Guidelines
2018 European (IUSTI/WHO)
International Union against sexually
transmitted infections (IUSTI) World
Health Organisation (WHO) guideline on
the management of vaginal discharge
Jackie Sherrard
1
, Janet Wilson
2
, Gilbert Donders
3
,
Werner Mendling
4
and Jørgen S Jensen
5
Abstract
Four common pathological conditions are associated with vaginal discharge: bacterial vaginosis, aerobic vaginitis, can-
didosis, and the sexually transmitted infection, trichomoniasis. Chlamydial or gonococcal cervical infection may result in
vaginal discharge. Vaginal discharge may be caused by a range of other physiological and pathological conditions including
atrophic vaginitis, desquamative inflammatory vaginitis, cervicitis, and mucoid ectopy. Psychosexual problems may pre-
sent with recurrent episodes of vaginal discharge and vulval burning. These need to be considered if tests for specific
infections are negative. Many of the symptoms and signs are non-specific and a number of women may have other
conditions such as vulval dermatoses or allergic and irritant reactions.
Keywords
Trichomoniasis (Trichomonas vaginalis), women, diagnosis, bacterial vaginosis, candida, aerobic vaginitis, vaginal discharge
Date received: 30 May 2018; accepted: 5 June 2018
Introduction
Four common pathological conditions are associated
with vaginal discharge: bacterial vaginosis (BV), aero-
bic vaginitis (AV), candidosis, and the sexually
transmitted infection, trichomoniasis. Chlamydial or
gonococcal cervical infection may result in vaginal dis-
charge. Vaginal discharge may be caused by a range of
other physiological and pathological conditions includ-
ing atrophic vaginitis, desquamative inflammatory
vaginitis (DIV), cervicitis, and mucoid ectopy.
Psychosexual problems may present with recurrent epi-
sodes of vaginal discharge and vulval burning. These
need to be considered if tests for specific infections are
negative. Many of the symptoms and signs are
non-specific and a number of women may have other
conditions such as vulval dermatoses or allergic and
irritant reactions.
1
Department of Genitourinary Medicine, Sexual Health Department,
Buckinghamshire Healthcare NHS Trust, Amersham, UK
2
Department of Genitourinary Medicine, Leeds Teaching Hospitals NHS
Trust, Leeds, UK
3
Department of Obstetrics and Gynecology, Regional Hospital H Hart
Tienen, University Hospital Antwerp
4
Infektionen in Gyn
akologie und Geburtshilfe, Wuppertal, Germany
5
Research Unit for Reproductive Microbiology, Statens Serum Institut,
Copenhagen, Denmark
Lead editor: Jørgen S Jensen
This guideline is an update of the European IUSTI vaginal discharge
guideline 2011.
Corresponding author:
Jackie Sherrard, Department of Genitourinary Medicine, Sexual Health
Department, Buckinghamshire Healthcare NHS Trust, Amersham, UK.
International Journal of STD & AIDS
2018, Vol. 29(13) 1258–1272
! The Author(s) 2018
Article reuse guidelines:
sagepub.com/journals-permissions
DOI: 10.1177/0956462418785451
journals.sagepub.com/home/std
Aetiology and transmission
BV
BV is the commonest cause of abnormal vaginal dis-
charge in woman of childbearing age but may also be
encountered in perimenopausal women.
1,2
In
Caucasian women the prevalence is 5–15%; in Black
women it is higher at 45–55%. Women who have s ex
with women (WSW) share similar lactobacillary types,
are more likely to have concordant vaginal microbiota
(flora) patterns, and are at increased risk for BV.
3
BV is a dysbiosis of the vaginal microbiota. It is
characterised by an overgrowth of predominantly
anaerobic organisms (e.g. Gardnerella vaginalis,
Prevotella spp., Atopobium vaginae, Mycoplasma hom-
inis, Mobiluncus spp.) in the vagina leading to a
replacement of lactobacilli and an increase in vaginal
pH. Bacterial identification using PCR has demonstrat-
ed that there are many different, previously unculti-
vated bacteria present in women with BV including
BV-associated bacterium 1, 2, and 3, and Sneathia spe-
cies.
4
Since these bacteria are difficult to culture, the
antibiotic susceptibility of many is not known.
BV can arise and remit spontaneously and although
not strictly considered a sexually transmitted infection
(STI) it is associated with sexual activity. The exact
aetiology of BV is still unclear but current evidence
suggests that formation of a biofilm with G. vaginalis
is important in the switch from normal vaginal flora
to BV.
5,6
AV/DIV
AV presents with a purulent discharge, some degree of
atrophy, and vaginitis. Lactobacilli are decreased and
pH is elevated, but aerobic microorganisms, like
Escherichia coli, group B streptococci, and
Staphylococcus aureus predominate.
7
Mixed infections
are frequent. It is not known whether AV has an infec-
tious origin or whether it is an inflammatory process
followed by a dysbiosis. It can cause long-term symp-
toms with intermittent exacerbations, and recurrences
after treatment are common.
8
Atrophic vaginitis in lac-
tating women is probably a variant of AV. More severe
forms of AV and DIV are probably the same condition.
Candidosis
More than 60% of healthy premenopausal women are
colonised with candida, with higher rates in pregnancy,
and lower rates in children and postmenopausal
women without hormonal replacement therapy.
9,10
An estimated 75% of women will experience at least
one symptomatic episode during their lifetime and 6–
9% will experience chronic recurrent vulvovaginal can-
didosis (at least four episodes per year). Vulvovaginal
candidosis results from an overgrowth of Candida albi-
cans in 90% of women (remainder with other species,
e.g. Candida glabrata).
11,12
Precipitating factors include
antibiotic therapy, pregnancy, and endogenous or
exogenous immunosuppression (including diabetes
mellitus and immunosuppressive medication). In some
women, symptoms may occur with a low burden of
candida and it is thought this may be due to an allergic
or inflammatory response to the yeast.
Trichomoniasis
Trichomonas vaginalis (TV) is a flagellated protozoon,
which is a parasite of the genital tract. In adults, it is
almost exclusively sexually transmitted. Due to site spe-
cificity, infection only follows intravaginal or intraure-
thral inoculation of the organism. In women urethral
infection is present in 90% of episodes, although the
urinary tract is the sole site of infection in <5% of
cases. The most obvious host response to infection is
a local increase in polymorphonuclear leucocytes.
Table 1. Symptoms and signs.
Bacterial vaginosis Aerobic vaginitis Candidosis Trichomoniasis
Approximately 50%
asymptomatic
10–20% asymptomatic Approx. 60% women colonised.
Minority develop symptoms
10–50% asymptomatic and
5–15% no abnormal signs
Thin white homogenous
discharge, coating walls
of vagina and vestibule
Purulent discharge Vaginal discharge may be curdy
(non-offensive)
Offensive vaginal discharge in up
to 70% frothy and yellow in
10–30%
Offensive fishy odour Vulval burning or stinging Vulval soreness/itching
and erythema
Vulval itching/irritation
and erythema
Absence of vaginitis Superficial dyspareunia Vulval fissuring Dysuria
Vaginal er ythema and
oedema
Superficial dyspareunia Rarely low abdominal discomfort
Vaginal ulceration Satellite skin lesions Vaginitis
Vulval oedema Approx. 2% ‘strawberry’ cervix
visible to naked eye.
Sherrard et al. 1259
Clinical features
There are recognised symptoms and signs (Table 1).
The diagnosis of both BV and candidosis is syndromic,
i.e. based on clinical symptoms and signs supported by
laboratory test findings, which in themselves vary in
specificity and sensitivity. The classical features of TV
are frequently absent or non-specific.
13,14
Complications
Women with BV are at increased risk of acquiring
STIs. They have a two-fold increased risk of HIV
acquisition,
15
1.5–2-fold risk of Chlamydia
16
and gon-
orrhoea,
16
a nine-fold risk of TV,
17
and a two-fold risk
of herpes simplex virus type 2 (HSV-2)
18
compared to
women without BV. HIV-positive women with BV
have a three-fold risk of transmitting HIV.
19
Monthly
prophylaxis with metronidazole reduces the incidence
of STIs by almost 50%.
20
The BV-associated bacteria
are probably also implicated in the aetiology of pelvic
inflammatory disease (PID). A prospective study of
women with clinically-suspected PID reported
significant correlations between the presence of
BV-associated bacteria and the presence of endometri-
tis and recurrent PID.
21
There is an association with BV and post-
hysterectomy vaginal cuff infection,
22,23
post-abortion
endometritis,
24,25
and an increased risk of spontaneous
miscarriage and preterm birth.
26,27
Symptomatic preg-
nant women with BV should be treated in the usual
way but the latest Cochrane review concludes there is
insufficient evidence to recommend routine screening
and treating all pregnant women for asymptomatic
BV to prevent preterm birth.
28
Multiple reports support an epidemiological associ-
ation between HIV and trichomoniasis. There is grow-
ing evidence that trichomonal infection enhances HIV
transmission
29–32
and there may be an increased risk of
TV infection in those that are HIV-positive.
33
Trichomoniasis is associated with adverse pregnancy
outcomes.
34,35
The literature on metronidazole treat-
ment during pregnancy and preterm birth is not con-
clusive. The most recent Cochrane review found that
metronidazole is effective against trichomoniasis when
taken by women and their partners during pregnancy,
but it may harm the baby due to early birth.
36
Screening of asymptomatic individuals for TV infection
is therefore not currently recommended.
Although only recently described, moderate/severe
AV is associated with an increasing number of co-
infections and complications.
37
An increased risk of
preterm delivery and chorioamnionitis in women with
first trimester AV has been described.
38
Several studies in the last decade have shown a
decrease in preterm birth, if vaginal candida colonisa-
tion or infection had been treated with clotrimazole.
39
In a study by Holzer et al.
40
women who were colonised
with candida spp. during the second trimester of preg-
nancy had higher rates of preterm birth and lower neo-
natal birthweight than those who were colonised
during the first trimester of their pregnancy.
According to old studies the vaginal treatment of an
asymptomatic candida colonisation during the last six
weeks of pregnancy reduces the candida colonisation of
the newborn during vaginal delivery and thus reduces
oral thrush and napkin dermatitis of the baby during
the first four weeks of life.
41
Modern studies are urgent-
ly needed to confirm these findings.
Diagnosis
Women presenting with abnormal vulval or vaginal
symptoms should be tested to ensure that appropriate
treatment is given.
40,42–44
If this is not possible, then
examination and testing shoul d definitely be performed
in the following situations:
Severe or recurrent symptoms
Failure of vaginal discharge to respond to empiri-
cal treatment
Symptoms in pregnancy
Finding of TV on cervical cytology
Diagnosis of TV in sexual partner
Asymptomatic women do not require testing for BV,
AV, or candida. Testing asymptomatic women for TV
should be guided by local prevalence data.
The definitive diagnosis of each infection is based
upon clinical symptoms, examination, pH and micro-
scopic findings of the vaginal secretions, and for TV
additionally by laboratory tests. A sample of the dis-
charge is removed from the vaginal wall with a swab.
This can be performed by the clinician or be self-
collected by the woman.
45
The type of swab is not
important. An elevated pH (>4.5) is suggestive of BV
or trichomoniasis and is almost always normal in can-
dida infections. Direct microscopy should be done
immediately, if available.
BV
Gram-stained microscopy is the reference method for
diagnosing BV.
A. Nugent score
46
This is used as a gold standard for
studies and relies upon estimating the relative pro-
portions of bacterial morphotypes on a Gram-
stained vaginal smear to give a score between 0
1260 International Journal of STD & AIDS 29(13)
and 10. A score of <4 is normal, 4–6 is intermedi-
ate, and >6 is BV. However, it does not take bac-
terial morphotypes other than those associated with
BV into account. The clinical implications of ‘inter-
mediate flora’ are unclear but they are associated
with complications.
47
B. Hay–Ison criteria
48
These are also based on the
findings on a Gram-stained smear but are easier and
quicker to use in clinical practice and do include
non-BV-associated bacteria .
Grade 0: Not related to BV, epithelial cells only, no
lactobacilli, indicates recent antibiotics
Grade 1: (Normal): Lactobacillus morphotypes
predominate
Grade 2: (Intermediate): Mixed flora with some lac-
tobacilli present, but Gardnerella or Mobiluncus mor-
photypes also present
Grade 3 (BV): Predominantly Gardnerella and/or
Mobiluncus morphotypes, clue cells. Few or absent
Lactobacilli.
Grade 4: Not related to BV, Gram-positive cocci
only, no lactobacilli (AV flora)
Clinical criteria for diagnosis of BV (Amsel)
49
The presence of three of the four criteria is required;
as three are clinical criteria it is possible to make a
diagnosis of BV without microscopy or the use of a
microbiology laboratory. Compared to Gram-stained
microscopy, the presence of three of the four clinical
criteria has a sensitivity of 60–72% for the diagnosis
of BV.
50,51
1. Homogeneous grey-white discharge
2. pH of vaginal fluid > 4.5 (measured using narrow
gauge pH paper)
3. Fishy odour (if not recognisable, use few drops of
10% KOH)
4. Clue cells present on wet mount microscopy (>20%
of all epithelial cells)
Other methods of diagnosing BV
Commercial tests for BV are also available. OSOM
BV Blue (Sekisui Diagnostics, Framingham, MA,
USA) is a point-of-care test which measures sialidase
levels and has sensitivity of 91.7% compared to micros-
copy.
52
The BD MAX
TM
Vaginal Panel (Becton,
Dickinson and company, Franklin Lakes, NJ, USA)
is a microbiome-based, nucleic acid amplification
assay that detects BV, TV, and several candida species.
The manufacturer insert quotes a sensitivity of 90.7%
for the diagnosis of BV.
53
The Guidelines Group recommends that the current
best test to diagnose BV in women is microscopy using
the Hay–Ison Criteria.
Strength of recommendation: Grade 1, quality of evi-
dence: Grade A.
AV
Microscopy.
The gold standard for diagnosis is wet mount
microscopy.
54
The AV score combines information
about bacterial flora, epithelial disruption, and
inflammation creating a score from 0 to 10: 0–2
(no AV), 3–4 (mild AV), 5–6 (moderate AV), or
7–10 (severe AV) (Table 2).
Cultures . Although most women with AV have positive
cultures for aerobic bacteria such as Streptococcus aga-
lactiae, S. aureus, E. coli, a positive vaginal culture does
not indicate the woman has AV and is not recom-
mended for diagnosis. However, culture with antimi-
crobial susceptibility testing may aid in treatment.
Molecular detection. Tests based on molecular biology
are being developed which correlate well with moder-
ate-to-severe AV compared with microscopy but need
confirmation in larger trials assessing sensitivity and
specificity.
55
The Guidelines Group recommends that the current
best test to diagnose AV in women is mic roscopy.
Strength of recommendation: Grade 2, quality of evi-
dence: Grade B.
Table 2. Abbreviated template for assessing the aerobic vagi-
nitis score.
8
Enter score
Background flora
Unremarkable 0
Small coliforms 1
Cocci or chains 2
Lactobacillary grade
Predominant lactobacilli 0
Reduced lactobacilli 1
No lactobacilli 2
Number of leucocytes
<10/high-powered field 0
10/epithelial cells 1
>10/epithelial cells 2
Toxic leucocyte proportion
None or sporadic 0
50% leucocytes 1
>50% leucocytes 2
Parabasal cells proportion
None 0
10% epithelial cells 1
>10% epithelial cells 2
Sherrard et al. 1261
Candidosis
Microscopy.
Budding cells (and a positive candida culture) can
exist in asymptomatic, colonised women or in
women with candidosis. The diagnosis should be
based on a combination of the clinical signs and
microscopic findings. Pseudohyphae/mycelia are evi-
dence of candidosis.
55–58
Yeasts or pseudohyphae on wet mount microscopy
with either saline or 10–20% KOH solution (40–
60% sensitivity) of vaginal discharge.
Yeasts or pseudohyphae on Gram stain (up to 65%
sensitivity) of vaginal discharge.
Culture.
Vaginal culture positive for a candida species. If pos-
sible this should be delineated as C. albicans or non-
albicans. If directly inoculated on to a Sabouraud’s
plate results should be reported as light, medium, or
heavy growth as this correlates with specificity.
As a high number of women carry candida asymp-
tomatically, the significance of light and medium
growths should be interpreted with caution.
Repeated culture of the same species of non-albicans
Candida (usually C. glabrata) indicates reduced anti-
fungal susceptibility to azoles.
The Guidelines Group recommends that the current best
test to diagnose candida in women is microscopy.
Strength of r ecommendation: Grade 1, quality of
evidence: Grade B.
Trichomoniasis
Microscopy. Direct observation of the organism by wet
mount microscopy (normal saline) or acridine orange-
stained slide from the posterior vaginal fornix. The wet
preparation should be read within 10 min of collection,
as the trichomonads quickly lose motility and will be
more difficult to identify.
59
The sensitivity is highest in
women presenting with vaginal discharge. However,
the sensitivi ty is reported to be as low as 45–60%
60–62
so a negative result should be interpreted with caution.
The specificity with trained personnel is high.
Point-of-care tests. A number of point-of-care tests that
have the advantages of microscopy have been
described. The OSOM Trichomonas Rapid Test
(Sekisui Diagnostics, Framingham, MA, USA) has
demonstrated a sensitivity of 80–94% and a specificity
greater than 95%.
63,64
This test requires no instrumen-
tation and provides a result within 30 min and is a
suitable alternative to culture or molecular testing.
Although these tests are more sensitive than vaginal
wet mount microscopy, false positives might occur,
especially in populations with a low prevalence of
disease, so consideration should be given to confirming
positives in that situation.
Culture. Culture of TV has a higher sensitivity com-
pared to microscopy but is not widely available.
A commercially available culture system (InPouch
TV; BioMed Diagnostics, White City, OR, USA),
offers many advantages over previous culture media
such as Diamond’s medium.
65–67
Once inocul ated the
pouches can be transferred to the laboratory for incu-
bation and the entire pouch read microscopically each
day for five days, negating the need to prepare wet
preparations every day that only sample a portion of
the culture medium.
Molecular detection. Nucleic acid amplification tests
(NAATs) offer the highest sensitivity for the detection
of TV in comparison to both microscopy and cul-
ture.
68,69
They should be the test of choice where
resources allow. NAATs can detect TV in vaginal or
endocervical swabs and in urine samples from women
with sensitivities of 88–97% and specificities of 98–
99%, depending on the specimen and reference stan-
dard.
70–73
The Guidelines Group recommends that the current
best tests to diagnose TV in women are NAATs.
Strength of r ecommendation: Grade 1, quality of
evidence: Grade A.
Management
BV
It should be explained that the cause is unclear and that
although there is increasing evidence of an association
with sexual activity, and of sexual transmissibility, it is
not yet proven to be a an STI.
Indications for treatment of BV:
Symptoms
Positive direct microscopy with/without symptoms
in some pregnant women (those with a history of
prior idiopathic preterm birth or second trimes-
ter loss)
BV in women undergoing gynaecological surgical or
invasive diagnostic procedures
Optional: positive direct microscopy in women with-
out symptoms. They may report a beneficial change in
their discharge following treatment.
Recommended regimens for BV:
Metronidazole 400–500 mg orally twice daily for 5–7
days
or
1262 International Journal of STD & AIDS 29(13)
Intravaginal metronidazole gel (0.75%) once daily
for five days
or
Intravaginal clindamycin cream (2%) once daily for
seven days
Alternative regimens for BV:
Metronidazole 2 g orally in a single dose
or
Tinidazole 2 g orally in a single dose
or
Tinidazole 1 g orally for five days
or
Clindamycin 300 mg orally twice daily for seven days
or
Dequalinium chloride 10 mg vaginal tablet one daily
for six days
For BV, single dose therapies have lower cure rates
than prolonged treatment. Oral metronidazole for
seven days has a significantly highe r cure rate
than single dose treatment (88% versus 54%
74
and
82% versus 62%
75
at 3–4 weeks after completion of
therapy). Fourteen days of oral metronidazole com-
pared with seven days showed improved cure initially
but there was no difference in cure rates 21 days after
completion of therapy.
76
A systematic review of trials
comparing clindamycin versus metronidazole conclud-
ed they have equal efficacy, whether oral or vaginal
formulations, both after one week (combined RR
1.01, 95% CI 0.69–1.46) and after one month (com-
bined RR 0.91, 95% CI 0.70–1.18). Roughly, 58–88%
will be cured after five days treatment with metronida-
zole or clindamycin. However, in terms of side effects,
in most studies clindamycin tended to have fewer
adverse effects than metronidazole (RR 0.75, 95% CI
0.56–1.02). Combining seven days of oral metronida-
zole with vaginal clindamycin cream did not improve
the cure rate compared with seven days oral metroni-
dazole with placebo.
77
Vaginal dequalinium seems to
have similar cure rates to vaginal clindamycin cream.
78
The effectiveness of metronidazole and clindamycin is
the same, but the cost of oral metronidazole is signifi-
cantly less than vaginal metronidazole which is cheaper
than clindamycin vaginal cream, with dequalinium
being the most expensive. Oral metronidazole has
more side effects than the other treatments but post-
treatment symptomatic candida is more common with
intravaginal treatments.
Clindamycin cream as well as metronidazole gel
contain mineral oils that are known to diminish the
strength of condoms. Therefore, use of barrier contra-
ception is not considered safe during the treatment with
any of these vaginal products.
The Guidelines Group recommends that 5–7 days of
topical or oral metronidazole or seven days of intravagi-
nal clindamycin can be considered first line for uncom-
plicated BV in women depending on personal choice and
circumstances. Cost-effectiveness of the recommended
regimens should be considered when adapting the
guideline for local use.
Strength of recommendation: Grade 1, quality of
evidence: Grade A.
Recurrent BV
A longitudinal study of women, following treatment of
BV with oral metronidazole for seven days, reported
BV recurrence rates of 23% at one month, 43% at
three months, and 58% at 12 months.
79
BV is associ-
ated with smoking and vaginal douching
80
but there is
no evidence that stopping these reduces BV. Studies of
consistent condom use have shown a 50% reduction in
BV incidence; the combined oral contraceptive pill is
associated with a 16% reduction and progestogen
depot injections/implants are associated with a
19% reduction in BV incidence.
81
Small studies have
reported an increased incidence of BV with the copper
intrauterine contraceptive device but it is not known
what effect, if any, progestogen- containing levonorges-
trel intrauterine systems have on BV incidence.
Recurrence of BV is associated with a new or multiple
male partners and having had a female partner.
A number of trials have evaluated intravaginal and
oral therapies to reduce BV recurrences.
Intravaginal metronidazole. A placebo-controlled trial
using twice-weekly metronidazole vaginal gel or place-
bo for 16 weeks reported a significant reduction in BV
recurrence. The relative risk at 16 and 28 weeks was
RR 0.43 (95% CI 0.25–0.73) and RR 0.68 (95% CI
0.49–0.93) with 70 and 39%, and 34 and 18% of
women being BV free at 16 weeks and 28 weeks, respec-
tively. Episodes of candidosis were more common with
metronidazole gel.
82
Another placebo-controlled trial
assessed vaginal pessaries containing metronidazole
750 mg plus miconazole 200 mg with matched placebo
for five nights per month for 12 months. The women
were evaluated every two months and the proportion of
visits with BV compared to placebo were 21.2 and
32.5%; RR 0.65 (95% CI 0.48–0.87). There was no
increase in candidosis with the intervention.
83
Oral metronidazole. A placebo-controlled trial assessed
the effect of monthly oral treatment (m etronidazole
2 g plus fluconazole 150 mg) versus placebo for
12 months: the intervention reduced the incidence of
BV (hazard ratio 0.55 [95% CI 0.49–0.63]).
84
Sherrard et al. 1263
Intravaginal lactate gel. In a small placebo-controlled
trial of intravaginal lactate gel 5 ml used for three
days after menses for six months, 88% of women
using the lactate gel were BV free compared with
10% using placebo.
85
Probiotics. In a systematic review of probiotics for the
treatment of BV, the authors concluded that the results
do not provide sufficient evidence for or against
recommending probiotics for the treatment of BV.
86
A subsequent meta-analysis concluded probiotic inter-
ventions were effective for treatment and prevention of
BV but the quality of the studies varied.
87
More good
quality research is needed to strengthen the body of
evidence needed for application by clinicians.
The Guidelines Group recommends that the current
best treatment for persistent and recurrent BV in
women is intravaginal metronidazole.
Strength of recommendation: Grade 2, quality of evi-
dence: Grade B.
AV/DIV
Indications for treatment of AV/DIV:
In one study, 5% of women presenting with vaginal
discharge had AV scores of 5 and over.
8
However,
these were a very heterogeneous group and specific
pathologies such as atrophic change, lichen planus,
and lichen sclerosus should be identified and treated
appropriately.
Recommended regimens for AV:
Two per cent clindamycin cream 5 g intravaginally
for 7–21 days
8,88
Combination use of intravaginal clindamycin and
intravaginal steroids,
88
e.g. hydrocortisone 300–
500 mg intravaginally for 7–21 days or Predfoam
enema applied intravaginally (off-label use) for
more severe cases
In cases with a significant atrophy component, local
oestrogens can be add ed
Clindamycin is active against staphylococci and
streptococci as well as anaerobes. Other antimicrobials
which are used with success in AV include kanamycin
ovules or moxifloxacin.
The Guidelines Group recommends that the current
best treatment for uncomplicated AV in women is clin-
damycin cream.
Strength of recommendation: Grade 2, quality of evi-
dence: Grade C.
Vaginal candidosis
Indications for therapy of candidosis:
Symptomatic women found to have candida on
either microscopy or culture.
Asymptomatic women do not require treatment
Asymptomatic male partners do not
require treatment
Recommended regimens for vaginal candidosis:
9,89,90
Oral preparations include
Fluconazole 150 mg as a single dose
Itraconazole 200 mg twice daily for one day
Intravaginal treatments include
Clotrimazole vaginal tablet 500 mg as single dose or
200 mg once daily for three days
Miconazole vaginal ovule 1200 mg as a single dose
or 400 mg once daily for three days
Econazole vaginal pessary 150 mg as a single dose
Treatment with azoles results in relief of symptoms
and negative cultures among 80–90% of patients after
treatment is completed, whether administered orally or
intravaginally. Only topical preparations should be used
during pregnancy. Overall, standard single-dose treat-
ments are as effective as longer courses. In a severely
symptomatic attack there is proven to be better symp-
tomatic benefit in repeating fluconazole 150 mg after
three days.
91
This does not affect relapse rates.
There are a number of other intravaginal preparat ions
available which are all either azoles, of limited availabil-
ity, e.g. nystatin, or unlicensed. There are limited data to
suggest that vulval treatment maybe of added benefit to
intravaginal treatment.
92
Where itch is a significant
symptom a hydrocortisone-containing topical prepara-
tion may provide more rapid symptomatic relief. Any
benefit may be from the emollient effect. If oral antifun-
gals are used, then a moisturising cream is cheaper and
may be less likely to give an irritant reaction.
The Guidelines Group recommends that the current
best treatment for uncomplicated candida in women is
a single-dose azole (oral or vaginal).
Strength of r ecommendation: Grade 1, quality of
evidence: Grade A.
Recurrent candidosis
Defined as four or more symptomatic episodes
per year
93,94
Document frequency, establish diagnosis and con-
firm by culture: all such women should have at
least one speciated culture.
1264 International Journal of STD & AIDS 29(13)
Exclude risk factors (e.g. diabetes, underlying immu-
nodeficiency, corticosteroid use, frequent antibiot-
ic use).
Consider other diagnoses vulval dermatitis/
eczema/vestibulodynia are common either coexisting
or as a differential diagnosis.
Maintenance therapy needs to be given frequently
enough to prevent vaginal regrowth, but the optimal
dosing interval is not clear. There are differing opinions
on how aggressive maintenance therapy should be
weekly or monthly treatments
93,95
and comparative
trials have not been undertaken. The long-term anti-
fungal regimen aims to prevent two essential pathoge-
netic mechanisms: increased risk of recolonisation and
increased risk of transformation to a symptomatic state
primarily as a function of pathologic host intolerance
of the candida.
96
Current recommendations are for an initial intensive
regimen of fluconazole 150–200 mg daily for three days
to attempt mycologic remission before initiating a
maintenance regimen. Published maintenance regimens
include oral fluconazole (i.e. 100, 150, or 200 mg dose)
weekly for six months
93
or 200 mg fluconazole weekly
for two months, followed by 200 mg biweekly for four
months, and 200 mg monthly for six months, according
to the individual response to therapy.
95
If these regi-
mens are not feasible, topical treatments used intermit-
tently can also be considered.
Treatment of persistent vaginal yeast infection due
to species other than C. albicans is particularly
challenging.
97
General advice includes the use of a vulval moistur-
iser applied to dry skin and washed off as a soap sub-
stitute. Ovulation suppressing progesterone
contraception, e.g. medroxyprogesterone acetate
(Depo Provera), nomegestrol, or desogestrel, may
have some benefits in particular women but the evi-
dence for this is poor.
98
The Guidelines Group recommends that the current
best treatment for persistent and recurrent candida in
women is a three-day induction course of an azole
followed by long-term maintenance suppressive regimen
for at least six months.
Strength of recommendation: Grade 2, quality of evi-
dence: Grade C.
TV
As TV is a sexually transmitted organism, screening for
coexistent STIs should be undertaken. Sexual absti-
nence should be advised until treatment of all partners
is completed.
Indications for therapy of TV:
Positive test for TV regardless of symptoms
Epidemiological treatment of sexual partners
Recommended regimens fo r TV:
99–101
First choice:
Metronidazole 400–500 mg orally twice daily for 5–7
days
or
Metronidazole 2 g orally in a single dose
or
Tinidazole 2 g orally in a single dose
The nitroimidazoles are the only class of drugs
useful for the oral or parenteral therapy of trichomo-
niasis and most strains are highly susceptible. Due to
high rates of infection of the urethra and paraurethral
glands in women systemic chemotherapy should be
given to effect a cure. The use of metronidazole gel is
not recommended. Oral single dose treatment is asso-
ciated with more frequent side effects than longer treat-
ment and a recent meta-analysis
99
indicated higher
treatment failure for single-dose compared to multi-
dose. In patients with true metronidazole allergy,
desensitisation has been used.
102,103
Patients should be advised not to take alcohol for
the duration of treatment and for at least 48 h (72 h for
tinidazole) afterwards because of the possibility of a
disulfiram-like (Antabuse
V
R
effect) CSL Biotherapies,
New Zealand reaction.
The Guidelines Group recommends that the current
best treatment for uncomplicated TV in women are nitro-
imidazoles (metronidazole or tinidazole).
Strength of recommendation: Grade 1, quality of evi-
dence: Grade A.
Persistent TV. Persistent or recurrent TV is due to inad-
equate therapy,
104
reinfection, or resistance. Check for
compliance and exclude vomiting of metronidazole and
exclude the possibility of reinfection from new or
untreated partners.
Treatment protocol for non-response to standard
TV therapy (having excluded reinfection and
non-adherence)
1. Repeat course of seven-day s tandard therapy
Metronidazole 400–500 mg twice daily for seven
days in those who failed to respond to a first
course of treatment, 40% responded to a repeat
course of standard treatment.
104
2. Higher-dose course of nitroimidazole
Metronidazole or tinidazole 2 g daily for 5–
7 days
105
Sherrard et al. 1265
Metronidazole 800 mg three times daily for seven
days in those who failed to respond to a second
course of treatment, 70% responded to a higher-
dose course of metronidazole.
104
For those failing this regimen, resistance testing
should be performed if available as improved outcomes
were reported with a treatment protocol guided by the
results of a resistance test.
104
If resistance testing is not
available high-dose tinidazole regimens are recom-
mended as in the above study; 65% of women with
clinical treatment failure did not have tinidazole-
resistant isolates and 83% of those receiving the rec-
ommended high-dose treatment were cured compared
with 57% of women receiving a lower-than-recom-
mended dose.
105
Tinidazole has a longer serum half-
life, good tissue penetration, a better side effect profile,
and lower levels of resistance than metronidazole so
should be used when infections have not responded
to metronidazole even though it is more expensive.
3. Very high-dose course of tinidazole
Tinidazole 1 g twice or three times daily, or 2 g
twice daily for 14 days intravaginal tinidazole
500 mg twice daily for 14 days
105–107
–inthose
who had failed other treatments 92 and 90%
responded to a very high-dose course of tinidazole.
If very high-dose tinidazole has been unsuccessful it
is difficult to recommend specific further treatment.
There are anecdotal reports of treatment success with
a number of other treatments. The reports are based on
success in one or two women who had usually received
a wide variety of prior treatments. Consequently, for
each successful anecdote there are a number of reports
of treatment failure.
The Guidelines Group recommends that the current
best treatment for persistent and recurrent TV in women
is repeated course of nitroimidazole at a higher dose.
Strength of recommendation: Grade B, quality of evi-
dence: Grade B.
Management during pregnancy and
breast feeding
A recent retrospective, case–control study found an
association between the use of a number of antibiotics
prescribed in the first trimester of pregnancy and spon-
taneous abortion. Statistically significant associations
were found with metronidazole. Clindamycin was not
tested in this study. Sexually transmitted genital infec-
tions themselves can cause pregnancy loss so failure to
treat them effectively may also result in spontaneous
abortion. The associations found might result from
women being prescribed the antibiotics for genital
infections with the increased risk of pregnancy loss
being due to the infections rat her than the antibiotics,
i.e. confounding by indication.
108
Meta-analyses have concluded that there is no evi-
dence of teratogenicity from the use of metronidazole
in women during the first trimester of pregnancy.
109–112
Metronidazole can be used in all stages of pregnancy
and during breast feeding. Symptomatic women with
TV and BV shoul d be treated at diagnosis, although
some clinicians have preferred to defer treatment until
the second trimester. The British National Formulary
advises against high-dose regimens in pregnancy.
Metronidazole enters breast milk and may affect its
taste. The manufacturers recommend avoiding high
doses if breastfeeding or if using a single dose of met-
ronidazole, breastfeeding should be discontinued for
12–24 h to reduce infant exposure.
Tinidazole is pregnancy category C (animal studies
have demonstrated an adverse event, and no adequate,
well-controlled studies in pregnant women have been
conducted), and its safety in pregnant women has not
been well evaluated. The manufacturer states that the
use of tinidazole in the first trimester is contraindicated.
Topical azoles can be used at any stage of pregnancy
for treatment of symptomatic candidosis. Oral flucon-
azole is associated with early abortions and Fallot
tetralogy, if administered in the first weeks of pregnan-
cy.
113,114
There appears to be less risk with oral prep-
arations after the first trimester.
The Guidelines Group recommends that the current best
treatment for TV in pregnant women is metronidazole.
Strength of r ecommendation: Grade 1, quality of
evidence: Grade A.
The Guidelines Group recommends that the current
best treatment for BV in pregnant women is clindamycin.
Strength of r ecommendation: Grade 2, quality of
evidence: Grade C.
The Guidelines Group recommends that the current
best treatment for candida in pregnant women are
topical azole preparations.
Strength of recommendation: Grade 1, quality of evi-
dence: Grade B.
Management of sexual partners
BV
A systematic review assessing the effectiveness of anti-
biotic treatment for male sexual partners of women
treated for BV concluded that antibiotic treatment
does not lead to a lower recurrence rate in the
women.
115
Routine screening and treatment of male
partners is therefore not indicated.
1266 International Journal of STD & AIDS 29(13)
In WSW, regular female partners frequently have
concordant vaginal microbiota so if one has BV the
partner is more likely to also have BV. It is thought
this is from sexual behaviours that transfer vaginal
secretions between them.
3
If a WSW is found to have
BV, and she has a regular female partner, it would be
reasonable to suggest that her partner be checked for
BV and be treated if positive although there is no evi-
dence that this will reduce BV recurrences.
The Guidelines Group recommends that the current
advice for women diagnosed with BV is that male
sexual partners do not require treatment. Female part-
ners may be treated if they have BV.
Strength of recommendation: Grade 2, quality of evi-
dence: Grade B.
Candidosis and AV
Routine screening and treatment of male partner(s) is
not indicated.
116,117
The Guidelines Group recommends that the current
advice for women diagnosed with candidosis or is that
sexual partners, do not require treatment.
Strength of recommendation: Grade 1, quality of evi-
dence: Grade B.
Trichomoniasis
Current sexual partners should be screened for STIs
and treated for TV regardless of the results of their
tests.
118,119
Patients should be instructed to avoid sex
until they and their sex partners are cured (i.e. when
therapy has been completed and patient and partner[s]
are asymptomatic).
The Guidelines Group recommends that the current
advice for women diagnosed with TV, is that their
sexual partners should be treated for TV.
Strength of recommendation: Grade 1, quality of evi-
dence: Grade A.
Follow-up
BV
Only in women with persistent symptoms. If treatment
is prescribed in pregnancy to reduce the risk of preterm
birth, a repeat test should be made after one month and
further treatment offered if BV has recurred.
AV
Women with persistent or recurrent symptoms.
Candida
Only in women with persistent or recurrent symptoms.
Consider other diagnoses, e.g. vulval dermatitis.
Trichomoniasis
Follow-up is unnecessary for men and women who
become asymptomatic after treatment or who are ini-
tially asymptomatic. Tests of cure are only recom-
mended if the patient remains symptomatic following
treatment or if symptoms recur.
Acknowledgements
The authors wish to acknowledge the following individuals
for their helpful comments:
Prof Harald Moi
Dr Raj Patel
Dr Keith Radcliffe
Prof Jonathan Ross
Dr Andy Winter
Composition of editorial board (see http://www.iusti.org/
regions/Europe/pdf/2017/EditorialBoardSept2017.pdf)
List of contributing organisations (see www.iusti.org/regions/
Europe/euroguidelines.htm)
Declaration of conflicting interests
The authors declared no potential conflicts of interest with
respect to the research, authorship, and/or publication of
this article.
Funding
The authors received no financial support for the research,
authorship, and/or publication of this article.
Proposed review date: 2023
ORCID iD
Jackie Sherrard http://orcid.org/0000-0002-1310-0372
Jørgen S Jensen
http://orcid.org/0000-0002-7464-7435
References
1. Koumans EH, Sternberg M, Bruce C, et al. The preva-
lence of bacterial vaginosis in the United States, 2001–
2004; associations with symptoms, sexual behaviors,
and reproductive health. Sex Transm Dis 2007;
34: 864–869.
2. Lamont RF, Morgan DJ, Wilden SD, et al. Prevalence
of bacterial vaginosis in women attending one of three
general practices for routine cervical cytology. Int J
STD AIDS 2000; 11: 495–498.
3. Marrazo JM, Koutsky LA, Eschenbach DA, et al.
Characterization of vaginal flora and bacterial vaginosis
in women who have sex with women. J Infect Dis 2002;
185: 1307–1313.
4. Fredricks DN, Fiedler TL and Marrazzo JM. Molecular
identification of bacteria associated with bacterial vagi-
nosis. N Engl J Med 2005; 353: 1899–1911.
5. Schwebke JR, Muzny CA and Josey WE. Role of
Gardnerella vaginalis in the pathogenesis of bacterial
vaginosis: a conceptual model. J Infect Dis 2014;
10: 338–343.
Sherrard et al. 1267
6. Swidsinski A, Mendling W, Loening-Baucke V, et al.
Adherent biofilms in bacterial vaginosis. Obstet
Gynecol 2005; 106: 1013–1023.
7. Donders GG. Definition of a type of abnormal vaginal
flora that is distinct from bacterial vaginosis: aerobic
vaginitis. BJOG 2002; 109: 1–10.
8. Mason MJ and Winter AJ. How to diagnose and treat
aerobic and desquamative inflammatory vaginitis. Sex
Transm Infect 2017; 93: 8–10.
9. Lindner JG, Plantema FH and Hoogkamp K.
Quantitative studies of the vaginal flora of healthy
women and of obstetric and gynaecological patients.
J Med Microbiol 1978; 11: 233–241.
10. Sobel JD. Pathogenesis and epidemiology of vulvova-
ginal candidosis. Ann N Y Acad Sci 1988;
544: 547–557.
11. Fidel PL Jr, Barousse M, Espinosa T, et al. An intra-
vaginal live Candida challenge in humans leads to new
hypotheses for the immunopathogenesis of vulvovaginal
candidiasis. Infect Immun 2004; 72: 2939–2946.
12. Holland J, Young ML, Lee O, et al. Vulvovaginal car-
riage of yeasts other than Candida albicans. Sex Transm
Infect 2003; 79: 249–250.
13. Wolner-Hanssen P, Kreiger JN, Stevens CE, et al.
Clinical manifestations of vaginal trichomoniasis.
JAMA 1989; 264: 571–576.
14. Fouts AC and Kraus SJ. Trichomonas vaginalis: re-
evaluation of its clinical presentation and laboratory
diagnosis. J Infect Dis 1980; 141: 137–143.
15. Atashili J, Poole C, Ndumbe PM, et al. Bacterial vagi-
nosis and HIV acquisition: a meta- analysis of published
studies. AIDS 2008; 22: 1493–1501.
16. Brotman RM, Klebanoff MA, Tonia R, et al. Bacterial
vaginosis assessed by gram stain and diminished coloni-
zation resistance to incident gonococcal, chlamydial,
and trichomonal genital infection. J Infect Dis 2010;
202: 1907–1915.
17. Rathod SD, Krupp K, Klausner JD, et al. Bacterial
vaginosis and risk for Trichomonas vaginalis infection:
a longitudinal analysis. Sex Transm Dis 2011;
38: 882–886.
18. Cherpes TL, Meyn LA, Krohn MA, et al. Association
between acquisition of herpes simplex virus type 2 in
women and bacterial vaginosis. Clin Infect Dis 2003;
37: 319–325.
19. Cohen CR, Lingappa JR, Baeten JM, et al. Bacterial
vaginosis associated with increased risk of female-to-
male HIV-1 transmission: a prospective cohort analysis
among African couples. PLoS Med 2012; 9: e1001251.
20. Balkus JE, Manhart LE, Lee J, et al. Periodic presump-
tive treatment for vaginal infections may reduce the
incidence of sexually transmitted bacterial infections.
J Infect Dis 2016; 213: 1932–1937.
21. Haggerty CL, Totten PA, Tang G, et al. Identification
of novel microbes associated with pelvic inflammatory
disease and infertility. Sex Transm Infect 2016;
92: 441–446.
22. Persson E, Bergstrom M, Larsson PG, et al. Infections
after hysterectomy. A prospective nation-wide Swedish
study. The Study Group on Infectious Diseases. Acta
Obstet Gynecol Scand 1996; 75: 757–761.
23. Soper DE, Bump RC and Hurt WG. Bacterial vaginosis
and trichomoniasis vaginitis are risk factors for cuff cel-
lulitis after abdominal hysterectomy.
Am J Obstet
Gynecol 1990; 163: 1016–1021.
24. Charonis G and Larsson PG. Use of pH/whiff test or
QuickVue advanced pH and Amines test for the diag-
nosis of bacterial vaginosis and prevention of postabor-
tion pelvic inflammatory disease. Acta Obstet Gynecol
Scand 2006; 85: 837–843.
25. Miller L, Thomas K, Hughes JP, et al. Randomised
treatment trial of bacterial vaginosis to prevent post-
abortion complication. BJOG 2004; 111: 982–988.
26. Lamont RF, Nhan-Chang CL, Sobel JD, et al.
Treatment of abnormal vaginal flora in early pregnancy
with clindamycin for the prevention of spontaneous pre-
term birth: a systematic review and metaanalysis. Am J
Obstet Gynecol 2011; 205: 177–190.
27. Leitich H, Bodner-Adler B, Brunbauer M, et al.
Bacterial vaginosis as a risk factor for preterm delivery:
a meta-analysis. Am J Obstet Gynecol 2003;
189: 139–147.
28. Brocklehurst P, Gordon A, Heatley E, Milan SJ.
Antibiotics for treating bacterial vaginosis in pregnancy.
Cochrane Database of Systematic Reviews 2013, Issue 1.
Art. No.: CD000262. DOI: 10.1002/14651858.
CD000262.pub4.
29. Sorvillo F and Kernott P. Trichomonas vaginalis and
amplification of HIV-1 transmission. Lancet 1998;
351: 213–214.
30. Laga M, Manoka A, Kivuvu M, et al. Non ulcerative
sexually transmitted diseases as risk factors for HIV-1
transmission in women: results from a cohort study.
AIDS 1993; 7: 95–102.
31. McClelland RS, Sangere L, Hassan WM et al. Infection
with Trichomonas vaginalis increases the risk of HIV-1
acquisition. J Infect Dis 2007; 195: 698–702.
32. Tanton C, Weiss HA, Le Goff J et al. Correlates of
HIV-1 genital shedding in Tanzanian women. PLoS
One 2011; 6: e17480.
33. Mavedzenge SN, Pol BV, Cheng H et al.
Epidemiological synergy of Trichomonas vaginalis and
HIV in Zimbabwean and South African women. Sex
Transm Dis 2010; 37: 460–466.
34. Cotch MF, Pastorek JG, Nugent RP, et al. Trichomonas
vaginalis associated with low birth weight and preterm
delivery. Sex Transm Dis 1997; 24: 353–360.
35. French JI, McGregor JA, Draper D, et al. Gestational
bleeding, bacterial vaginosis, and common reproductive
tract infections: risk for preterm birth and benefit of
treatment. Obstet Gynecol 1999; 93: 715–724.
36. Gu
¨
lmezoglu AM and Azhar M. Interventions for
trichomoniasis in pregnancy. Cochrane Database Syst
Rev 2011; 5: CD000220.
37. Donders GGG, Bellen G, Grinceviciene S, et al. The
limited value of symptoms and signs in the diagnosis
of vaginal infections. Arch Intern Med 1990;
150: 1929–1933.
1268 International Journal of STD & AIDS 29(13)
38. Donders GG, Van CK, Bellen G, et al. Predictive value
for preterm birth of abnormal vaginal flora, bacterial
vaginosis and aerobic vaginitis during the first trimester
of pregnancy. BJOG 2009; 116: 1315–1324.
39. Roberts CL, Algert CS, Rickard KL, et al. Treatment of
vaginal candidiasis for the prevention of preterm birth: a
systematic review and meta-analysis. Syst Rev 2015;
4: 31.
40. Holzer I, Farr A, Hagmann M, et al. The colonization
with Candida species is more harmful in the second tri-
mester of pregnancy. Arch Gynecol Obstet 2017;
295: 891–895.
41. Abbott J. Clinical and microscopic diagnosis of vaginal
yeast infection: a prospective analysis. Ann Emerg Med
1995; 25: 587–591.
42. Eckert LO, Hawes SE, Stevens CE, et al. Vulvovaginal
candidosis: clinical manifestations, risk factors, manage-
ment algorithm. Obstet Gynecol 1998; 92: 757–765.
43. Sonnex C and Lefort W. Microscopic features of vagi-
nal candidosis and their relation to symptomatology.
Sex Transm Infect 1999; 75: 417–419.
44. Mendling W, Brasc h J , C ornely OA, et al. G uideline:
vulvovaginal candidosis (AWMF 015/072), S2k
(excluding mucocutaneous candidosis). Mycoses
2015; 58: 1–15.
45. van de Wigert J, Altini L, Jones H, et al. Two methods
of self-sampling compared to clinician sampling to
detect reproductive tract infections in Gugulethu,
South Africa. Sex Transm Dis 2006; 33: 516–523.
46. Nugent RP, Krohn MA and Hillier SL. Reliability of
diagnosing bacterial vaginosis is improved by a stan-
dardized method of gram stain interpretation. J Clin
Microbiol 1991; 29: 297–301.
47. Gue
´
dou FA, Van Damme L, Mirembe F, et al.
Intermediate vaginal flora is associated with HIV
prevalence as strongly as bacterial vaginosis in a
cross-sectional study of participants screened for a rand-
omised controlled trial. Sex Transm Infect 2012;
88: 545–551.
48. Ison CA and Hay PE. Validation of a simplified grading
of gram stained vaginal smears for use in genitourinary
medicine clinics. Sex Transm Infect 2002; 78: 413–415.
49. Amsel R, Totten PA, Spiegel CA, et al. Nonspecific
vaginitis. Diagnostic criteria and microbial and epidemi-
ologic associations. Am J Med 1983; 74: 14–22.
50. Gallo MF, Jamieson DJ, Cu US, et al. Accuracy of
clinical diagnosis of bacterial vaginosis by human
immunodeficiency virus infection status. Sex Transm
Dis 2011; 38: 270–274.
51. Singh RH, Zenilman JM, Brown KM, et al. The role of
physical examination in diagnosing common causes of
vaginitis: a prospective study. Sex Transm Infect 2013;
89: 185–190.
52. Myziuk L, Romanowski B and Johnson SC. BVBlue
test for diagnosis of bacterial vaginosis. J Clin
Microbiol 2003; 41: 1925–1928.
53. Gaydos CA, Begaj S, Schwebke J et al. Clinical valida-
tion of a test for the diagnosis of vaginitis. Obstet
Gynecol 2017; 130: 181–189.
54. Donders GG, Vereecken A, Bosmans E, et al. Definition
of a type of abnormal vaginal flora that is distinct from
bacterial vaginosis: aerobic vaginitis. BJOG 2002;
109: 34–43.
55. Rumyantseva TA, Bellen G, Savochkina YA, et al.
Diagnosis of aerobic vaginitis by quantitative real-time
PCR. Arch Gynecol Obstet
2016; 294: 109–114.
56. Hopwood V, Crowley T, Horrocks CT, et al. Vaginal
candidosis: relation between yeast counts and symptoms
and clinical signs in non-pregnant women. Genitourin
Med 1988; 64: 331–334.
57. Odds FC, Webster CE, Mayuranathan P, et al. Candida
concentrations in the vagina and their association with
signs and symptoms of vaginal candidosis. J Med Mycol
1988; 26: 277–283.
58. Priestley CJ, Jones BM, Dhar J, et al. What is normal
vaginal flora? Genitourin Med 1997; 73: 23–28.
59. Kingston MA, Bansal D and Carlin EM. ‘Shelf life’ of
Trichomonas vaginalis. Int J STD AIDS 2003; 14: 28–29.
60. Bickley LS, Krisher KK, Punsalang A, et al.
Comparison of direct fluorescent antibody, acridine
orange, wet mount and culture for detection of
Trichomonas vaginalis in women attending a public
sexually transmitted disease clinic. Sex Transm Dis
1989; 127–131.
61. Kreiger JN, Tam MR, Stevens CE, et al. Diagnosis of
trichomoniasis: comparison of conventional wet- mount
examination with cytological studies, cultures, and
monoclonal antibody staining of direct specimens.
JAMA 1988; 259: 1223–1227.
62. Nye MB, Schwebke JR and Body BA. Comparison of
APTIMA Trichomonas vaginalis transcription-mediated
amplification to wet mount microscopy, culture, and
polymerase chain reaction for diagnosis of trichomoni-
asis in men and women. Am J Obstet Gynecol 2009; 200:
188.e181–188.e187.
63. Hegazy MM, El-Tantawy NL, Soliman MM, et al.
Performance of rapid immunochromatographic assay
in the diagnosis of Trichomoniasis vaginalis. Diagn
Microbiol Infect Dis 2012; 74: 49–53.
64. Campbell L, Woods V, Lloyd T, et al. Evaluation of the
OSOM Trichomonas rapid test versus wet preparation
examination for detection of Trichomonas vaginalis vag-
initis in specimens from women with a low prevalence of
infection. J Clin Microbiol 2008; 46: 3467–3469.
65. Borchardt KA, et al. A comparison of the sensitivity of
the InPouch TV, diamond’s and trichosel media for
detection of Trichomonas vaginalis. Genitourin Med
1997; 73: 297–298.
66. el Naga IF, Khalifa AM and el Azzouni MZ. In-pouch
TV culture system in diagnosis of Trichomonas vaginalis
infection. J Egypt Soc Parasitol 2001; 31: 647–656.
67. Levi MH, Torres J, Pina C, Klein RS. Comparison of
the InPouch TV culture system and diamond’s modified
medium for detection of Trichomonas vaginalis. J Clin
Microbiol 1997; 35: 3308–3310.
68. Van Der Shee C, van Belkum A, Zwiggers L et al.
Improved diagnosis of Trichomonas vaginalis infection
by PCR using vaginal swabs and urine specimens
Sherrard et al. 1269
compared to diagnosis by wet mount, culture and fluo-
rescent staining. J Clin Microbiol 1999; 37: 4127–4130.
69. Radonjic IV, Dzamic AM, Mitrovic SM, et al.
Diagnosis of Trichomonas vaginalis: the sensitivities
and specificities of microscopy, culture and PCR
assay. Eur J Obstet Gynecol Reprod Biol 2006;
126: 116–120.
70. Hardick A, Hardwick J, Wood BJ, Gaydos C.
Comparison between the Gen-Probe transcription-
mediated amplification Trichomonas vaginalis research
assay and real-time PCR for Trichomonas vaginalis
detection using a Roche LightCycler instrument with
female self-obtained vaginal swab samples and male
urine samples. J Clin Microbiol 2006; 44: 4197–4199.
71. Munson E, Napierala M, Olson R et al. Impact of
Trichomonas vaginalis transcription-mediated amplifica-
tion-based analyte-specific reagent testing in a metropol-
itan setting of high sexually transmitted disease
prevalence. J Clin Microbiol 2008; 46: 3368–3374.
72. Schwebke JR, Hobbs MM, Taylor SN, et al. Molecular
testing for Trichomonas vaginalis in women: results from
a prospective U.S. clinical trial. J Clin Microbiol 2011;
49: 4106–4111.
73. Ginocchio CC, Chapin K, Smith JS, et al. Prevalence of
Trichomonas vaginalis and coinfection with Chlamydia
trachomatis and Neisseria gonorrhoeae in the United
States as determined by the Aptima Trichomonas vagi-
nalis nucleic acid amplification assay. J Clin Microbiol
2012; 50: 2601–2608.
74. Larsson P-G. Treatment of bacterial vaginosis. Int J
STD AIDS 1992; 3: 239–247.
75. Joesoef MR and Schmid GP. Bacterial vaginosis: review
of treatment options and potential clinical indications
for therapy. Clin Infect Dis 1995; 20: S72–S79.
76. Schwebke JR and Desmond RA. A randomized trail of
the duration of therapy with metronidazole plus or
minus azithromycin for the treatment of symptomatic
bacterial vaginosis. Clin Infect Dis 2007; 44: 213–219.
77. Bradshaw CS, Pirotta M, De GD, et al. Efficacy of oral
metronidazole with vaginal clindamycin or vaginal pro-
biotic for bacterial vaginosis: randomised placebo-
controlled double-blind trial. PLoS One 2012; 7: e34540.
78. Weissenbacher ER, Donders G, Unzeitig V, et al.
A comparison of dequalinium chloride vaginal tablets
(Fluomizin) and clindamycin vaginal cream in the treat-
ment of bacterial vaginosis: a single-blind, randomized
clinical trial of efficacy and safety. Gynecol Obstet Invest
2012; 73: 8–15.
79. Bradshaw CS, Morton AN, Hocking J, et al. High
recurrence rates of bacterial vaginosis over the course
of 12 months after oral metronidazole therapy and fac-
tors associated with recurrence. J Infect Dis 2006;
193: 1478–1486.
80. Brotman RM, Klebanoff MA, Nansel TR et al. A lon-
gitudinal study of vaginal douching and bacterial vagi-
nosis a marginal structural modeling analysis. Am J
Epidemiol 2008; 168: 188–196.
81. Bradshaw CS, Vodstrcil LA, Hocking JS et al.
Recurrence of bacterial vaginosis is significantly
associated with posttreatment sexual activities and hor-
monal contraceptive use. Clin Infect Dis 2013;
56: 777–786.
82. Sobel JD, Ferris D, Schwebke J, et al. Suppressive anti-
bacterial therapy with 0.75% metronidazole vaginal gel
to prevent recurrent bacterial vaginosis. Am J Obstet
Gynecol 2006; 194: 1283–1289.
83. McClelland RS, Balkus JE, Lee J, et al. Randomised
trial of periodic presumptive treatment with high dose
intravaginal metronidazole and miconazole to prevent
vaginal infections in HIV-negative women. J Infect Dis
2015; 211: 1875–1882.
84. McClelland RS, Richardson BA, Hassan WM, et al.
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: 1361–1368.
85. Andersch B, Lindell D, Dahlen I, et al. Bacterial vagi-
nosis and the effect of intermittent prophylactic treat-
ment with an acid lactate gel. Gynecol Invest 1990;
30: 114–119.
86. Senok AC, Verstraelen H, Temmerman M, et al.
Probiotics for the treatment of bacterial vaginosis.
Cochrane Database Syst Rev 2009; 4: CD006289.
87. Hanson L, VandeVusse L, Jerme M, et al. Probiotics for
treatment and prevention of urogenital infections in
women: a systematic review. J Midwifery Womens
Health 2016; 61: 339–355.
88. Sobel JD, Reichman J, Misra D, et al. Prognosis and
treatment of desquamative inflammatory vaginitis.
Obstet Gynecol 2011; 117: 850–855.
89. Watson MC, Grimshaw JM, Bond CM, et al. Oral
versus intra-vaginal imidazole and triazole anti-fungal
agents for the treatment of uncomplicated vulvovaginal
candidiasis (thrush): a systematic review. BJOG 2002;
109: 85–95.
90. Nurbhai M, Grimshaw J, Watson M, et al. Oral versus
intra-vaginal imidazole and triazole anti-fungal treatment
of uncomplicated vulvovaginal candidiasis (thrush).
Cochrane Database Syst Rev 2007; 4: CD002845.
91. Sobel JD, Kapernick PS, Zervos M, et al. Treatment of
complicated Candida vaginitis: comparison of single and
sequential doses of fluconazole. Am J Obstet Gynecol
2001; 185: 363–369.
92. Mendling W and Schlegelmilch R. Three-day combina-
tion treatment for vulvovaginal candidosis with 200 mg
clotrimazole vaginal suppositories and clotrimazole
cream for the vulva is significantly better than treatment
with vaginal suppositories alone an earlier, multi-
centre, Placebo-Controlled Double Blind Study.
Geburtsh Frauenheilk 2014; 74: 355–360.
93. Sobel J, Wiesenfeld H, Martens M, et al. Maintenance
fluconazole therapy for recurrent vulvovaginal candidi-
asis. N Engl J Med 2004; 351: 876–883.
94. Cooke G, Watson C, Smith J, et al. Treatment for recur-
rent vulvovaginal candidiasis (thrush). Cochrane
Database Syst Rev 2011; 5: CD009151.
95. Donders G, Bellen G, Byttebier G et al. Individualized
decreasing-dose maintenance fluconazole regimen for
1270 International Journal of STD & AIDS 29(13)
recurrent vulvovaginal candidiasis (ReCiDiF trial). Am
J Obstet Gynecol 2008; 199: 613–619.
96. Rosa MI, Silva BR, Pires PS et al. Weekly fluconazole
therapy for recurrent vulvovaginal candidiasis: a sys-
tematic review and meta-analysis. Eur J Obstet
Gynecol Reprod Biol 2013; 167: 132–136.
97. Nyirjesy P, ZhaoDavies S, Johnson E, et al. How to
treat persistent vaginal yeast infection due to species
other than Candida albicans. Sex Transm Infect 2013;
89: 165–166.
98. Dennerstein GJ. Depo-Provera in the treatment of
recurrent vulvovaginal candidiasis. J Reprod Med
1986; 31: 801–803.
99. Forna F and Gu
¨
lmezoglu AM. Interventions for treat-
ing trichomoniasis in women. Cochrane Database Syst
Rev 2003; 2: CD000218.
100. Thin RN, Symonds MAE, Booker R, et al. Double-
blind comparison of a single dose and a five-day
course of metronidazole in the treatment of trichomoni-
asis. Br J Vener Dis 1979; 55: 354–356.
101. Howe K and Kissinger PJ. Single-dose compared with
multidose metronidazole for the treatment of trichomo-
niasis in women: a meta-analysis. Sex Transm Dis 2017;
44: 30–35.
102. Pearlman MD, Yashar C, Ernst S, et al. An incremental
dosing protocol for women with severe vaginal tricho-
moniasis and adverse reactions to metronidazole. Am J
Obstet Gynecol 1996; 174: 934–936.
103. Kurohara ML, Kwong FK, Lebherz TB, et al.
Metronidazole hypersensitivity and oral desensitization.
J Allergy Clin Immunol 1991; 88: 279–280.
104. Das S, Huengsberg M and Shahmanesh M. Treatment
failure of vaginal trichomoniasis in clinical practice. Int
J STD AIDS 2005; 16: 284–286.
105. Bosserman EA, Helms DJ, Mosure DJ, et al. Utility of
antimicrobial susceptibility testing in Trichomonas vagi-
nalis-infected women with clinical treatment failure. Sex
Transm Dis 2011; 38: 983–987.
106. Sobel JD, Nyirjesy P and Brown W. Tinidazole therapy
for metronidazole-resistant vaginal trichomoniasis. Clin
Infect Dis 2001; 33: 1341–1346.
107. Mammen-Tobin A and Wilson JD. Management of
metronidazole-resistant Trichomonas vaginalis a new
approach. Int J STD AIDS 2005; 16: 488–490.
108. Muanda FT, Sheehy O and Be
´
rard A. Use of antibiotics
during pregnancy and risk of spontaneous abortion.
CMAJ 2017; 189: E625–E633.
109. Gu
¨
lmezoglu AM and Azhar M. Interventions for
trichomoniasis in pregnancy. Cochrane Database Syst
Rev 2011; 5: CD000220.
110. Czeizel AE and Rockenbauer M. A population based
case-control teratologic study of oral metronidazole
treatment during pregnancy. BJOG 1998; 105: 322–327.
111. Burtin P, Taddio A, Ariburnu O, et al. Safety of met-
ronidazole in pregnancy: a meta-analysis. Am J Obstet
Gynecol 1995; 172: 525–529.
112. Caro-Paton T, Carvajal A, de Diego IM, et al. Is met-
ronidazole teratogenic: a meta-analysis. Br J Clin
Pharmacol 1997; 44: 179–182.
113. Molgaard-Nielsen D, Pasternak B and Hviid A. Use of
fluconazole during pregnancy and risk of birth defects.
N Engl J Med 2013; 369: 830–839.
114. Mølgard-Nielsen D, Svanstr
om H, Melbye M, et al.
Association between use of oral fluconazole during
pregnancy and risk of spontaneous abortion and still-
birth. JAMA 2016; 315: 58–67.
115. Amaya-Guio J, Viveros-Carre
~
no DA, Sierra-Barrios
EM, et al. Antibiotic treatment for the sexual partners
of women with bacterial vaginosis. Cochrane Database
Syst Rev 2016; 10: CD011701.
116. Bisschop MP, Merkus JM, Scheygrond H, et al. Co-treat-
ment of the male partner in vaginal candidosis: a double-
blind randomized control study. BJOG 1986; 93: 79–81.
117. Fong IW. The value of treating the sexual partners of
women with recurrent vaginal candidiasis with ketoco-
nazole. Genitourin Med 1992; 68: 174–176.
118. Lyng J and Christensen J. A double blind study of treat-
ment with a single dose tinidazole of partners to females
with trichomoniasis. Acta Obstet Gynecol Scand 1981;
60: 199–201.
119. Schwebke JR and Desmond RA. A randomized con-
trolled trial of partner notification methods for preven-
tion of trichomoniasis in women. Sex Transm Dis 2010;
37: 392–396.
Appendix 1
Review of the literature
An extensive literature review was performed using
Medline for the years 2009–2017. MEDLINE search-
keywords: vulvovaginal candidosis, vaginal candidosis,
vaginal candida, Trichomonas vaginalis, trichomonia-
sis, Bacterial vaginosis, non-specific vaginitis, abnor-
mal vaginal flora, vaginal dysbiosis. The resulting
articles were handsearched and sorted. Further refer-
ences were obtained from these articles.
The Cochrane Library was searched; search-
keywords were:
vulvovaginal candidosis, vaginal candidosis, vaginal
candida, Trichomonas vaginalis, in women, bacteri-
al vaginosis.
The 2015 US CDC guidelines for the treatment of
Sexually Transmitted Diseases and the related UK
national guidelines (www.bashh.org) were reviewed.
Tables of levels of evidence and grading of
recommendations:
(see http://www.iusti.org/regions/Europe/pdf/2017/
ProtocolForProduction2017.pdf).
Appendix 2. Declarations of interests
Jackie Sherrard: JS has received consultancy fees from
Becton Dickinson
Janet Wilson: JW has received speaker fees from BD
Diagnostics; unconditional research grants in the form
Sherrard et al. 1271
of diagnostic tests from Hologic; and remuneration for
contract research from Starpharma.
Gilbert Donders: received consultancy fees and /or
speakers fees from Medinova, Alfa-Wasserma n,
Bayer, Phacobel and GSK.
Werner Mendling: WM declares in the last three
years royalties by Aristo Pharma GmbH Berlin, Dr
August Wolff GmbH & Co. KG Arzneimittel
Bielefeld, Dr Kade Pharmazeutische Fabrik GmbH
Berlin, Pierre Fabre Pharma GmbH Freiburg,
SymbioPharm Herborn and Johnson & Johnson
GmbH Neuss (all Germany), and Medinova AG
Zurich/Switzerland
Jørgen S Jensen: JSJ has received speaker fees from
Hologic and SSI has received remuneration for con-
tract research from Hologic, SpeeDx, NYtor,
Diagenode, Nabriva, Angelini, GSK, and Osel
1272 International Journal of STD & AIDS 29(13)