical trial. Benef Microbes 2019, 10, (1), 19-26.
31. Baldacci, F.; Baldacci, M.; Bertini, M., Lactobacillus rhamnosus BMX 54 + Lactose, A Symbiotic Long-Lasting Vaginal
Approach to Improve Women’s Health. Int J Womens Health 2020, 12, 1099-1104.
32. Nasioudis, D.; Linhares, I. M.; Ledger, W. J.; Witkin, S. S., Bacterial vaginosis: a critical analysis of current knowledge.
Bjog 2017, 124, (1), 61-69.
33. Miranda, M.; Saccone, G.; Ammendola, A.; Salzano, E.; Iannicelli, M.; De Rosa, R.; Nazzaro, G.; Locci, M., Vaginal lactoferrin
in prevention of preterm birth in women with bacterial vaginosis. J Matern Fetal Neonatal Med 2021, 34, (22), 3704-3708.
34. Vieira-Baptista, P.; De Seta, F.; Verstraelen, H.; Ventolini, G.; Lonnee-Hoffmann, R.; Lev-Sagie, A., The Vaginal Microbi-
ome: V. Therapeutic Modalities of Vaginal Microbiome Engineering and Research Challenges. J Low Genit Tract Dis
2022, 26, (1), 99-104.
35. MacAlpine, J.; Daniel-Ivad, M.; Liu, Z.; Yano, J.; Revie, N. M.; Todd, R. T.; Stogios, P. J.; Sanchez, H.; O’Meara, T. R.; Tomp-
kins, T. A.; Savchenko, A.; Selmecki, A.; Veri, A. O.; Andes, D. R.; Fidel, P. L., Jr.; Robbins, N.; Nodwell, J.; Whitesell, L.; Cow-
en, L. E., A small molecule produced by Lactobacillus species blocks Candida albicans filamentation by inhibiting a
DYRK1-family kinase. Nat Commun 2021, 12, (1), 6151.
36. Rose Jorgensen, M.; Thestrup Rikvold, P.; Lichtenberg, M.; Ostrup Jensen, P.; Kragelund, C.; Twetman, S., Lactobacillus
rhamnosus strains of oral and vaginal origin show strong antifungal activity in vitro. J Oral Microbiol 2020, 12, (1), 1832832.
37. Jang, S. J.; Lee, K.; Kwon, B.; You, H. J.; Ko, G., Vaginal lactobacilli inhibit growth and hyphae formation of Candida
albicans. Sci Rep 2019, 9, (1), 8121.
38. De Gregorio, P. R.; Silva, J. A.; Marchesi, A.; Nader-Macias, M. E. F., Anti-Candida activity of beneficial vaginal lactoba-
cilli in in vitro assays and in a murine experimental model. FEMS Yeast Res 2019, 19, (2).
39. Baeten, J. M.; Hassan, W. M.; Chohan, V.; Richardson, B. A.; Mandaliya, K.; Ndinya-Achola, J. O.; Jaoko, W.; McClelland, R.
S., Prospective study of correlates of vaginal Lactobacillus colonisation among high-risk HIV-1 seronegative women.
Sex Transm Infect 2009, 85, (5), 348-53.
40. McClelland, R. S.; Richardson, B. A.; Hassan, W. M.; Graham, S. M.; Kiarie, J.; Baeten, J. M.; Mandaliya, K.; Jaoko, W.;
Ndinya-Achola, J. O.; Holmes, K. K., Prospective study of vaginal bacterial flora and other risk factors for vulvovaginal
candidiasis. J Infect Dis 2009, 199, (12), 1883-90.
41. Cotch, M. F.; Hillier, S. L.; Gibbs, R. S.; Eschenbach, D. A., Epidemiology and outcomes associated with moderate to
heavy Candida colonization during pregnancy. Vaginal I nfections and Prematurity Study Group. Am J Obstet Gynecol
1998, 178, (2), 374-80.
42. Tortelli, B. A.; Lewis, W. G.; Allsworth, J. E.; Member-Meneh, N.; Foster, L. R.; Reno, H. E.; Peipert, J. F.; Fay, J. C.; Lewis,
A. L., Associations between the vaginal microbiome and Candida colonization in women of reproductive age. Am J
Obstet Gynecol 2020, 222, (5), 471 e1-471 e9.
195
43. Brown, S. E.; Schwartz, J. A.; Robinson, C. K.; O’Hanlon, D. E.; Bradford, L. L.; He, X.; Mark, K. S.; Bruno, V. M.; Ravel, J.;
Brotman, R. M., The Vaginal Microbiota and Behavioral Factors Associated With Genital Candida albicans Detection
in Reproductive-Age Women. Sex Transm Dis 2019, 46, (11), 753-758.
44. Kovachev, S. M.; Vatcheva-Dobrevska, R. S., Local Probiotic Therapy for Vaginal Candida albicans Infections. Probiotics
Antimicrob Proteins 2015, 7, (1), 38-44.
45. Pirotta, M.; Gunn, J.; Chondros, P.; Grover, S.; O’Malley, P.; Hurley, S.; Garland, S., Effect of lactobacillus in preventing
post-antibiotic vulvovaginal candidiasis: a randomised controlled trial. BMJ 2004, 329, (7465), 548.
46. Russo, R.; Superti, F.; K aradja, E.; De Seta, F., Randomised clinical trial in women with Recurrent Vulvovaginal Candid-
iasis: Efficacy of probiotics and lactoferrin as maintenance treatment. Mycoses 2019, 62, (4), 328-335.
47. Ehrstrom, S.; Daroczy, K.; Rylander, E.; Samuelsson, C.; Johannesson, U.; Anzen, B.; Pahlson, C., Lactic acid bacteria
colonization and clinical outcome after probiotic supplementation in conventionally treated bacterial vaginosis
and vulvovaginal candidiasis. Microbes Infect 2010, 12, (10), 691-9.
48. Van de Wijgert, J.; Verwijs, M. C., Lactobacilli-containing vaginal probiotics to cure or prevent bacterial or fungal
vaginal dysbiosis: a systematic review and recommendations for future trial designs. Bjog 2020, 127, (2), 287-299.
49. Xie, H. Y.; Feng, D.; Wei, D. M.; Mei, L.; Chen, H.; Wang, X.; Fang, F., Probiotics for vulvovaginal candidiasis in non-preg-
nant women. Cochrane Database Syst Rev 2017, 11, CD010496.
50. Shenoy, A.; Gottlieb, A., Probiotics for oral and vulvovaginal candidiasis: A review. Dermatol Ther 2019, 32, (4), e12970.
51. Heczko, P. B.; Tomusiak, A.; Adamski, P.; Jakimiuk, A. J.; Stefański, G.; Mikołajczyk-Cichońska, A.; Suda-Szczurek, M.;
Strus, M., Supplementation of standard antibiotic therapy with oral probiotics for bacterial vaginosis and aerobic
vaginitis: a randomised, double-blind, placebo-controlled trial. BMC Womens Health 2015, 15, 115.
52. Gatski, M.; Martin, D. H.; Levison, J.; Mena, L.; Clark, R. A.; Murphy, M.; Henderson, H.; Schmidt, N.; Kissinger, P., The
influence of bacterial vaginosis on the response to Trichomonas vaginalis treatment among HIV-infected women.
Sex Transm Infect 2011, 87, (3), 205-8.
53. Holmes, K. K.; Chen, K. C.; Lipinski, C. M.; Eschenbach, D. A., Vaginal redox potential in bacterial vaginosis (nonspecific
vaginitis). J Infect Dis 1985, 152, (2), 379-82.
54. Sgibnev, A.; Kremleva, E., Influence of Hydrogen Peroxide, Lactic Acid, and Surfactants from Vaginal Lactobacilli on
the Antibiotic Sensitivity of Opportunistic Bacteria. Probiotics Antimicrob Proteins 2017, 9, (2), 131-141.
55. Sgibnev, A.; Kremleva, E., Probiotics in addition to metronidazole for treatment Trichomonas vaginalis in the pres-
ence of BV: a randomized, placebo-controlled, double-blind study. Eur J Clin Microbiol Infect Dis 2020, 39, (2), 345-351.
56. Friedman, M.; Tam, C. C.; Cheng, L. W.; Land, K. M., Anti-trichomonad activities of different compounds from foods,
marine products, and medicinal plants: a review. BMC Complement Med Ther 2020, 20, (1), 271.
196
Unrestricted educational grant:
SCYNEXIS
... V ulvovaginal candidiasis (VVC) is a common condition, with over half of the women population experiencing at least one episode during their lifetime (1,2). The current classification systems for VVC divide the disease into uncomplicated versus complicated or acute versus recurrent (3)(4)(5)(6)(7)(8)(9)(10). The bulk of VVC cases is acute and uncomplicated, occurring in otherwise healthy premenopausal women as symptomatic vaginitis due to Candida albicans. ...
... International and national VVC guidelines vary in their classification system, definition of recurrent, chronic, and severe, requirement for confirmatory tests, inclusion of postme nopausal women as an affected group, and management recommendations (Table 1) (3)(4)(5)(6)(7)(8)(9)(10). No VVC guideline explicitly names vulvar cutaneous candidiasis, illustrates it with a photograph, details the clinical presentation, and outlines its histopathologic features. ...
... The Society of Obstetricians and Gynecolo gists of Canada practice recommendations mirror the CDC, except for the use of four episodes as the threshold for recurrent VVC (RVVC) and the application of 'severe' as a descriptor for symptoms rather than signs (5). The International Society for the Study of Vulvovaginal Disease (ISSVD) recommendations also draw on the CDC guideline for the classification and definition of RVVC and management protocols (6). The ISSVD document describes a positive yeast culture as crucial before initiating therapy. ...
Article
Vulvovaginal candidiasis (VVC) affects over half of women during their lifetime. There are two categorization systems for VVC: uncomplicated versus complicated and acute versus recurrent. Most uncomplicated or acute cases occur in postpubertal premenopausal girls and women as sporadic vaginitis due to Candida albicans . Complicated VVC includes recurrent, chronic, or severe cases, presence of non- albicans species, and/or disease occurring in people with diabetes, immunosuppression, or pregnancy. These classification systems fail to distinguish the two distinct clinical categories of genital candidiasis: estrogen-dependent VVC and estrogen-independent cutaneous candidiasis. These entities are characterized by different pathogenesis, patient demographics, predisposing conditions, symptoms, signs, investigations, differential diagnosis, treatment, and ancillary measures. The current international and national guidelines on VVC are inadequate in their description of the clinical presentation, role and limitations of culture, biopsy findings, and management of cutaneous candidiasis. Progress toward improved patient outcomes will require the interdisciplinary collaboration of researchers and guideline authors to separate these two entities, unify terminology for each, explore the roles of medications and comorbid dermatoses, detail pragmatic and accessible diagnostic processes, define treatment goals, and discuss the long-term management strategies pertinent to each condition.
... BV and AV pose significant risks, including obstetrical complications and an increased risk of sexually transmitted infection acquisition (e.g., human papilloma virus, human immunodeficiency virus, Trichomonas vaginalis, and Chlamydia trachomatis). Therefore, it is crucial to treat these infections promptly [28]. ...
... The aim of BV and AV treatment is to restore the balance of the vaginal flora and thereby limit the excessive growth of harmful microorganisms. Current international guidelines recommend the use of antibiotics, in particular, metronidazole or clindamycin, topically in the first instance and then, depending on disease evolution, systemically [28][29][30][31][32][33][34]. Local treatments are generally preferred as a first-line treatment because oral options can result in systemic side effects such as headache and gastro-intestinal disorders. ...
... Local treatments are generally preferred as a first-line treatment because oral options can result in systemic side effects such as headache and gastro-intestinal disorders. In contrast, vaginal applications can reach local concentrations up to 30 times higher, resulting in comparable or slightly improved cure rates with fewer side effects [28]. As clindamycin is active against both staphylococci and streptococci, as well as anaerobes, the International Union Against Sexually Transmitted Infections states that the best treatment for uncomplicated AV is currently vaginally administered clindamycin, whereas metronidazole if advocated for the treatment of persistent and recurrent BV [3,28,29]. ...
Article
Full-text available
Background/Objectives: Aerobic vaginitis (AV) and bacterial vaginosis (BV) are vaginal infections requiring the fast elimination of pathogens. The frequent confusion of these infections may justify the use of a rapidly acting broad-spectrum antibiotic treatment. Methods: This study investigated the bactericidal kinetics of the neomycin-polymyxin B-nystatin (NPN) combination compared to those of two reference antibiotics (clindamycin and metronidazole) against 22 bacteria commonly implicated in AV and BV. Results: NPN exhibited bactericidal activity against the aerobic Gram-positive bacteria, with particularly high bactericidal activity being observed against streptococci, S. aureus, and C. amycolatum after 1 h at low dilutions and after 4 h for all dilutions. Enterococci were less sensitive to NPN. Clindamycin demonstrated poor rapid bactericidal activity against all Gram-positive bacteria tested. NPN manifested high bactericidal activity against all aerobic Gram-negative bacteria tested, whereas clindamycin showed bactericidal activity only after 4 h at a 1/2 dilution. With respect to the four anaerobic strains tested, NPN demonstrated high bactericidal activity at all tested dilutions with concentration-dependent effects. Metronidazole exhibited lower or no rapid bactericidal activity. Conclusions: These results suggest that NPN has very fast bactericidal action against the main bacteria involved in AV and BV compared to clindamycin and metronidazole, highlighting its potential in managing bacterial vaginal infections.
... Moreover, a substantial proportion (5-8%) suffer from recurrent VC (RVC), experiencing four or more episodes per year [2]. The symptoms of VC extend beyond physical manifestations such as vaginal discharge, itching, and burning; they also significantly affect mental, sexual, and social health [3][4][5]. Epidemiological data indicate that RVC poses a significant global public health challenge, impacting over 130 million women annually [1]. Additionally, the diagnosis and treatment of VC, particularly RVC, impose significant economic burdens due to social withdrawal and workforce loss [1,6]. ...
... Despite their impact, Candida species are typically a low-abundance component of the vaginal flora and may be present at an increased abundance in 20-30% of reproductive-aged women without symptoms (Candida colonization) [7]. Historically, the primary pathogen in both acute and recurrent VC (AVC and RVC, respectively) is C. albicans (85-95%) [3,8,9]. However, misdiagnosis and improper treatment, including the use of over-the-counter antifungals, have contributed to the rise in treatment-refractory C. albicans and non-albicans VC [10][11][12]. ...
... Azole antifungals, such as fluconazole, show decreased activity at pH 4 [17,18]. Therefore, conducting AFST at pH 4, reflecting vaginal acidity, provides a more realistic assessment of antifungal efficacy [3,9,14]. ...
Article
Full-text available
Vaginal candidiasis (VC) is a prevalent condition among women of reproductive age and poses a significant global public health challenge. However, the disease is often diagnosed and treated without mycological information. We investigated the epidemiology, laboratory diagnostics, and antifungal susceptibility of VC. We included 300 women from Çukurova University Obstetrics and Gynecology outpatient clinic in Adana, Türkiye. Participants underwent a health survey and provided vaginal swab samples for microscopic examination and fungal culture. The microscopic analysis involved wet-mount and gram-stained slides, whereas fungal identification involved CHROMAgar Candida, matrix-assisted laser desorption/ionization time-of-flight mass spectrometry (MALDI-TOF MS), and real-time polymerase chain reaction high-resolution melting analysis (RT-PCR HRMA). Antifungal susceptibility tests were conducted at pH 7 and 4 using the CLSI document M44-A2. Of the 106 women with positive fungal cultures, 86.8% were diagnosed with VC, whereas 13.2% showed Candida colonization. Among those with VC, 55.4% had acute and 44.6% had recurrent VC; a family history of allergies increased the risk for both types. We recovered 115 yeast isolates, predominantly C. albicans, C. glabrata, and C. krusei. Diagnostic accuracy of CHROMAgar Candida was 91.3% for the most common isolates, and HRMA was consistent in differential diagnosis. Antifungal resistance varied with pH; susceptibility to fluconazole, itraconazole, and ketoconazole decreased at pH 4, whereas susceptibility to miconazole increased. Our findings underscore the need for a diagnostic algorithm and enhanced collaboration between clinicians and microbiologists to improve VC management. Recommendations include using Gram staining, CHROMAgar Candida, MALDI-TOF MS, and antifungal susceptibility tests at both pH levels.
... 10,11 Therefore, conjugated equine estrogen remains a recommended treatment option for vulvovaginal atrophy in women. 12 The histological report confirmed exuberant granulation tissue. No foreign body or residual surgical suture was identified in the specimen. ...
... In contrast, topical conjugated equine estrogen vaginal cream is patient-administered, rejuvenates the vaginal tissue amidst the relative hypoestrogenic state that occurs in postpartum and breastfeeding mothers, 7 and this informed the preference for its use after surgical excision. Although topical vaginal estrogen may be safe, 26 it is best administered as a low-or ultralow-dose preparation (such as conjugated equine estrogen, estriol, estradiol, and promestriene [3-propyl 17β -methyl diether estradiol]) 12 to minimize systemic absorption that can increase endometrial thickness. While honey is effective in the treatment of obstetric wounds and shows promise in animal studies as a protective agent for genital atrophy, there are no long-term studies and definitive conclusions about its efficacy in the treatment of urogenital atrophy in humans as in the index patient with atrophic vaginal mucosa. ...
Article
Full-text available
Background Poor wound healing may limit body functionality and is an indication for clinical intervention. Excessive formation of granulation tissue above the edge of the skin surrounding a wound without re-epithelization is termed exuberant granulation, or proud flesh. It is uncommonly reported as a complication of an episiotomy wound. Aim This study aimed to report exuberant granulation that complicates an episiotomy wound with a friable vaginal epithelium and to describe the successful treatment of the lesion with surgical excision and topical conjugated equine estrogen vaginal cream. Case Report A 24-year-old para 1 had spontaneous vaginal birth of a normal baby at term in a district hospital. Five months later, she presented to a regional hospital with complaints of pain and incomplete wound healing at the episiotomy site. She had used topical povidone-iodine ointment with no success. Following a physical examination, an exuberant granulation at the episiotomy wound was diagnosed. The lesions were located mostly at 5 to 7 o’clock position in the vagina which had a thin and friable mucosa. The patient was treated with surgical excision and postoperative topical conjugated equine estrogen vaginal cream 0.625 mg per 1 g at a dose of 0.5 g per intravaginal application twice weekly for two weeks, and thereafter once weekly for one week. A review after 6 weeks, 12 weeks, and 6 months confirmed complete wound healing and normal function of the genitalia. Conclusion Exuberant granulation that complicates an episiotomy wound with friable vaginal mucosa is amenable to surgical excision and postoperative intermittent intermediate doses of topical conjugated equine estrogen vaginal cream.
... Another possible explanation for this decrease in consumption may be a lower prevalence of infections due to reduced opportunities for human contact due to the implementation of non-pharmaceutical interventions, such as lockdowns where gatherings and meetings were banned and public spaces were closed. This reduction in infection rates could have led to an overall decrease in consumption of broad-spectrum antibiotics, which is a main risk factor associated with recurrent Vulvovaginal Candidosis [11]. ...
... We wonder if this reflects a medical preference or if there is another reason for this decrease. The International Society for the Study of Vulvovaginal Disease indicates isoconazole vaginal suppository as one of the available treatment options for patients with uncomplicated Vulvovaginal Candidosis (mild symptoms); however, it does not mention vaginal cream [11]. ...
Article
Full-text available
Background/Objectives: Excessive or inadequate use of antimicrobial drugs may lead to the emergence of resistant strains. For this reason, it is important to monitor consumption indicators to assess drugs’ utilization over time. This study aimed to analyze the consumption of medically prescribed azole antifungal drugs in mainland Portugal from 2014 to 2023, focusing on those directed to genital infections: fluconazole, isoconazole, itraconazole, and sertaconazole. Methods: For each drug, the evaluated parameters were the total number of packages, number of packages per 1000 inhabitants, defined daily dose (DDD) per 1000 inhabitants per day, and total costs. For this purpose, we used data from community pharmacies’ sales, which are available through INFARMED (the Portuguese national authority on medicines and health products). Results: Several trends emerged from data analysis. The COVID-19 pandemic negatively affected the consumption of all azole antifungal drugs included in this study. However, after 2020, fluconazole and sertaconazole consumption has been increasing. In the specific case of fluconazole, there was an increase in expenditure, although the total number of packages suffered a decrease over the 10-year study period. Additionally, the defined daily dose (DDD) per 1000 inhabitants per day for fluconazole and itraconazole was lower compared to estimates from the last available survey (2009). Conclusions: Although our findings represent a lesser pressure on fungi, further monitoring is needed to better understand the evolution of fluconazole and itraconazole consumption over time, particularly due to the trends observed in this study.
... В заключение необходимо отметить рекомендации Международного общества по изучению вульвовагинальных заболеваний (ISSVD, 2023), согласно которым в основе диагностики ГУМС также лежит оценка клинико-анамнестических данных с дополнительной оценкой микроскопии мазка и определением уровня вагинального pH [14]. ...
... В настоящее время, по данным отечественных и международных клинических рекомендаций, «золотым стандартом» лечения ГУМС является локальная гормональная терапия [10,13]. При этом, согласно рекомендациям ISSVD (2023), имеется несколько вариантов локальной гормональной терапии, представленных в табл. 2 [14]. ...
Article
Genitourinary syndrome of menopause still represents one of the most pressing unresolved issues of modern gynecology. High prevalence and the data suggesting a marked decline in the quality of life in general and sexual life in particular in patients of this cohort demonstrate the need to develop effective clinical management programs. The review provided considers the key aspects of genitourinary syndrome of menopause as a medical and social problem, reports modern approaches to clinical management in the wake of new international guidelines.
... В основную группу вошла 41 пациентка 18-45 лет с жалобами на патологические выделения из половых путей и с установленным диагнозом БВ. Согласно рекомендациям Международного общества по изучению вульвовагинальных заболеваний (International Society for the Study of Vulvovaginal Disease -ISSVD) [7] диагноз БВ устанавливали на основании критериев Амселя [8] при наличии 3 из 4 перечисленных признаков: ...
Article
Full-text available
Aim. To evaluate the effectiveness of treatment of bacterial vaginosis (BV) with the nifuratel+nystatin complex and its effect on the lactobacillus profile (LBP) in the vagina. Materials and methods. A multicenter, prospective, observational study included 41 patients diagnosed with BV (main group) who were treated with vaginal capsules containing 500 mg of nifuratel and 200,000 IU of nystatin (Macmiror® Complex). Efficacy was assessed at 2 weeks and 4–6 weeks after treatment based on clinical data (absence of pathological discharge). The control group for the comparative assessment of LBP included 25 healthy women. Lactobacillus index (LI, the proportion of lactobacilli in the total bacterial mass) and the abundance of Lactobacillus crispatus, L. gasseri, L. jensenii, and L. iners were assessed using a polymerase chain reaction before treatment with BV, at 2 weeks and 4–6 weeks after the end of therapy, as well as in the control group. Additionally, the vaginal pH level was measured. Results. Clinical recovery was achieved in all main group patients and persisted throughout the observation. In all patients in this group, lactobacilli comprised 0.01–100% of the total bacterial mass. The frequency of lactobacilli species in the control group was relatively uniform, with a relatively uniform abundance of lactobacilli and a slight dominance of L. crispatus, with frequent isolation of two species in samples and high LI (70%) in 80% of females. Before treatment in the main group, the majority of vaginal samples contained one species of lactobacilli; 40% had LI70%, and in 62.5%, L. iners was isolated. LBP did not change significantly within 2 weeks after treatment, and after 4–6 weeks, it became comparable to that in the control group. A decrease in pH was associated with LBP recovery. However, there was no strong correlation between LI and individual species of lactobacilli in samples with or without BV. Conclusion. Nifuratel+nystatin complex is highly effective in BV treatment. The treatment had no adverse effect on LBP, which did not differ significantly immediately after treatment compared to the baseline. However, 4–6 weeks after the treatment, LBP recovered spontaneously, reaching comparable indicators with a group of clinically healthy females regarding species composition and LI.
... Кроме того, это состояние связано с неблагоприятными последствиями для беременности и плода, включая внутриутробные инфекции, преждевременный разрыв плодных оболочек, поздний самопроизвольный выкидыш, преждевременные роды, а также послеродовые и послеабортные гнойно-септические осложнения [3,4]. В настоящее время первой линией лечения БВ остаются антибактериальные препараты, такие как пероральный метронидазол по 500 мг дважды в день в течение семи дней, интравагинальный метронидазол по 5 г один раз в день в течение пяти дней, а также интравагинальный клиндамицин по 5 г один раз в день в течение семи дней [5]. Однако известно, что рецидивы БВ после стандартной антибактериальной терапии возникают у 43% женщин в течение трёх месяцев и у 58% -в течение 12 месяцев, что может быть связано с устойчивостью к антимикробной терапии, формированием биоплёнок и последующей персистенцией микроорганизмов, а также реинфекцией [6]. ...
Article
BACKGROUND. Bacterial vaginosis is a common infectious non-inflammatory vaginal disease that increases the likelihood of contracting sexually transmitted infections, which has a negative impact on perinatal outcomes and generally reduces the quality of life. Given the low long-term efficacy of antibiotic therapy, as well as the high recurrence rate and side effects associated with antibiotic use, there is a need to find alternative approaches to its treatment. AIM. To evaluate the efficacy and tolerability of a comprehensive two-stage treatment of BV including clindamycin or dequalinium chloride and lactic acid in women of reproductive age. METHODS. An open randomized clinical study was carried out, which involved 54 women aged 18 to 45 years with a diagnosis of bacterial vaginosis confirmed by Amsel's criteria. The participants were randomly assigned into three groups: 17 women from the first group used lactic acid, 20 women from the second group used a combination of clindamycin and lactic acid, and 17 women from the third group used dequalinium chloride with lactic acid. Treatment efficacy was evaluated after 14 days using Amsel's criteria. Three months after the completion of treatment, complaints were assessed and vaginal pH was measured. RESULTS. During the study, it was noted that two weeks after the completion of treatment, whitish-gray vaginal discharge ceased in all women from the second and third groups. In the first group, where only lactic acid was used, discharge continued in three patients. Positive dynamics in the pH change of vaginal secretions were observed in all groups both two weeks and three months after the end of treatment, with the most noticeable effect in women using dequalinium chloride together with lactic acid. Three months after the end of treatment, complaints of vaginal discharge persisted in two women from the first group and two from the second group. CONCLUSION. The conducted study confirmed the high effectiveness of two-stage treatment. However, the combination of dequalinium chloride and lactic acid demonstrated more stable results in both the short and long term, making this method of non-antibacterial therapy the most preferable in the context of the growing problem of antibiotic resistance.
Article
Aim: to assess the effectiveness of combination therapy in reproductive age patients with mixed vaginitis. Materials and Methods. A retrospective analysis of the therapeutic effectiveness for mixed vaginitis in 158 patients aged 18 to 45 years was carried out. For treatment, the drugs Gainomax (standard dose – thioconazole 100 mg, tinidazole 150 mg in the form of a suppository) and Gainomax plus (standard dose – thioconazole 200 mg, tinidazole 300 mg, lidocaine 100 mg in the form of a suppository) were used for 3 days. The criteria for treatment effectiveness were disappearance of clinical symptoms and normalization of laboratory parameters. Results. Clinical and laboratory efficacy of therapy reached 95.6 %. Lack of therapeutic effectiveness presented as persistent complaints from patients was recorded in 7 out of 158 (4.4 %) of patients, at the same time, the normalization of vaginal pH was established (pH = 4.1 ± 0.4), the Lactobacilli count increased up to 10 ⁷ ± 2.8 CFU/cm ³ , eradication of other opportunistic bacteria and fungi Candida occurred. No side effects were reported while assessing drug safety and tolerability. Conclusion. Therapy combining thioconazole and tinidazole in treatment of patients with mixed vaginitis demonstrated high effectiveness.
Article
Full-text available
The literature review presents modern aspects of the etiology and pathogenesis of urogenital candidiasis, highlights issues of immunological resistance and the significance of polymorphic and pathological gene loci, the functioning of which leads to insufficient activation of the immune system in the disease. The mechanisms of formation of biofilms by fungi of the genus Candida on mucous membranes and the role of lactobacillary microflora in preventing their spread, as well as ideas about the prerequisites for the development of vulvovaginal candidiasis are considered. The modern possibilities of drug therapy for urogenital candidiasis, presented in domestic and foreign clinical recommendations, are discussed. The advantages of prescribing fenticonazole, which is active against various types of fungi of the genus Candida, as well as a wide range of anaerobes and microorganisms associated with bacterial vaginosis, are considered, and the results of studies on the effectiveness of the drug are analyzed.
Article
Full-text available
Objective: To assess whether CO2- laser treatment is more effective than sham application in relieving the Most Bothersome Symptom (MBS) in women with Genitourinary Syndrome of Menopause (GSM). Design: Single center, sham controlled, double-blind, randomised trial. Setting: A tertiary center in Belgium. Population: Sixty women with moderate to severe GSM symptoms. Methods: All participants eventually received three consecutive laser and three consecutive sham applications, either first laser followed by sham, or conversely. Main outcome measures: The primary outcome was the participant-reported change in severity of the MBS at 12 weeks. Secondary outcomes included subjective (patient satisfaction, sexual function, urinary function) and objective (pH, Vaginal Health Index Score, in vivo microscopy) measurements assessing the short-term effect and the longevity of treatment effects at 18 months after start of the therapy. Adverse events were reported at every visit. Results: The MBS severity score decreased from 2.86(0.35) to 2.17(0.93) (-23.60%; CI -36.10 to -11.10) in women treated with laser as compared to 2.90(0.31) to 2.52(0.78) (-13.20%; CI -22.70 to -3.73) in those receiving sham applications (p=0.13). There were no serious adverse events reported up to 18 months. Conclusions: In women with GSM, the treatment response 12 weeks after laser application was comparable to that of sham applications. There were neither obvious differences for secondary outcomes. No serious adverse events were reported.
Article
Full-text available
Bacterial vaginosis (BV), the overgrowth of diverse anaerobic bacteria in the vagina, is the most common cause of vaginal symptoms worldwide. BV frequently recurs after antibiotic therapy, and the best probiotic treatments only result in transient changes from BV-associated states to “optimal” communities dominated by a single species of Lactobacillus. Therefore, additional treatment strategies are needed to durably alter vaginal microbiota composition for patients with BV. Vaginal microbiota transplantation (VMT), the transfer of vaginal fluid from a healthy person with an optimal vaginal microbiota to a recipient with BV, has been proposed as one such alternative. However, VMT carries potential risks, necessitating strict safety precautions. Here, we present an FDA-approved donor screening protocol and detailed methodology for donation collection, storage, screening, and analysis of VMT material. We find that Lactobacillus viability is maintained for over six months in donated material stored at − 80 °C without glycerol or other cryoprotectants. We further show that species-specific quantitative PCR for L. crispatus and L. iners can be used as a rapid initial screening strategy to identify potential donors with optimal vaginal microbiomes. Together, this work lays the foundation for designing safe, reproducible trials of VMT as a treatment for BV.
Article
Full-text available
Vulvovaginitis is a common and challenging gynaecological problem in prepubertal and pubertal girls. Such an infection, owing to a wide range of aetiologies, if not responding to hygienic measures, needs further investigation through vaginal cultures, since treatment should be tailored accordingly. This study aimed to investigate the pathogens isolated in prepubertal and pubertal girls with signs and symptoms of vulvovaginitis. A total of 2314 symptomatic girls, 1094 prepubertal and 1220 pubertal, aged 2 to 16 years, were included. Vaginal samples were inoculated on specific culture plates followed by incubation in aerobic, anaerobic or CO2 atmosphere at 37 °C for 24 or 48 h, as appropriate. The identification of the isolated pathogens was carried out using Gram stain, conventional methods and the automated system VITEK 2 (BioMerieux, Marcy l’Etoile, France). Positive cultures were obtained from 587 (53.7%) of prepubertal girls and 926 (75.9%) of pubertal girls. A total of 613 and 984 pathogens were detected in prepubertal and pubertal subjects, respectively. Isolated bacteria included 40.1% and 22.8% Gram-positive cocci, 35.6% and 24.8% Gram-negative rods in the prepubertal and pubertal groups, respectively, with faecal pathogens being the most prevalent. Bacterial vaginosis was diagnosed in 22.8% of prepubertal and 37.9% of pubertal girls. Candida species were isolated mostly in the pubertal girls (14.5%). Conclusion: Culture results should be evaluated with caution in children with vulvovaginitis. In the prepubertal girls, the most common isolated pathogens were opportunistic bacteria of faecal origin while girls in late puberty were more susceptible to bacterial vaginosis and vulvovaginal candidiasis.What is Known: • Vulvovaginitis is the most frequent and challenging reason for referral to paediatric and adolescent gynaecology services. • Microbiological examination can prove to be a significant tool to help diagnosis although results should be evaluated with caution in children. What is New: • Significantly more positive vaginal cultures and pathogens were recorded in symptomatic pubertal girls compared to prepubertal children. • The prevalence of bacterial vaginosis was increased in both prepubertal and pubertal girls with vulvovaginitis although significantly more in girls at puberty.
Article
Full-text available
Background Women treated for breast cancer (BC) often suffer genitourinary syndrome of menopause. These symptoms may be alleviated by vaginal estrogen therapy (VET) or menopausal hormone therapy (MHT). However, there are concerns of risks of recurrence of BC and death following treatment. Methods Our study included longitudinal data from a national cohort of postmenopausal women, diagnosed 1997-2004 with early-stage invasive estrogen receptor–positive nonmetastatic BC, who received no treatment or 5 years of adjuvant endocrine therapy. We ascertained prescription data on hormone therapy, VET or MHT, from a national prescription registry. We evaluated mortality and risk of recurrence associated with use of VET and MHT vs non-use using multivariable models adjusted for potential confounders. Results Among 8461 women who had not received VET or MHT before BC diagnosis, 1957 and 133 used VET and MHT, respectively, after diagnosis. Median follow-up was 9.8 years for recurrence and 15.2 years for mortality. The adjusted relative risk of recurrence was 1.08 (95% confidence interval [CI] = 0.89 to 1.32) for VET (1.39 [95% CI = 1.04 to 1.85 in the subgroup receiving adjuvant aromatase inhibitors]) and 1.05 (95% CI = 0.62 to 1.78) for MHT. The adjusted hazard ratios for overall mortality were 0.78 (95% CI = 0.71 to 0.87) and 0.94 (95% CI = 0.70 to 1.26) for VET and MHT, respectively. Conclusions In postmenopausal women treated for early-stage estrogen receptor–positive BC, neither VET nor MHT was associated with increased risk of recurrence or mortality. A subgroup analysis revealed an increased risk of recurrence, but not mortality, in patients receiving VET with adjuvant aromatase inhibitors.
Article
Full-text available
Despite being one of the most common sexually transmitted infections (STIs) in the United States, the epidemiology of trichomoniasis remains understudied. One population that has been historically overlooked regarding STIs is that of older adults, despite many individuals remaining sexually active well into their older years. We investigated the reported prevalence and incidence of trichomoniasis in adults aged ≥45 years in the United States using a systematic literature review. Twelve articles were included in the review, all assessing prevalence of trichomoniasis in this age group. Notably, no included articles assessed trichomoniasis incidence. Data collected encompassed several decades, from 1993 to 2016. Estimates of infection prevalence varied widely and ranged from 0.2% to 21.4% in included populations, with the highest prevalence typically seen among individuals seeking diagnostic testing for STIs. Several studies found increased risk for trichomoniasis in older patients compared to younger age groups. This is the first review to examine the risk of trichomoniasis in older adults, and the surprisingly high prevalence suggests that older adults may merit increased screening for trichomoniasis and sexual health education.
Article
Full-text available
Introduction: Single-dose 2-g oral secnidazole (SEC), newly approved by the U.S. Food and Drug administration (FDA) for treatment of trichomoniasis, is a potent 5-nitroimidazole with selective toxicity against various micro-organisms. It has been used internationally to treat trichomoniasis, bacterial vaginosis, and other infections for decades. Trichomoniasis is the most common non-viral sexually transmitted infection worldwide and is associated with significant morbidity. In comparison to the only other FDA-approved treatments for trichomoniasis in the United States-metronidazole and tinidazole-SEC has favorable pharmacokinetics, including a longer half-life and a lower minimal lethal concentration. Areas covered: This work summarizes the chemistry and pharmacology of SEC and reviews the evidence on its efficacy, tolerability, and safety for the treatment of trichomoniasis. Expert opinion: SEC is an efficacious, well tolerated, and safe treatment for patients aged ≥12 years with trichomoniasis. Single-dose administration makes it a favorable treatment option for patients, especially in cases where adherence to multi-dose treatment regimens may be low.
Article
Background Trichomonas vaginalis (T. vaginalis) is an extracellular flagellated protozoan parasitizing the human genital and urinary tracts. T. vaginalis infection impacts human reproductive function, but whether it causes infertility is still a matter of debate. Methods In this work, we consulted 205 relevant articles, which were classified into three categories: epidemiological investigations (100), review articles (43), and research articles (62). RevMan 5.4 was used to conduct a meta-analysis of the articles reporting epidemiological investigations comparing the incidence of T. vaginalis infection between infertile and fertile groups. Review and research articles were used to summarize the pathogenesis of infertility caused by T. vaginalis. Results The results indicated that rate of T. vaginalis infection in the infertile group was significantly higher than that in the fertile group. Moreover, the epidemiological surveys showed that the infertility rates of population infected with T. vaginalis were significantly higher than that of population without T. vaginalis infection. Nine out of ten (90%) related review articles stated that T. vaginalis infection causes infertility, and the review and research articles indicated the main pathogenic mechanisms of infertility caused by T. vaginalis were as follows: T. vaginalis impairs sperm quality, resulting in infertility; the immune response triggered by T. vaginalis infection impacts human reproductive function. Conclusion Our results confirm that there is a correlation between T. vaginalis infection and infertility, and T. vaginalis infection can lead to infertility. The study provides a foundation for further investigations into its pathogenesis.
Article
(Abstracted from JAMA 2021;326:1381–1389) An estimated 40% to 60% of women experience vaginal symptoms associated with menopause, which can lead to physical discomfort, as well as negatively affecting sexual function, relationships, and quality of life. Preliminary data have shown that fractional laser treatments used as nonhormonal alternatives for postmenopausal vaginal symptoms result in improvement in vaginal atrophy after treatment.