Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/334404210

Gestational diabetes mellitus

Article in Nature Reviews Disease Primers · July 2019


DOI: 10.1038/s41572-019-0098-8

CITATIONS READS

914 18,272

6 authors, including:

Harold David Mcintyre Patrick Catalano


The University of Queensland Tufts Medical Center
358 PUBLICATIONS 24,454 CITATIONS 60 PUBLICATIONS 1,753 CITATIONS

SEE PROFILE SEE PROFILE

Cuilin Zhang Gernot Desoye


Harvard University Medical University of Graz
408 PUBLICATIONS 16,885 CITATIONS 543 PUBLICATIONS 13,529 CITATIONS

SEE PROFILE SEE PROFILE

All content following this page was uploaded by Gernot Desoye on 11 March 2020.

The user has requested enhancement of the downloaded file.


PRIMER
­­Gestational diabetes mellitus
H. David McIntyre1*, Patrick Catalano2, Cuilin Zhang3, Gernot Desoye4,
Elisabeth R. Mathiesen5 and Peter Damm6
Abstract | Hyperglycaemia that develops during pregnancy and resolves after birth has been
recognized for over 50 years, but uniform worldwide consensus is lacking about threshold
hyperglycaemic levels that merit a diagnosis of ‘gestational diabetes mellitus’ (GDM) and thus
treatment during pregnancy. GDM is currently the most common medical complication of
pregnancy, and prevalence of undiagnosed hyperglycaemia and even overt diabetes in young
women is increasing. Maternal overweight and obesity, later age at childbearing, previous
history of GDM, family history of type 2 diabetes mellitus and ethnicity are major GDM risk
factors. Diagnosis is usually performed using an oral glucose tolerance test (OGTT), although a
non-​fasting, glucose challenge test (GCT) is used in some parts of the world to screen women for
those requiring a full OGTT. Dietary modification and increased physical activity are the primary
treatments for GDM, but pharmacotherapy, usually insulin, is used when normoglycaemia is
not achieved. Oral hypoglycaemic agents, principally metformin and glibenclamide (glyburide),
are also used in some countries. Treatment improves immediate pregnancy outcomes, reducing
excess fetal growth and adiposity and pregnancy-​related hypertensive disorders. GDM increases
the risk of long-​term c­om­pl­ic­at­ions, including obesity, impaired glucose metabolism and c­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­a­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­r­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­d­­­­
­­­­­­­­­­­­­i­­­­­­­­­­­­­­­­­­­o­­­­­­v­­­­­as­­cular d­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­i­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­­s­­­­­­­­­­­­­­e­­a­­se, i­­­­n b­­­­o­­t­h t­­h­e m­­­­o­­t­­­her a­nd infant. Optimal management of mother and infant
during long-​term follow-​up remains challenging, with very limited implementation of preventive
strategies in most parts of the world.

Although the term ‘gestational diabetes’ was coined by societal trend for childbearing at a later age, especially
Carrington in 1957 (ref.1), it only gained wider recog- in regions in which obesity and early-​onset diabetes are
nition after the publications of John O’Sullivan in 1961 also common. For example, in the USA, the reported
(ref.2) and 1964 (ref.3). However, the phenomenon of prevalence of overt diabetes in adults 20–22 years of age
hyperglycaemia that develops during pregnancy (gen- is 4.5%, and a further 29.3% of these individuals have
erally detected in the late second trimester (13–26 com- detectable pre-​diabetes (IFG, IGT or impaired glycated
pleted weeks of gestation) or early in the third trimester haemoglobin (HbA1c))6. In the context of this epidemic
(27–40 weeks)) and resolves following delivery, was of hyperglycaemia outside pregnancy, it is highly likely
noted some time before4. that many cases diagnosed as GDM actually repre-
Complexity and controversy have shadowed the sent undiagnosed pre-​pregnancy hyperglycaemia of
diagnosis of gestational diabetes mellitus (GDM) ever varying severity.
since, partly owing to the very broad definition that was Globally, no single diagnostic protocol or set of diag­
initially promoted5, which allowed for “any degree of nostic criteria for GDM has gained universal accept-
hyperglycaemia first recognized in pregnancy”, irrespec- ance, making international comparisons difficult.
tive of the treatment required or the extent of postpar- However, the International Association of Diabetes
tum resolution of hyperglycaemia. Consequently, ‘GDM’ in Pregnancy Study Groups (IADPSG) 2010 criteria7,
encompassed a broad range of hyperglycaemia, from which were endorsed by the WHO in 2013 (ref.8), clas-
mild impaired glucose tolerance (IGT) or impaired fast- sify women who are first diagnosed during pregnancy
ing glucose (IFG) detected in late pregnancy to glucose but who would be diagnosed with diabetes if hyper-
levels characteristic of overt diabetes (or, rarely, even glycaemia were detected outside of pregnancy as
new-​onset type 1 diabetes mellitus (T1DM)) detected distinct from women who have ‘standard’ GDM.
in early pregnancy (<20 weeks of gestation). Although IADPSG described these women as having ‘overt dia-
*e-​mail: David.McIntyre@
mater.org.au more severe hyperglycaemia was initially uncommon, it betes’ whereas the WHO preferred the term ‘diabetes
https://doi.org/10.1038/ has become much more prominent with the expanding in pregnancy’. However, both groups recognized that
s41572-019-0098-8 worldwide epidemics of diabetes and obesity and the these women had severe hyperglycaemia that merited

NATURE REvIEWS | DiSeASe PRiMeRS | Article citation I­D:­ (2019) 5:47 ­ 1

0123456789();
Primer

Author addresses review by including studies published between August


2015 and December 2018. The overall updated GDM
1
Mater Research and University of Queensland, Brisbane, Queensland, Australia. prevalence map by WHO regions and country-​specific
2
Mother Infant Research Institute, Department of Obstetrics and Gynecology, Tufts estimates of GDM prevalence are illustrated in Figs 1
University School of Medicine, Friedman School of Nutrition Science and Policy, Boston,
and 2. The prevalence of GDM is highest in the Middle
MA, USA.
East and some North African countries, with a median
3
Division of Intramural Population Health Research, Eunice Kennedy Shriver National
Institute of Child Health and Human Development, NIH, Bethesda, MD, USA. of 15.2% (interquartile range 8.8–20.0%), followed by
4
Department of Obstetrics and Gynecology, Medical University of Graz, Graz, Austria. South-​East Asia (median 15.0%; range 9.6–18.3%), the
5
Department of Endocrinology, Center for Pregnant Women with Diabetes, Western Pacific (median 10.3%; range 4.5–20.3%), South
Rigshospitalet and The Institute of Clinical Medicine, Faculty of Health and Medical and Central America (median 11.2%; range 7.1–16.6%),
Sciences, University of Copenhagen, Copenhagen, Denmark. sub-​Saharan Africa (median 10.8%; range 8.5–13.1%)
6
Department of Obstetrics, Center for Pregnant Women with Diabetes, Rigshospitalet and North America and the Caribbean (median 7.0%;
and The Institute of Clinical Medicine, Faculty of Health and Medical Sciences, University range 6.5–11.9%). The lowest GDM prevalence and
of Copenhagen, Copenhagen, Denmark. widest variation in prevalence is observed in Europe
(median 6.1%; range 1.8–31.0%). Even within each
immediate treatment and might also have established region, considerable variation is observed, both within
microvascular complications of diabetes that require and between countries. For example, in the Western
timely intervention9. Thus, these women are at high risk Pacific region, the prevalence ranges from 4.5% in Japan to
and are clinically distinct from women with less marked 18.0% in Singapore. By contrast, among countries in
hyperglycaemia. However, in this Primer, we consider North America, GDM prevalence is relatively consistent.
primarily those women who are classified as having As few studies are available to estimate GDM prevalence
GDM using the IADPSG or WHO 2013 criteria (that is, in Africa and South and Central America, more studies
hyperglycaemia that is first diagnosed during pregnancy, are clearly warranted in these regions.
with glucose levels below those considered diagnostic of
overt diabetes outside of pregnancy) (Table 1). Risk factors
Irrespective of diagnostic criteria or temporal trends, Epidemiological studies have identified a number of GDM
GDM is important for two reasons. First, GDM increases risk factors (Box 1), such as advanced maternal age, ethni­
the risk of complications during pregnancy for both the city, previous history of gestational diabetes and family
mother (especially hypertensive disorders of pregnancy) history of type 2 diabetes mellitus (T2DM). Although the
and the fetus (especially those related to excessive fetal traditional focus has been on risk factors detected during
growth and adiposity). Second, a GDM diagnosis pregnancy, data support the important role of risk factors
identifies a group of women and their offspring who during the periconception and preconception periods in
are at higher risk of diabetes, obesity and premature the development of GDM14,15.
cardiovascular disease in the long term10–12.
In this Primer, we explore the clinical definition of Age. Advanced maternal age has been related to increased
gestational diabetes and examine its underlying mol­ risk of GDM. In a large prospective study in the USA
ecular and clinical pathophysiology. We consider the (>95% white ethnicity), women >40 years of age had a
epidemiology and relationship of gestational diabetes to more than twofold increased risk of GDM compared
immediate adverse perinatal outcomes, including mater- with women <30 years of age (prevalence 9.8% versus
nal quality of life (QOL). We also discuss the treatment 4.1%, respectively), even after adjustment for other
of gestational diabetes during pregnancy and evolv- major risk factors16. Women carrying a male fetus
ing trends in risk factors, prevention and longer-​term seem to have a higher risk of developing GDM17, and
follow-​up of affected mothers and offspring. Finally, some reports18 suggest a higher risk of GDM in twin
we outline future perspectives for treating and under- pregnancies, although this is not a universal finding19.
standing gestational diabetes, including areas requiring
further research across the translational spectrum. Geography and ethnicity. It should be noted that even
when the same diagnostic criteria were applied, consid­
Epidemiology erable variability in prevalence estimates of GDM was
Prevalence observed between different countries (Fig. 2), which
The documented prevalence of GDM varies substan- indi­cates that variations in the distributions of inherent
tially worldwide, ranging from 1% to >30%. Owing to a charac­teristics of study populations may contribute to the
lack of consensus and uniformity in the screening stan­ variability. Furthermore, in countries with multi-​ethnic
dards and diagnostic criteria for GDM, it is challenging populations (such as Australia, the USA and Canada),
to compare the prevalence across countries and regions. notable differences in the prevalence of GDM between
Furthermore, historical definitions of GDM make it dif- ethnicities have been observed. For example, in northern
ficult to differentiate between undiagnosed diabetes and California, GDM prevalence was highest among women
GDM, and GDM diagnostic criteria have changed over from the Philippines (10.9%) and Asians (10.2%) and
the years. To capture the contemporary burden of GDM lowest among non-​Hispanic white (4.5%) and African-​
with the consideration of variations due to these factors, American (4.4%) women20. In Australia21, women of
a review of the global prevalence of GDM was carried South Asian origin had more than fourfold higher risk
out on the basis of studies between 2005 and 2015 of GDM than women of Australian or New Zealand
(ref.13). Using the same methods, we have expanded this origin, which is consistent with the higher prevalence

2 | Article citation ID: (2019) 5:47 www.nature.com/nrdp

0123456789();
Primer

Table 1 | Glucose threshold values in the OGTT for a GDM diagnosis


Criteria or trial Sample Fasting plasma 1-hour glucose 2-hour glucose 3-hour glucose
size (n) glucose (in mmol l–1 (in mmol l–1 (in mmol l–1 (in mmol l–1
(mg dl–1)) (mg dl–1)) (mg dl–1)) (mg dl–1))
IADPSG 2010 recommended glucose thresholds
IADPSG OR 1.5 23,316 ≥5.0 (90) ≥9.3 (167) ≥7.9 (142) NA
IADPSG OR 1.75a 23,316 ≥5.1 (92) ≥10.0 (180) ≥8.5 (153) NA
IADPSG OR 2.0 23,316 ≥5.3 (95) ≥10.6 (191) ≥9.0 (162) NA
Glucose thresholds for a NA ≥7.0 (126) NA ≥11.1 (200) NA
diagnosis of overt diabetes
(IADPSG) or diabetes in
pregnancy (WHO)b
OGTT thresholds used for exclusion or inclusion in randomized trials
Crowther et al. (2005)109 1,000 ≥7.0c (126) NA ≥7.8d (140) NA
Landon et al. (2009) 110e
958 ≥5.3 (95)
c
≥10.0 (180)
d
≥8.6d (155) ≥7.8 (140)
Carpenter–Coustan NA ≥5.3 (95) ≥10.0 (180) ≥8.6 (155) ≥7.8 (140)
criteriaf (ACOG primary
recommendation)177
NDDG criteriaf (ACOG NA ≥5.8 (105) ≥10.6 (190) ≥9.2 (165) ≥8.0 (145)
alternative)
ACOG, American College of Obstetricians and Gynecologists; GDM, gestational diabetes mellitus; IADPSG, International
Association of Diabetes and Pregnancy Study Groups; NA , not applicable; NDDG, National Diabetes Data Group; OGTT, oral
glucose tolerance test. aThresholds are glucose values at which the odds of birthweight, neonate percent body fat and cord blood
C-​peptide level greater than the ninetieth percentile are 1.75 times the estimated odds of these outcomes at average glucose
values. The alternative diagnostic thresholds corresponding to OR 1.5 and 2.0 compared to the mean are also shown. At least one
elevated value is sufficient for a GDM diagnosis. bThresholds used to define overt diabetes or diabetes in pregnancy are the same
as those used outside of pregnancy. cValue for exclusion from trials. dValue for inclusion in trials. eTwo or more elevated values at 1,
2 or 3 hours after glucose ingestion were required for exclusion or inclusion in the Landon trial. fTwo or more elevated values are
required for a GDM diagnosis.

of GDM among the general South-​East Asian popula- include greater consumption of fibre29,37 and nuts35, a
tion. The reasons underlying the ethnic differences are prudent dietary pattern characterized by a high intake
likely to be multifactorial, including but not limited to of fruit, green leafy vegetables, poultry and fish25, and a
the major risk factors for GDM (differences in body adi- ‘Mediterranean’ diet38.
posity, lifestyle (diet and physical activity) and genetic Overall, findings from observational studies suggest
susceptibility). that approximately 45% of GDM cases might be prevent-
able by the adoption of a healthy diet before pregnancy,
Modifiable lifestyle factors. Being overweight or obese maintaining a BMI <25 kg m–2, exercising for ≥30 min-
before pregnancy (body mass index (BMI) ≥25 kg m–2) utes per day and avoiding cigarette smoking39. However,
is the most significant GDM risk factor22. Cigarette owing to the lack of data on these risk factors both before
smoking by pregnant women and whether their parents pregnancy and during pregnancy before the time of a
smoked are related to increased risk of GDM, inde- GDM diagnosis, it is unclear whether these behaviours
pendent of pre-​pregnancy BMI and other risk factors23. have a chronic effect on insulin sensitivity and pancreatic
Physical activity both during and before pregnancy β-​cell function or an acute effect due to continuing these
reportedly reduces GDM risk24,25. In addition, a num- behaviours during pregnancy. If the effects of these risk
ber of dietary factors affect GDM risk. To date, no con- factors are chronic rather than acute, then prevention
crete conclusions can be drawn about the role of specific of GDM would require starting lifestyle interventions
dietary factors during pregnancy in the development before pregnancy. Indeed, preventive efforts may need
of GDM22. However, there is suggestive evidence that to be made at the population level to prevent childhood
low plasma levels of vitamin D26 and vitamin C27 in early overweight and obesity, reduce excess adolescent weight
pregnancy and increased dietary fat intake28,29 dur- gain and promote an optimal lifestyle pattern in both
ing pregnancy increase the risk of developing GDM22. parents before pregnancy40.
Large observational studies on pre-​pregnancy diet iden-
tified various dietary factors that potentially increase Emerging risk factors. In addition to diet and lifestyle
GDM risk, independent of body adiposity and physi- factors, emerging data indicate a possible contribution
cal acti­vity, including higher consumption of sugar-​ of environmental and psychosocial factors to the risk of
sweetened beverages30, potatoes31, fried foods32, haem developing GDM. For example, higher exposure to
iron33 and animal fat34 and protein35. Furthermore, a diet persistent organic pollutants and endocrine disruptors,
low in carbohydrates but high in animal fat and protein36, such as polybrominated diphenyl ethers41 and perfluoro­
as well as an overall ‘western’ dietary pattern (high intake octanoic acid42, has been associated with increased GDM
of red meat, processed meat, refined grain products, risk. Furthermore, depression in the first and second
sweets, french fries and pizza25) are associated with trimester has been prospectively related to increased
increased GDM risk. Potential healthful dietary factors GDM risk43.

NATURE REvIEWS | DiSeASe PRiMeRS | Article citation I­D:­ (2019) 5:47 ­ 3

0123456789();
Primer

Europe
6.1 (1.8–31.0)

North America Western


and Caribbean Pacific
7.0 (6.5–11.9) 10.3 (4.5–20.3)

South and South-East


Central America Asia
11.2 (7.1–16.6) 15.0 (9.6–18.3)

Middle East and Sub-Saharan


North Africa Africa
15.2 (8.8–20.0) 10.8 (8.5–13.1)

Fig. 1 | Global prevalence of GDM in 2005–2018. Median (interquartile range) prevalence (%) of gestational diabetes
mellitus (GDM) by WHO region, 2005–2018 (map generated from WHO website). A literature search was conducted in
PubMed supplemented by cross-​checking relevant references of eligible studies on the prevalence of GDM from 1 January
2005 to 1 December 2018 to capture the contemporary burden of GDM. Among the eligible studies that met the search
criteria13, data from countries reported in the studies were included to derive country-​specific estimates for GDM
prevalence. The region-​specific prevalence of GDM was estimated by calculating the median prevalence of country-specific
estimates within each WHO region.

Genetic factors. Although genetic heritability is impli- pregnancy. The metabolic adaptations during preg-
cated in the aetiology of GDM, studies that have nancy place additional stress on β-​cells (Figs 3,4). The
examined associations of specific genetic factors with increased risk of T2DM in the years after pregnancy in
GDM are limited and findings are inconsistent. A sys- women with a history of GDM is related both to pre-​
tematic review and meta-​analysis suggested that the existing (often undiagnosed) baseline abnormalities
minor alleles of nine single-​nucleotide polymorphisms and to further, progressive β-​cell dysfunction after the
(SNPs) in seven genes (such as rs7903146 (in TCF7L2), index GDM pregnancy, which are associated with factors
rs12255372 (in TCF7L2), rs1799884 (−30G/A, in GCK) such as retention of excessive gestational weight gain and
and rs5219 (E23K, in KCNJ11)), most of which are increases in insulin resistance. Only a small proportion
involved in regulating insulin secretion, are associated of women with GDM (2–13%) have antibodies against
with increased risk of GDM, which supports an impor- specific β-​cell antigens48, whereas approximately 5% of
tant role of pancreatic islet β-​cell compensation in the women with GDM have monogenic variants of dia­betes
pathogenesis of GDM44. In the only genome-​wide associ- mellitus, which most commonly involve mutations in
ation study of GDM in an Asian population, two genetic GCK (encoding glucokinase) in white populations49.
variants, rs10830962 (near MTNR1B) and rs7754840 As glucokinase phosphorylates glucose to produce
(in CDKAL1) were associated with GDM45. Subsequently, glucose-6-phosphate in the pancreas and liver, a hetero­
rs10830962 was associated with β-​cell compensation for zygous GCK mutation in the mother often results in
insulin resistance in Hispanic women with prior GDM46. mildly elevated fasting glucose levels with the attendant
In a study of 112 SNPs among 8,722 white women (2,636 risk of excess fetal growth if the fetus does not have the
with GDM and 6,086 without GDM) from two inde- GCK mutation. Interestingly, if the GCK mutation is pres-
pendent populations47, 8 novel GDM-​associated SNPs ent in both the mother and the fetus, then fetal growth
were identified. Larger genetic studies of GDM that is normal, whereas if only the fetus has the GCK muta-
included information about the fetal and/or paternal tion, there is an increased risk of fetal growth restriction
genome, gene–gene and gene–environmental inter- because of altered glucose sensing by the fetal pancreas50.
actions and in non-​white populations are rare. Future
studies with larger sample sizes, likely through consor- Metabolic changes during normal pregnancy
tium efforts, are warranted to improve understanding of To understand the pathophysiology of GDM, one needs
genetic contributions to the aetiology of GDM. to recognize the metabolic alterations that occur in a
normal pregnancy. To meet the fasting energy needs
Mechanisms/pathophysiology of pregnancy, basal endogenous glucose production
The metabolic abnormalities underlying GDM include (primarily hepatic) increases by 30% in healthy preg-
increased insulin resistance and β-​cell defects. These nant women by the end of gestation, despite a substantial
defects most likely exist before conception in many increase in fasting insulin levels51. However, circulating
cases, especially in populations with high rates of dia- fasting glucose concentrations decrease during preg-
betes and obesity. However, these defects are almost nancy, most likely because of an increase in plasma vol-
entirely asymptomatic and are generally detected only ume in early pregnancy and increased glucose utilization
because of widespread testing of glucose levels during in later gestation by the feto-​placental unit. Peripheral

4 | Article citation ID: (2019) 5:47 www.nature.com/nrdp

0123456789();
Primer

South and North Sub-


Western South-East Central America and Middle East and Saharan
Pacific Asia America Caribbean North Africa Europe Africa

35

30
Prevalence of GDM (%)

25

20

15

10

0
Th ore
Vi nd

Ne C m
Ze a
Au and
lia

al n
sia
ng dia
Sr esh
ka
ba

rb l
os

To A
Ar Ca o
Em da

Q s
Ba ar
n

l
No an
ay

Be K
Hu uim

Fr in
ce
itz key

ee d
Ge and

Sw ny

Ire n
nd
Ta eria

ia
Ba razi

ae
te

ar
w hin

ai

e
g

Gr rlan
U
S

an
na

a
at

rw
ad
an
ra

Ir

an

a
p

Cu

ed
ab na
ba
la

la
ay

an U

Isr
hr
ira

Sp
Ba In

ng
ap

Sw Tur
d

rm
lg

g
Ja
al

nl
B

nz
st
et
ai

iL
la

Ni
e
ng

d
Si

ad

d
id

ite
in

WHO NDDG C&C IADPSG Other


Un
Tr

Fig. 2 | Country-​specific prevalence of GDM according to different diagnostic criteria. Graph of prevalence
of gestational diabetes mellitus (GDM) in selected countries according to the Carpenter–Coustan criteria (C&C),
International Association of Diabetes and Pregnancy Study Groups (IADPSG) criteria, National Diabetes Data Group
(NDDG) criteria, WHO 2013 criteria and International Classification of Diseases codes and local guidelines or criteria
(other). A literature search was conducted in PubMed, supplemented by cross-​checking relevant references of eligible
studies on the prevalence of GDM from 1 January 2005 to 1 December 2018 to capture the contemporary burden of GDM.
Among the eligible studies that met the search criteria13, data from countries reported in the studies were included to
derive country-​specific estimates of GDM prevalence on the basis of different diagnostic criteria. The median of all
available source data was used if more than one estimate of GDM prevalence was available for a country.

insulin sensitivity (defined as the ability of insulin to who are later diagnosed with GDM53. By contrast,
increase glucose uptake in skeletal muscle and adipose women who develop GDM have less suppression (by
tissue) decreases by approximately 50% by late gesta- 80–85% compared with almost 100%) of endogenous
tion52. In women with normal glucose tolerance, there glucose production, thus contributing to postprandial
is a 2–3-fold increase in insulin secretion in response to hyperglycaemia in this population.
the decreased insulin sensitivity that maintains eugly­ In non-​pregnant women with normal glucose toler­
caemia. Maternal amino acid and lipid metabolism ance, the binding of insulin to the cell surface insulin
are also affected by the decreased insulin sensitivity receptor in peripheral tissues, such as skeletal muscle,
during pregnancy and are also related to the increased results in glucose uptake by cells. This interaction acti-
risk of increased fetal growth and adiposity but are not vates then results in autophosphorylation by the tyro­
discussed owing to space considerations. sine kinase domain of the insulin receptor β-​subunit
(IRβ), which activates a signalling cascade that induces
Pathophysiology of GDM redistribution of glucose transporter type 4 (GLUT4) to
Insulin resistance. In women who are normoglycae- the cell surface to enable glucose uptake by the cell. As
mic before pregnancy but go on to develop GDM in noted previously, there is a decrease in sensitivity with
late gestation, there is evidence of decreased peripheral advancing gestation during pregnancy, and this is further
insulin sensitivity before conception53 (Fig. 5). These decreased in women who develop GDM, both before
women initially adaptively maintain normoglycaemia in and during pregnancy55 (see Fig. 7 for the defects in the
early pregnancy because of the ability of the pancreatic insulin signalling cascade during pregnancy and known
β-cells to increase their insulin response. However, by defects associated with GDM). In late pregnancy, skeletal
late pregnancy, as insulin resistance increases, the insu- muscle content of one of the signalling molecules, insulin
lin response is inadequate. This defect in β-​cell function receptor substrate 1 (IRS1), is lower than in non-​pregnant
exists before pregnancy in many cases but only clinically women. In addition to the decrease in IRS1 levels, auto-
manifests with the increased insulin resistance of preg- phosphorylation of IRβ is lower in women with GDM
nancy, resulting in hyperglycaemia54 (Fig. 6). Although than in pregnant women with normal glucose tolerance,
endogenous glucose production is increased by 30% in which results in 25% lower glucose uptake in biopsied
women with normal glucose tolerance in pregnancy, it skeletal muscle56.
is almost completely suppressed during insulin infu- Historically, insulin resistance during pregnancy has
sion with a hyperinsulinaemic–euglycaemic clamp been ascribed to the effects of hormones released by
in women with normoglycaemia before conception. the placenta, such as human placental lactogen (HPL;
However, suppressed endogenous glucose production also known as choriomammotropin) and placental
may contribute to fasting hyperglycaemia in women growth hormone (PGH)57. Although concentrations of
with (again often undetected) IFG before pregnancy, these hormones are higher in later gestation, a specific

NATURE REvIEWS | DiSeASe PRiMeRS | Article citation I­D:­ (2019) 5:47 ­ 5

0123456789();
Primer

Box 1 | Risk factors for GDM supports the concept that pregnancy need not have a
long-​term detrimental effect on a woman’s metabolic
• Overweight or obesity (body mass index (BMI) ≥25 kg m–2) health62. However, weight gain postpartum that is asso-
• Advanced age ciated with increased parity could increase the risk of
• Non-​white ancestry T2DM owing to progressive insulin resistance and fur-
• Family history of type 2 diabetes mellitus ther deterioration of β-​cell function63, which may be par-
• Previous history of gestational diabetes mellitus (GDM) ticularly important for women who developed GDM and
• Parity (number of pregnancies >20 weeks) have a greater lifetime risk of future metabolic problems.
• Male fetus
Mechanisms of fetal consequences of GDM
• Multiple pregnancy
The placenta is central in the maternal–fetal supply line,
• Genetic factors as it integrates maternal exposures and provides oxygen,
• Polycystic ovary syndrome macronutrients and micronutrients to the developing fetus.
• Cigarette smoking The interactions of these factors with the fetal geno­type
• Psychosocial factors (for example, depression in (including epigenotype) determine the fetal pheno­type64.
pregnancy) The placenta partitions maternal fuels to cover its own
• Unhealthy dietary factors before pregnancy needs and at the same time to sustain fetal growth and
• Physically inactive lifestyle before and during pregnancy thereby mediates the effects of maternal metabolic dis-
turbances on the fetus. These metabolic changes encom-
pass foremost hyperglycaemia, which is both the means
mechanism relating to decreased insulin-​mediated glu- for diagnosing GDM and the principal therapeutic target.
cose disposal has not been described. Inflammation, However, it is important to remember that the metabolic
often linked to obesity, has been identified as another derangements also include elevated fatty acid and amino
factor that disrupts the insulin signalling cascade58. acid concentrations in the maternal circulation.
Tumour necrosis factor (TNF) activates a signalling
pathway that increases levels of sphingomyelinase and Placental nutrient transport. The placenta is richly
ceramides, which interfere with insulin receptor tyro­sine endowed with transporter molecules that ensure an ade-
autophosphorylation, and promotes serine phospho- quate supply of glucose, lipids and amino acids when
rylation of IRS1, which disrupts the insulin signalling the mother is metabolically normal, although it does not
cascade. In pregnancy, circulating TNF levels are signifi­ protect the fetus from oversupply in GDM65.
cantly negatively correlated with insulin sensitivity, even Transplacental glucose transfer becomes saturated
after adjustment for maternal fat mass59. only when the glucose concentration difference between
the maternal and fetal circulation is ≥25 mmol l–1 (ref.66).
Postpartum consequences of GDM. In the postpartum This high efficiency explains why glucose transfer is
period, within days of delivery of the placenta, there is a unaltered in GDM at the level of the placenta itself67.
rapid and substantial 120% increase in insulin sensitivity Therefore, the glucose concentration gradient between
and decrease in insulin response in comparison with late the maternal and fetal circulation is the most important
gestation60. A series of studies examining women with determinant of the amount of maternal glucose reaching
normal glucose tolerance and GDM in late pregnancy, the fetus. The gradient is determined not only by mater-
involving skeletal muscle biopsies and clamp studies at nal glycaemia but also by the glucose level in the fetus.
1 year postpartum, revealed that improvement in insulin Fetal glycaemia is modified by fetal insulin levels, which
resistance was associated with weight loss and increased are usually elevated in GDM (see below). Fetal hyper-
skeletal muscle expression of IRβ and IRS1. The change insulinaemia facilitates glucose uptake into peripheral
in IRS1 levels was correlated with the change in insulin tissues and steepens the concentration gradient. In these
sensitivity (regression coefficient 0.84, P < 0.007)59. conditions, the fetus also pulls (‘steals’) glucose from the
In women with previous GDM in whom there was maternal circulation, the ‘fetal glucose steal’ phenom­
no significant improvement in insulin sensitivity, body enon68, with the consequence that more maternal glucose
weight or body composition measures and circulating reaches the fetal circulation.
and skeletal muscle TNF concentrations remained ele- Placental transfer systems are much less efficient
vated 1 year postpartum. Although skeletal muscle IRβ for fatty acids than for glucose, and therefore only ~3%
and IRS1 levels improved, insulin-​stimulated insulin of maternal fatty acids reach the fetal circulation69,70.
receptor autophosphorylation and receptor tyrosine Efficient transfer is not needed, as the fetus can synthe-
kinase activity did not improve. Levels of skeletal mus- size its own non-​essential fatty acids using glucose as a
cle 312Ser-​IRS1 also did not improve and correlated with precursor (Fig. 8). Consequently, only 20% of the fatty
TNF expression61. acids in neonatal fat are derived from maternal sources71.
These data are consistent with a state of chronic Fatty acid transfer does not seem to be altered by GDM,
inflammation and insulin resistance in GDM and stress except perhaps that of docosahexaenoic acid72, which is
the importance of postpartum retention of excess weight important for the development of the brain and retina.
gained during pregnancy as a significant risk factor for The fetus depends on maternal supply of this essen-
the development of obesity and T2DM. Returning to pre-​ tial fatty acid. GDM reduces the placental levels of
pregnancy body weight is associated with a substantial the docosahexaenoic acid transporter NLS1 (encoded
improvement in the overall metabolic condition, which by MFSD2A) by ~30%73. NLS1 levels correlate with

6 | Article citation ID: (2019) 5:47 www.nature.com/nrdp

0123456789();
Primer

cord blood concentrations of docosahexaenoic acid. red blood cells reflect some degree of fetal hypoxia.
The reduced placental NLS1 levels may explain the lower The placenta responds to the increased fetal oxygen
cord blood levels of docosahexaenoic acid in women demand by increasing its capillary surface75. Low oxygen,
with GDM than in healthy women. hyperinsulinaemia and changes in the levels of several
At the end of pregnancy, only ~9–10% of the placen- other angiogenic factors in the fetal circulation in GDM
tal surface is involved in mediating nutrient transfer to stimu­late placental angiogenesis76,77. Whereas these
the fetus74, and this proportion is unaltered in GDM. regu­latory signals are derived from the fetus, others may
Nutrients taken up across the vast majority of the placen- come from the trophoblast and macrophages, both of
tal surface instead enter metabolic pools in the placenta which are essential cell types for placental function78,79.
to sustain placental functions65. The number and function of these cell types may also
Collectively, at the end of a GDM pregnancy, the pla- be altered in GDM, including changes in the molecules
centa does not actively enhance the quantum of mater- they secrete, which contribute to regulation of placental
nal nutrients reaching the fetal circulation and thus vasculari­zation. Overall, multiple signals give rise to
does not directly contribute to excessive fat accretion placental hypervascularization in GDM.
that leads to the characteristic phenotype of fetuses in Other examples of placental adaptations that ‘buffer’
GDM pregnancies. the potentially adverse effects of the maternal environ­
ment in GDM on fetal growth and development include
Buffering capacity of the placenta. Many of the changes an enhanced placental capacity to cope with increased
in the placenta of women with GDM are adaptive cholesterol synthesis in placental endothelial cells.
responses to protect both the placenta and fetus, of Multiple cellular and molecular mechanisms that facili­
which placental hypervascularization is the best stud- tate cholesterol removal from the feto-placental circu­
ied example. Fetal aerobic metabolism is stimulated lation to avoid the formation of pre-atherosclerotic
by hyperinsulinaemia in GDM pregnancies, and ele- lesions (which would reduce blood flow) are upregulated
vated cord blood concentrations of erythropoietin and in GDM80,81.
The placenta seems to have evolved some capacity
to buffer the intrauterine environment by adapting its
Mother Fetus functions to altered conditions in this environment,
Pre-pregnancy risk factors although this adaptive capacity is likely to be limited82.
(e.g. obesity and inflammatory cytokines (such as TNF)) Thus, extreme perturbations of the maternal milieu, as
in untreated GDM or GDM combined with obesity, may
Excessive peripheral Insufficient insulin override the placental buffering capacity and thereby
insulin resistance production contribute to pathological effects in the fetus83. Some evi-
dence exists to suggest that placental adaptive responses
are more pronounced in female fetuses84–87.
As a fetal tissue, the placenta is under fetal con-
trol, especially in the second half of gestation, when
Hyperglycaemia
fetal organs have formed. Consequently, the placenta
Excessive endogenous Excessive peripheral Placental-
glucose production insulin resistance related and is less vulnerable to an adverse maternal environment
hormones hyperinsulinaemia in this period than in early pregnancy, when the pla-
centa is mostly under maternal control88. For example,
↑ Glucose production ↓ Glucose uptake
• Short-term the placenta has poor antioxidative defences (such as
consequences lower levels of the antioxidant enzyme catalase) in the
(e.g. macrosomia
first 10–12 weeks of pregnancy89, resulting in the pla-
and neonatal
Hyperglycaemia hypoglycaemia) centa being especially sensitive to oxidative and meta-​
• Long-term inflammatory stress, which often occurs in women
consequences with hyperglycaemia, obesity and/or GDM90,91. Future
• Short-term consequences (e.g. pre-eclampsia) (e.g. T2DM and studies should investigate whether and how early hyper­
• Long-term consequences (e.g. T2DM) obesity)
glycaemic events in women who will develop GDM
later in pregnancy affect the growth and developmental
Fig. 3 | Pathophysiology of GDM. Women who develop gestational diabetes mellitus trajectories of the placenta and, subsequently, the fetus92.
(GDM) during pregnancy have evidence of metabolic dysfunction before conception,
such as pancreatic β-​cell defects and increased insulin resistance. In high-​income Fetal phenotype and long-​term effects. Maternal glucose
countries, many women who develop GDM are overweight or obese, which is associated is the main macronutrient that sustains fetal growth
with an inflammatory milieu. With the onset of pregnancy and associated metabolic (Fig. 8). In pregnant women with T1DM, prolonged expo-
changes (increased insulin resistance and demand for increased β-​cell response because sure of the fetal pancreas to hyperglycaemia from the
of placental factors), insulin is less effective in suppressing endogenous (primarily hepatic) early stages of pregnancy, which may also occur in GDM
glucose production and glucose uptake by peripheral skeletal muscle and adipose tissue,
(but remain undetected until diagnosis), accelerates mat-
which results in clinical hyperglycaemia. Maternal hyperglycaemia results in increased
placental transfer of glucose and (fetal) β-​cell secretagogues, such as amino acids, to the uration of the stimulus–secretion coupling mechanism in
fetus, leading to fetal hyperinsulinaemia. Fetal hyperinsulinaemia then results in fetal pancreatic β-​cells and results in early hyperinsulinaemia,
metabolic reprogramming that leads to short-​term problems, such as fetal overgrowth with ensuing fetal hyperglycaemia. Some amino acids,
and/or adiposity, and long-​term problems, such as metabolic dysfunction in later life. such as arginine, also stimulate the fetal pancreas and
T2DM, type 2 diabetes mellitus; TNF, tumour necrosis factor. contribute to hyperinsulinaemia. Free fatty acids (FFAs)

NATURE REvIEWS | DiSeASe PRiMeRS | Article citation I­D:­ (2019) 5:47 ­ 7

0123456789();
Primer

100
GDM included pre-​eclampsia, polyhydramnios, operative
T2DM delivery, shoulder dystocia, birth canal lacerations, fetal

Insulin resistance and β-cell


80 overgrowth (also called macrosomia), neonatal hypo-
glycaemia, jaundice and, in some studies of untreated

dysfunction (%)
60 Normal glucose GDM, perinatal mortality97–102. Furthermore, a graded
tolerance increase in the risk of maternal, fetal and neonatal com-
40 plications occurs with rising maternal glucose, even
within what is generally considered the normal plasma
20 glucose range103–106.
However, women with GDM often have other risk
0 factors for poor outcomes, including maternal over-
0 20 40 60 weight, increasing age, reduced physical activity or
Lifespan (years and increasing weight)
belonging to an ethnic minority. Thus, for many years it
Fig. 4 | Pregnancy as a metabolic stress test for future was intensely debated whether the poor outcomes asso-
metabolic disorders. Schematic representation of the risk ciated with GDM were due to maternal hyperglycaemia
of type 2 diabetes mellitus (T2DM) in women with normal per se or other risk factors107. Subsequently, the large
glucose tolerance and gestational diabetes mellitus (GDM), multinational landmark Hyperglycemia and Adverse
based on pregravid metabolic status of insulin sensitivity and Pregnancy Outcomes (HAPO) study108 clearly docu-
β-​cell dysfunction, with increasing age and body weight. mented that maternal hyperglycaemia independently and
Higher insulin resistance and β-​cell dysfunction preceding
in a graded linear way (without obvious cut-off points)
pregnancy in women who develop GDM increases risk of
T2DM in later life (that is, exceeding the glycaemic threshold increases the risk of pre-​eclampsia, preterm delivery, cae-
for T2DM (dashed line)), whereas women with normal sarean section, large for gestational age (LGA) infants,
glucose tolerance (that is, no insulin resistance and normal shoulder dystocia, neonatal hypogly­caemia, hyperbiliru-
β-cell function) who undergo the metabolic changes in binaemia and admission to neonatal special care units108.
pregnancy but return to their normal trajectory of metabolic The absolute risk of these complications in women with
changes (that occur owing to increasing age and weight) are GDM diagnosed using the IADPSG criteria ranges from
at lower risk of T2DM in later life. 1.8% for shoulder dystocia to 16.6% for neonatal adi-
posity (absolute outcome frequencies are summarized
are released from maternal lipoproteins by lipolysis, but in Table 2). Overall, fasting plasma glucose values from
only a small proportion crosses the placenta and contrib- the OGTT were more strongly associated with poor out-
utes to the fetal FFA pool. This pool mainly comprises comes than were the 1-hour and 2-hour values. Two large
FFAs produced by de novo lipogenesis in the fetal liver, randomized controlled trials have clearly shown that
using glucose as a precursor, which is present in excess treatment of GDM is effective in reducing or preventing
in maternal overnutrition. Fetal insulin stimulates tri- maternal and fetal short-​term complications, in particu-
glyceride synthesis and, thus, fat storage in white adipo- lar with reduction in LGA frequency to within the normal
cytes in the fetus in a sex-​dependent manner, which is expected range and of pre-​eclampsia by ~50%109,110.
reflected by a stronger association of cord blood insulin
with neonatal fat deposition in males than in females93. Long-​term maternal consequences. It has been known
GDM also leads to long-​term metabolic effects in since the original diagnostic criteria for GDM by
offspring. The pathogenetic mechanisms underlying O’Sullivan3 that women with elevated glucose levels
these abnormal metabolic characteristics are not known, in pregnancy are at an increased risk of subsequently
but maternal hyperglycaemia-​induced changes in DNA developing diabetes (primarily T2DM). Risk estimates
methylation and microRNA (miRNA) content in fetal have been obtained for different populations and vary
blood, skeletal muscle and adipose tissue94–96 and other depending on the population studied and the GDM
factors are most likely involved. criteria used. A meta-​analysis found a more than sev-
enfold increased risk of T2DM in women with GDM
Clinical consequences compared with women with normoglycaemic preg-
The first description of GDM arose from the observation nancies111. Thus, GDM is the best-​known risk factor
that parous pregnant women with overt diabetes often for T2DM112. Increasing BMI, GDM diagnosis early
had the same complications in pregnancies antedating in pregnancy, higher glucose levels at the time of diag-
their own diagnosis of diabetes as those in pregnant nosis during pregnancy, need for insulin treatment
women with diabetes, which was speculated to be due during pregnancy and IGT in the postpartum OGTT
to undetected prediabetic hyperglycaemia in previous are some of the risk factors for subsequent diabetes in
pregnancies. The GDM diagnostic criteria were based on women with previous GDM113–115. In 2018, the HAPO
the long-​term risk of maternal diabetes rather than the Follow-​Up Study (HAPO-​FUS)116 provided long-​term
short-​term risks of poor perinatal outcomes3. data about maternal and infant outcomes in women who
were diagnosed with GDM post hoc using IADPSG cri-
Short-​term consequences for mother and offspring. Later teria but who were untreated in the index pregnancy.
retrospective and prospective observational studies using This study provided data about the natural history of
these and similar diagnostic criteria clearly indicated untreated GDM (outcomes in the immediate peri­natal
that GDM was indeed associated with poor maternal period and after a mean of 11.4 years follow-up are
and offspring outcomes. The short-​term complications summarized in Table 2). Untreated GDM clearly has

8 | Article citation ID: (2019) 5:47 www.nature.com/nrdp

0123456789();
Primer

substantial long-​term risks for both mother and child. potential benefits from healthful diet and lifestyle
In some populations, risk of T1DM is also increased changes have been inconsistent135–137. Meta-​analyses
after GDM114. Furthermore, women with previous GDM and systematic reviews138–143 have also reached divergent
have an increased risk of the metabolic syndrome and conclusions. Both the timing of intervention (early to
cardiovascular, kidney, liver and retinal disease12,117–120. mid or late pregnancy) and the intervention approaches
themselves (yoga, aerobic exercises, resistance training
Long-​term offspring consequences. Studies using ani- and targeting different aspects of diet) were diverse in
mal models that are intended to simulate GDM have these studies. These methodological issues may have
documented that the offspring of GDM mothers have contributed to the markedly heterogeneous findings.
increased risk of hyperglycaemia, diabetes, obesity, car- A meta-​analysis published in 2015 provides some evi-
diovascular disease and structural hypothalamic changes dence that lifestyle modification (diet, physical activity
during their subsequent pregnancies and that these or both) initiated before the fifteenth gestational week
abnormal outcomes can be prevented by normalization can reduce the risk of GDM141. Antenatal dietary sup-
of maternal blood glucose levels during pregnancy121,122. plementation with myo-​inositol (a derivate of second-
These adverse outcomes are similar to clinical obser- ary messengers involved in several signalling pathways,
vations in children from different populations that including the insulin pathway) for the prevention of
included women with different types of diabetes melli- GDM is a comparatively new intervention. In two small
tus, which noted an increased risk of diabetes and obesity clinical trials, myo-​inositol supplementation reduced the
in children of women with diabetes mellitus123–125. risk of GDM144,145, whereas in a larger trial, supplemen-
In a follow-​up study from Denmark of the offspring tation with an inositol combination in early pregnancy
(18–27 years of age) of women with GDM, 21% of the did not prevent GDM in women with a family history of
offspring had pre-​diabetes or diabetes — an eightfold diabetes146. However, of note, the overall quality of this
increased risk compared with the background popu- evidence has been assessed as low or very low and the
lation11. Furthermore, the risk of overweight and the overall risk of bias is considered unclear144,147.
metabolic syndrome was higher (twofold and fourfold,
respectively)10 and insulin sensitivity and secretion were Probiotics
reduced126. In a study of nearly 100,000 pregnant women, Probiotics have potentially beneficial effects on glucose
children of women with GDM had increased fasting metabolism outside of pregnancy and have therefore
glucose levels, insulin resistance, adiposity and cardio­ been suggested as an intervention to prevent GDM.
vascular risk profile127. The HAPO-​FUS confirmed Initial enthusiasm was based mostly on positive findings
these findings but suggests that although maternal adi- in a randomized controlled trial from Finland. In that
posity is a strong risk factor for offspring obesity, GDM study, a probiotic preparation containing Lactobacillus
remains a significant risk factor, even after adjustment rhamnosus GG and Bifidobacterium animalis subsp. lactis
for maternal BMI116,128,129. BB-12 reduced GDM prevalence from 35% to 13%148.
Although differing results have been reported about However, findings in subsequent studies have been
the effects of GDM on cognitive function in offspring, largely negative149–152, and systematic reviews yielded con-
there is no solid evidence that maternal GDM inde- tradictory and inconclusive results153–155. Taking a global
pendently causes impaired cognitive function125,130. view, there is currently no firm evidence supporting the
Some studies have found that the offspring of women use of probiotics for the prevention of GDM.
with an early diagnosis of GDM have an increased risk
of autism spectrum disorder131, whereas the offspring of 0.3
Insulin sensitivity index

women with GDM needing medical treatment have Control


an increased risk of attention-​deficit hyperactivity 0.2 GDM
disorder (ADHD)132.
In contrast to the well-​d ocumented benefits of
GDM treatment in reducing immediate maternal and 0.1

fetal complications and the findings in animal studies,


treatment of GDM does not seem to improve the long-​ 0.0
term prognosis of offspring133,134. However, postpartum Pregravid Early pregnancy Late pregnancy
follow-​up studies are still of relatively short duration Fig. 5 | Changes in insulin sensitivity during pregnancy
(4–10 years) and long-​term results are awaited. in normoglycaemic women and women with GDM.
In conclusion, GDM is part of a vicious circle — dia­ Longitudinal changes in insulin sensitivity during a
betes begets diabetes, without any proven interventions hyperinsulinaemic–euglycaemic clamp (40 mU m–2 per
to interrupt or mitigate this cycle (Fig. 9). minute insulin infusion) in women with normal glucose
tolerance (control) and those who developed gestational
Diagnosis, screening and prevention diabetes mellitus (GDM). Data are presented as the mean
and s.d. The x-​axis indicates longitudinal changes before and
Prevention of GDM
during pregnancy. The y-​axis indicates the change in insulin
A number of clinical trials of dietary and lifestyle sensitivity index, which is defined as the glucose infusion
changes, dietary supplements and pharmacological rate to maintain euglycaemia (90 mg dl–1) plus residual
medications for the prevention of GDM have been con- endogenous glucose production during the clamp/mean
ducted. Despite promising findings from observational insulin concentration during insulin infusion at 40 mU m–2
studies, intervention studies seeking to demonstrate per minute. Adapted with permission from REF.53, Elsevier.

NATURE REvIEWS | DiSeASe PRiMeRS | Article citation I­D:­ (2019) 5:47 ­ 9

0123456789();
Primer

1,000 that two elevated values would be required for a GDM


900 Normoglycaemic diagnosis. Using this definition, GDM prevalence in
800 GDM the cohort was 1.9% and was predictive of risk of later
Third trimester developing T2DM.
Insulin secretion rate

700
Postpartum These original ‘O’Sullivan criteria’ for a GDM diag-
600
nosis were subsequently modified in the late 1980s
500 and early 1990s (to account for changes in laboratory
400 methodology) by Carpenter and Coustan160 and the
300 National Diabetes Data Group (NDDG)161. The NDDG
200 criteria erroneously failed to correct for the measure-
ment of substances other than glucose in the original
100
O’Sullivan studies and thereby produced higher diagnos-
0
0.0 0.1 0.2 0.3 0.4
tic threshold glucose levels for GDM. However, both the
Insulin sensitivity index
Carpenter–Coustan and NDDG criteria are still recog­
nized by the American College of Obstetricians and
Fig. 6 | Insulin sensitivity–secretion relationships in normoglycaemic women and
Gynecologists (ACOG)162, despite the fact that they were
women with GDM. Prehepatic insulin secretion was assessed during steady-​state
hyperglycaemia using measurement of plasma insulin and C-​peptide concentrations derived from an empirical analysis of a small cohort of
and kinetics in individual patients during the third trimester and postpartum. The women in Boston in the late 1950s and developed as pre-
disposition index depicts the relationship of changes of insulin secretion together with dictors of later T2DM, without consideration of their
insulin sensitivity. In women who developed gestational diabetes mellitus (GDM) during relationship to pregnancy outcomes (despite these short-
pregnancy, when not pregnant (postpartum, yellow circles) there is decreased insulin comings and owing to their widespread use in the USA,
sensitivity and slightly lower insulin secretion than in postpartum women with normal we include these criteria in Table 1).
glucose tolerance. In the third trimester of pregnancy (blue circles), however, although The HAPO study 108,163–165 provided large-​s cale
there is a further decrease in insulin sensitivity in both groups, in women who develop blinded data of the association between mild hyper­
GDM there is a less marked increase in insulin secretion to match the decreased glycaemia in pregnancy and maternal and fetal out-
insulin sensitivity, which then results in hyperglycaemia. Adapted with permission from
comes. In a cohort of 23,316 women (>30-fold larger
REF.54, Oxford University Press.
than the O’Sullivan cohort) who completed a blinded
75 g OGTT at 24–32 weeks of gestation, fasting, 1-hour
Pharmacological interventions and 2-hour glucose values in the OGTT were linearly
Pharmacological interventions during pregnancy for associated with a broad range of pre-​defined, carefully
GDM prevention are rare, with most studies testing ascertained and adjudicated adverse clinical and bio-
metformin in high-​r isk populations, such as over- chemical outcomes of pregnancy. These independent
weight and obese women156,157 or women with a history associations of hyperglycaemia with pregnancy out-
of polycystic ovary syndrome158,159. Overall, data from comes remained strong after extensive adjustment for
existing studies do not demonstrate a protective role potential confounders, including maternal BMI, age,
of metformin in the prevention of GDM. Large-​scale, height, mean arterial pressure and parity.
multiple-​arm inter­vention studies, in which the inter-
vention is initiated early in or even before pregnancy, Current screening and diagnostic criteria. Following the
among multi-​ethnic popu­lations are sorely needed. HAPO study, it was evident that there are no ‘natural
Ideally, these studies should be sufficiently powered to inflection points’ in the associations between glycaemia
detect effects on maternal and neonatal outcomes and and adverse outcomes that readily identify a ‘natural’
should obtain long-​term follow-up data for both mother set of diagnostic thresholds for GDM. After an exten-
and child. sive consensus process, taking into consideration the
results of both the HAPO study and other published
Screening and diagnosis studies, the IADPSG published recommendations in
Historical definition of GDM. The prevailing defini- 2010 for the identification and classification of hyper­
tions of GDM are based on the landmark 1964 paper glycaemia in pregnancy166. The underlying principles of
by O’Sullivan and Mahan3. These authors reported on the IADPSG consensus process were that women with
an unselected cohort of 986 women from Boston (USA) equivalent levels of glycaemia-​associated risk of adverse
who were enrolled over a 4-month period. A subgroup pregnancy outcomes should be classified in a similar
(752 women) completed both a 50 g non-​fasting glucose manner and that threshold blood glucose values should
challenge test (GCT) at their first antenatal presentation be standardized internationally. The IADPSG recom-
and later a formal 3-hour, 100 g oral glucose tolerance mended a ‘one-​step’ method involving an OGTT at
test (OGTT), and their results are the historical basis for 24–28 weeks of gestation and proposed GDM diagnos-
the diagnosis of GDM in the USA. Their OGTT values tic thresholds (Table 1) that were based on an adjusted
were used to derive 97.7 percentile levels (2 s.d. above the odds ratio threshold of 1.75 (compared with the odds at
cohort mean) for the 100 g OGTT in pregnancy. After the HAPO cohort mean) of delivering an infant affected
rounding of the 2-hour and 3-hour values, these results by key fetal complications of maternal hyperglycaemia,
provided the initial threshold whole blood glucose values namely, increased size at birth, increased adiposity and
for GDM (fasting 90 mg dl–1; 1 hour 165 mg dl–1; 2 hour elevated cord blood C-​p eptide levels. The IADPSG
145 mg dl–1; 3 hour 125 mg dl–1), with the equally arbi- also recognized that undiagnosed T2DM in pregnant
trary decision (noted as “it was considered expedient…”) women is increasingly prevalent in certain populations

10 | Article citation ID: (2019) 5:47 www.nature.com/nrdp

0123456789();
Primer

Normal glucose tolerance Normoglycaemic GDM


(non-pregnant) pregnancy

Insulin Glucose Glucose Glucose

GLUT4
Insulin P P P
receptor P P
P P P
P P
P P P
pSer pTyr pSer pTyr pSer pTyr
IRS1 IRS1 IRS1
Insulin Insulin Insulin
PI3K action PI3K action PI3K action

Fig. 7 | Changes in insulin signalling in normoglycaemic and GDM pregnancies. Schematic representation of pregnancy-
related changes in insulin signalling. Insulin signalling during pregnancy in women with normal glucose tolerance requires
tyrosine autophosphorylation of the insulin receptor in skeletal muscle. This is the initial step in the insulin signalling
cascade, allowing recruitment and activation of downstream effectors, such as insulin receptor substrate 1 (IRS1)
and phosphatidylinositol 3-kinase (PI3K), resulting in translocation of glucose transporter type 4 (GLUT4) to the plasma
membrane and thereby leading to increased glucose uptake into skeletal muscle55. In late pregnancy, skeletal muscle IRS1
content is lower (red arrow) than in non-​pregnant women. In gestational diabetes mellitus (GDM) pregnancies, in addition
to the decrease in IRS1, tyrosine autophosphorylation in the intracellular domain of the insulin receptor β-subunit is
reduced, which results in 25% lower in vitro glucose uptake than in pregnant women with normal glucose tolerance56.

and recommended that these cases should ideally be The two-​step method imposes less of a diagnostic
detected early in pregnancy and classified as ‘overt burden on many women than the one-​step method but,
diabetes’ (also termed ‘diabetes in pregnancy’ in the depending on the thresholds used to determine the need
subsequent WHO adaptation of the IADPSG criteria)8 for a full diagnostic OGTT, requires that up to 30% of
to both identify them as a particularly high-​risk group women complete a second test. However, a systematic
and ensure rapid treatment (see Table 1 for a summary review demonstrated that the two-​step method misses
of the recommended classification). approximately 25% of women with GDM on a formal
Subsequently, the IADPSG recommendations were OGTT180, whereas another systematic review concluded
widely endorsed as the preferred diagnostic criteria by that the one-​step method is associated with improved
major national bodies167–171 and international bodies8,172 pregnancy outcomes181. Ensuring reliable follow-​up
but were disputed by others162,173,174. The IADPSG crite- after a positive GCT result is also important; only 36%
ria are largely followed in Australia169 and Japan170 and of women in a study from eastern Canada received
are officially endorsed in Europe175 and globally by the appropriate further testing after a positive GCT result182,
International Federation for Gynecology and Obstetrics whereas 75% of women in a study from western Canada
(FIGO) and the International Diabetes Federation172,176. received appropriate follow-​up testing within 2 weeks183.
However, they are not widely used in the USA or Insistence on two or more elevated values on the
Canada, although the 2018 ACOG guidelines177 and 2018 OGTT to make a GDM diagnosis is essentially a histor-
Canadian guidelines178 include the IADPSG process as ical quirk related to the empirical decision in favour of
one (not the preferred) option. The UK guidelines issued this approach by O’Sullivan and Mahan in their land-
by the National Institute for Clinical Excellence (NICE) mark 1964 publication3. The two-​step method limits the
remain at variance with other countries in their contin- number of women who are diagnosed with GDM but
ued support for selective risk-​factor-based testing179. The not in a logical manner, as demonstrated in multiple case
principal points of contention relate to increased num- series (reviewed in ref.184).
bers of women potentially diagnosed with GDM owing In summary, GDM is diagnosed by the detection
to the change in thresholds and from a ‘two-​step’ method of hyperglycaemia in pregnancy, which, although less
(screening test followed by a diagnostic test) to a one-​step severe than overt diabetes, is associated with increased
method (a diagnostic test only, the 75 g glucose 2-hour pregnancy complications, particularly those related to
test). In most settings, the two-​step method involves a excessive fetal growth. In the absence of natural thresh­
non-​fasting GCT in which plasma glucose is tested 1 hour olds, the precise numeric cut-off values of maternal
after ingestion of a 50 g or 75 g glucose load, which is fol- glycaemia used for diagnosis are the subject of ongoing
lowed by a full fasting OGTT if the GCT falls above a pre-​ debate. Overall consensus currently favours the IADPSG
defined threshold. The other issue, principally relevant and WHO criteria (outlined in Table 1), which are based on
to the USA, is the change from a requirement that two a consensus overview of available large-scale epidemio-
values on the OGTT should meet or exceed the threshold logical data and randomized controlled trials and relate
value to make a GDM diagnosis177 to the acceptance by the diagnostic thresholds to the risk of hyperglycaemia-​
IADPSG that only a single elevated value on the OGTT related pregnancy complications. However, they may
is sufficient to make a GDM diagnosis7. not be suitable for uniform worldwide application, as a

NATURE REvIEWS | DiSeASe PRiMeRS | Article citation I­D:­ (2019) 5:47 ­ 11

0123456789();
Primer

Maternal Placental Fetal be initiated shortly after diagnosis186,187. The diet should
contain sufficient macronutrients and micronutrients,
Amino limit postprandial glucose excursions and prevent
acids
excessive maternal gestational weight gain. The recom-
+ mended diet is similar to a general diabetes diet, includes
Glucose
Glucose ↑ Insulin
mainly low-​glycaemic-index carbohydrates and often
apportions daily caloric intake into three main meals and
2–4 snacks. To ensure sufficient fetal growth and cere-
bral development and function, the Institute of Medicine
Macrophage + (IOM) recommends a diet with ≥175 g of carbohydrates
daily188,189. Maternal ketonaemia and/or ketonuria in
Lipogenesis pregnant women with diabetes has been associated with
Liver lower mental and/or motor function in the offspring190,
Lipoproteins + and the minimum carbohydrate intake recommended
EL by IOM ensures sufficient carbohydrates to avoid star-
FFA FFA
FFA vation ketonaemia. A low-​glycaemic-index diet has been
pool
shown in randomized trials to be associated with less
White adipose frequent insulin use and lower birthweight than control
Syncytiotrophoblast Endothelium
tissue diets188. The average total caloric intake was 1,600 kcal
daily in the pregnant women in these trials, which
Fig. 8 | Pathophysiology of fetal phenotype in GDM. Maternal glucose is the main
probably reflects some degree of caloric restriction188.
macronutrient that sustains fetal growth. In women with gestational diabetes mellitus
Caloric restriction up to 30% of the required intake has
(GDM), prolonged fetal exposure to hyperglycaemia and/or some amino acids (for
example, leucine and arginine) induces hyperinsulinaemia. Free fatty acids (FFAs) derived been suggested as safe in obese pregnant women187, but
from maternal lipoproteins are released by endothelial lipase (EL)-mediated lipolysis on evidence is lacking in those who are not obese.
the surface of the placenta. Only a small proportion of these FFAs cross the placenta, Lifestyle intervention also includes recommendations
where they contribute to the fetal FFA pool, which comprises predominantly FFAs for daily physical activity, such as walking, cycling and
resulting from hepatic de novo lipogenesis, using the excess glucose (due to maternal swimming191. Diet and physical activity are sufficient to
overnutrition) as a precursor. Fetal insulin stimulates triglyceride synthesis and, thus, control the glycaemic status in ~70–85% of women with
fat storage in white adipocytes in the fetus in a sex-​dependent manner93. Solid lines GDM187,191. In addition to more appropriate fetal growth,
represent macronutrient flow; dotted lines represent regulating effects of fetal insulin. this lifestyle intervention is also associated with reduced
occurrence of postnatal depression among mothers109,
study in Denmark found a >40% GDM diagnostic rate and a systematic review demonstrated that exercise
in a low-​risk population using the IADPSG and WHO interventions reduce postpartum weight retention in
criteria, without a clear relationship of GDM diagnosis to pregnant women192.
adverse pregnancy outcomes185. The 2015 guidelines from
FIGO172 recommend a more flexible approach that allows Patient involvement and treatment goals
for differing diagnostic processes and glucose thresholds Women with GDM are encouraged to self-​monitor their
in specific geographic regions and ethnic groups. preprandial and postprandial blood glucose levels before
and 1–2 hours after main meals193. Glucose targets vary
Management between various international guidelines. At present, the
The primary goal of GDM treatment is the prevention American Diabetes Association (ADA) recommends fast-
of fetal overgrowth and pregnancy complications. This ing glucose values <95 mg dl–1 (5.3 mmol l–1) and either
aim is usually achieved by dietary modification, and 1-hour postprandial glucose <140 mg dl–1 (7.8 mmol l–1)
promotion of physical activity to minimize postprandial or 2-hour postprandial glucose <120 mg dl–1 (6.7 mmol l–1)
glucose elevations and pharmacotherapy is required only and, if measured, HbA1c levels <6%191. Initially, prepran-
in a minority of women. dial and postprandial glucose measurements with the
three main meals are recommended daily. If the gly­caemic
Lifestyle intervention goals are easily obtained within 2 weeks of initiating life-
Glucose is the main nutrient that promotes fetal growth style interventions, women may reduce the frequency
(although other nutrients, such as lipids and amino of preprandial and postprandial glucose measure­ments
acids, also influence growth). Increased maternal gly- to twice weekly as long as the values are within target.
caemia is directly associated with fetal overgrowth and Continuous glucose monitoring has become available at
many of the pregnancy complications in GDM (the a reasonable price and may replace finger-​prick tests in
Pedersen hypothesis). Therefore, control of maternal selected cases. In our opinion, an average blood glucose
hyperglycaemia logically remains the primary goal of level of 5–7 mmol l–1 can be recommended, although
GDM treatment. In two large, randomized controlled precise targets vary widely between and even within
studies, lifestyle advice (supplemented with insulin countries and regions. In the future, more detailed recom­
administration if necessary) reduced the prevalence of mendations for time in target and time below target for
fetal overgrowth (LGA; infant mass >90th percentile) by women with GDM may be developed.
~50% to within the normal expected population range In addition to hyperglycaemia, excessive gestational
and reduced the risk of shoulder dystocia109,110. Lifestyle weight gain is also associated with fetal overgrowth in
modification is the cornerstone of treatment and should both healthy and GDM pregnancies189,194, which has

12 | Article citation ID: (2019) 5:47 www.nature.com/nrdp

0123456789();
Primer

led to recommendations for weekly gestational weight-​ safe to postpone initiation of pharmacological treat-
gain goals that are dependent on maternal BMI before ment. Conversely, excessive fetal growth may lead to
pregnancy. For obese, overweight and healthy-​weight intensification of treatment with lower glycaemia goals195.
women, the IOM189 recommends a maximum weekly
weight gain of 220 g, 280 g and 420 g, respectively. There Insulin. Traditionally, insulin has been the primary med-
is some evidence that even lower targets for obese women ical treatment if the glycaemic treatment goals are not
may be safe and are associated with more appropriate achieved with lifestyle intervention within 1–2 weeks.
fetal growth194. Many clinicians will therefore accept a Insulin is effective and safe for the fetus, as it does not cross
diet aimed at weight stability, and some clinicians even the placenta. Human insulin and several insulin analogues
recommend some weight loss in obese women after (for example, insulin aspart, insulin lispro and insulin
diagnosis of GDM. To follow the recommendations for detemir) have been formally tested and are considered safe
weekly weight gain, the weight of the mother should be to use in pregnancy191. Insulin glargine is also often used
monitored carefully at frequent intervals after diagnosis and, although there are no randomized trials supporting
of GDM. Both caregivers and patients should be aware of the use of insulin glargine, a synthesis of published cohort
both the blood glucose targets and the weight-​gain goals. studies has not raised specific concerns about its use in
pregnancy196,197. The main issue with insulin therapy is
Medical treatment of GDM the burden for the women, which can include discomfort,
When glycaemia remains elevated after ≥1–2 weeks fear of needles, the cost of treatment and the risk of hypo-
of lifestyle interventions, daily glucose testing should glycaemia. Episodes of mild hypoglycaemia occur often,
be continued and pharmacological treatment should be whereas episodes of severe hypoglycaemia (that require
initiated. Ultrasonography-​based assessment of fetal help from a third party) are rare.
growth may also assist in guiding the intensity of glu- During medical treatment, the goals for glucose con-
cose control that is needed in an individual woman. trol and weight gain are the same as those for lifestyle
If the baby is growing appropriately, in particular if fetal interventions alone. The insulin treatment can be a
abdominal circumference is <75th percentile, it may be basal–bolus regimen using intermediate-​acting or long-​
acting insulin once daily and a rapid-​acting insulin before
main meals. Twice-​daily injections of a mixture of fast-​
Table 2 | Outcomes from the HAPO study and the HAPO-​FUS
acting and long-​acting insulin are also effective198. Insulin
Outcome GDMa Non-​GDM Statistical detemir has been associated with less hyperglycaemia
(%) (%) significance and hypoglycaemia than slow-​acting human neutral
Perinatal outcomesb protamine Hagedorn (NPH) insulin199. However, when
Pre-​eclampsia 9.1 4.5 P < 0.001 choosing the type of insulin, the cost of the insulin must
also be considered. An initial dose of 0.3 international
Preterm delivery (<37 weeks) 9.4 6.4 P < 0.001
units (IU) per kg (body weight) per 24 hours can be used
Primary caesarean delivery 24.4 16.8 P < 0.001 when initiating the insulin treatment, and a final insulin
Shoulder dystocia or birth injury 1.8 1.3 P < 0.01 dose close to 1 IU per kg (body weight) is often needed198.
Birthweight greater than ninetieth percentile 16.2 8.3 P < 0.001 Insulin treatment is time consuming for caregivers
and requires training and education of the pregnant
Neonate percentage body fat greater than 16.6 8.5 P < 0.001 women, and patient contact is frequently needed to
ninetieth percentile
adjust the insulin dose. Consequently, oral glucose-​
Cord blood C-​peptide level greater than 17.5 6.7 P < 0.001 lowering medications, such as two well-​established oral
ninetieth percentile
agents, metformin and the sulfonylurea glibenclamide,
Clinical neonatal hypoglycaemia 2.7 1.9 P < 0.001 have been studied in women with GDM. Other glucose-​
Admission to newborn intensive care 9.1 7.8 P < 0.01 lowering agents are generally not advocated for use in
Long-​term outcomes c pregnancy owing to documented complications, such as
neonatal hypoglycaemia, fear of unexpected fetal com-
Maternal diabetes 10.7 1.6 P < 0.001 plications and possible metabolic or epigenetic changes
Maternal pre-​diabetes 41.5 18.4 P < 0.001 in the developing fetus.
Offspring overweight or obesity 39.5 28.6 P < 0.001
Metformin. Metformin acts mainly by suppressing
Offspring obesity 19.1 9.9 P < 0.001
hepatic glucose production, leading to a reduction in
Offspring percentage body fat greater than 21.7 13.9 P < 0.001 fasting plasma glucose levels and HbA1c. Metformin is
eighty-​fifth percentile the first-​line treatment in non-​pregnant patients with
Offspring impaired fasting glucose (ADA 9.2 7.4 Not significant T2DM200. Although it is inexpensive and easy to use,
threshold of ≥5.6 mmol l–1) metformin commonly causes gastrointestinal symptoms,
Offspring impaired glucose tolerance 10.6 5.0 P < 0.001 may cause low vitamin B12 and, rarely, may increase the
Offspring diabetes 0.3 0.2 Not significant risk of lactic acidosis. Perhaps more importantly, met-
ADA, American Diabetes Association; GDM, gestational diabetes mellitus; HAPO, Hyperglycemia
formin crosses the placenta and thus can potentially
and Adverse Pregnancy Outcomes; HAPO-​FUS, HAPO Follow-​Up Study. aWomen classified affect the developing fetus. In randomized clinical trials,
post hoc as having GDM by International Association of Diabetes and Pregnancy Study Groups metformin seems to be comparable to insulin in glycae-
(IADPSG) criteria were compared to women without GDM. Women with GDM were not treated
during or after the index pregnancy. bOutcomes are from the HAPO study108. cOutcomes are from mic control and immediate neonatal outcomes201,202.
the HAPO-​FUS116. At least one-​third of women with GDM who are treated

NATURE REvIEWS | DiSeASe PRiMeRS | Article citation I­D:­ (2019) 5:47 ­ 13

0123456789();
Primer

Pregnancy registry study found higher rates of LGA, neonatal hypo-


• Increased insulin glycaemia, birth injury and neonatal admission to the
resistance intensive care unit with glibenclamide than with insu-
(obesity or GDM)
lin212. Together, these findings argue against the use of
• Metabolic
Metabolic inflammation glibenclamide as first-​line pharmacotherapy in women
ageing with GDM.
The major international guidelines for treatment of
women with GDM, including those from the ADA191,
Adult metabolic syndrome, Fetal–neonatal metabolic the Endocrine Society168, FIGO172 and the British NICE
T2DM and obesity programming of obesity
guidelines179, all recommend lifestyle interventions and
insulin as the cornerstones of GDM treatment but differ
regarding the possible use of metformin or glibencla-
• Childhood obesity mide in pregnancy. In summary, insulin remains the
• Pre-metabolic gold standard for pharmacotherapy of GDM, but met-
syndrome?
formin or glibenclamide may be chosen in individual
Fig. 9 | Application of the DOHAD hypothesis to GDM. cases depending on convenience and cost (Box 2).
Schematic representation of the cycle of intrauterine
exposure of offspring to maternal metabolic disturbances, Pharmacotherapy immediately after delivery. Pharma­
resulting in subsequent fetal metabolic programming, cotherapy for treatment of GDM can be stopped imme­
childhood obesity and metabolic disturbances and later diately after delivery, although glucose monitoring for a
leading to overt adult disease that contributes to increased few days to exclude marked ongoing hyperglycaemia
exposure in the next generation. DOHAD, Developmental is recommended. Healthy eating, which during breast-
Origins of Health and Disease; GDM, gestational diabetes feeding should include a carbohydrate intake of at
mellitus; T2DM, type 2 diabetes mellitus. least 210 g daily, is recommended. If hyperglycaemia
consistent with overt diabetes persists postpartum
with metformin also need additional treatment with (fasting glucose >7.0 mmol l –1 and/or postprandial
insulin203. Because metformin crosses the placenta, the glucose >11.0 mmol l–1), lifestyle interventions can be
long-​term effects of treatment in utero on the child must initiated and pharmacotherapy can possibly be reiniti­
be considered204. The adiposity and blood pressure in ated with insulin, metformin or glibenclamide, which
2-year-old offspring of mothers with GDM were compar­ are considered safe during lactation (Box 2).
able for metformin and insulin treatment205,206, although
children exposed to metformin in utero had, on average, Long-​term maternal complications
more subcutaneous fat at the upper arm. The 9-year-old Epidemiological studies indicate that maintaining
offspring of women treated with metformin were taller physical activity, adopting healthful dietary patterns213,
than those whose mothers were treated with insulin207. avoiding weight gain after pregnancy and frequent, pro-
In another randomized trial, offspring were heavier at longed and more intensive breastfeeding reduce the risk
1 year of age and both taller and heavier at 18 months of of progression to overt diabetes214–216. Consistent with
age if they had been exposed to metformin in utero208. this, lifestyle interventions and medical treatment both
decrease progression to diabetes by ~50% in women
Sulfonylureas. Sulfonylureas augment insulin secretion, with previous GDM in randomized controlled trials with
and the resulting hyperinsulinaemia leads to a decline in up to 10 years follow-​up217–219. Therefore, there is great
fasting plasma glucose levels and HbA1c. In a randomized potential for preventing or delaying the onset of T2DM
clinical trial, glibenclamide (also known as glyburide in and cardiovascular disease in these women. Guidelines
the USA) was as effective as insulin in achieving gly- recommend breastfeeding, a lifelong healthy life-
caemic control, and rates of LGA and overall perinatal style (including weight loss if necessary), an OGTT
outcomes were similar209. However, contrary to initial 2–6 months after delivery and thereafter assessment of
findings with early, less-​sensitive assays, glibenclamide glucose tolerance every 1–3 years191 using either fast-
crosses the placenta210,211, and the fetal:maternal concen- ing glucose levels, OGTT or HbA1c levels. In addition,
tration ratio is highly variable210 and may have an effect women with prior GDM should be screened for cardio­
on the developing fetus. Follow-​up studies of children vascular risk factors. Although these guidelines and
who have been exposed to glibenclamide antenatally are recom­mendations have been available for years, they
not available. are not generally implemented and followed in everyday
In a meta-​analysis comparing the efficacy and safety clinical practice, for various reasons114,220.
of insulin, metformin and glibenclamide203, glibencla-
mide was associated with higher birthweight and higher Quality of life
rates of macrosomia and neonatal hypoglycaemia than For most women, a GDM diagnosis will cause a consid-
insulin. Compared with metformin, glibenclamide was erable change in their perception of pregnancy. A med­
associated with higher birthweight and higher rates of ical diagnosis of GDM will change their pregnancy from
macrosomia. Insulin and/or metformin treatment is ‘normal’ to ‘abnormal’ and could potentially be associ-
therefore considered superior to glibenclamide treat- ated with anxiety for maternal and fetal health221,222.
ment. Glibenclamide has been widely used in women However, as formal studies are rare, little is known about
with GDM in the USA, and a US health-​care insurance this topic.

14 | Article citation ID: (2019) 5:47 www.nature.com/nrdp

0123456789();
Primer

A systematic review reported that QOL is more fre- population obesity. Therefore, many cases currently clas-
quently reduced in women with GDM than in pregnant sified as GDM probably represent pre-existing undiag­
women without GDM223. However, it is not clear whether nosed pre-diabetes, which is not specifically related
this is caused by GDM itself or is due to confounding to pregnancy226.
factors, such as obesity, gestational weight gain, socio-​ GDM is also clearly associated with increased risk
economic status or pre-​existing depression. Interestingly, of later maternal diabetes and cardiovascular disease.
in the landmark ACHOIS randomized controlled trial of Preventive strategies have been clearly elucidated227,228 and
treatment versus no treatment of GDM, treated women proved to be efficacious, but long-​term implementation
with GDM had a higher QOL score at 3 months post- on a broad population scale remains elusive.
partum than those who were not treated109. This result Unfortunately, current treatment protocols have not
most likely reflects the beneficial effect of the generally been proved to effectively reduce the risks of later obesity
increased care that women receive during pregnancy and impaired glucose metabolism in offspring133,134.
after a GDM diagnosis.
Depression has been reported more often in women Areas for future GDM research
with GDM, both before and after pregnancy, than in Despite many years of attempting to achieve interna-
pregnant women without GDM, indicating that GDM tional consensus on GDM diagnostic criteria, uniform
is not the cause of depression43,224,225. Thus, a knowledge recommendations remain elusive and the ‘pragmatic’
gap still exists about the effect of GDM on QOL, and approach of FIGO172, which allows for regional vari­
further studies are needed. ations so long as the overarching principles of uniform
testing and treatment are adhered to, may represent the
Outlook best available compromise solution.
Currently accepted features of GDM A key area of uncertainty is the definition of healthy
The evidence discussed above clearly supports the or desirable glucose levels and by inference ‘GDM equiv-
existence of the phenomenon of gestational diabetes alent’ glucose levels in early pregnancy (<20 weeks of
— hyperglycaemia arising during pregnancy owing to gestation), and both epidemiological studies and clinical
pathophysiological changes in gestation. Furthermore, trials in this area are eagerly awaited229,230. Furthermore,
although there is no clear or natural set of diagnostic additional biomarkers that predict pregnancy compli-
cut-​off levels for hyperglycaemia, current treatment cations are highly desirable, as are non-​fasting alterna-
protocols, including diet, exercise and insulin, reduce tives to the complex and burdensome OGTT. Plasma
immediate pregnancy complications, in particular exces- glycated CD59 is the most promising current candidate,
sive fetal growth and its consequences and maternal but further research is required before it can be widely
hypertensive disorders of pregnancy. adopted231. In addition, differences in fetal sex-​related
Conversely, the factors that influence the risk of GDM pregnancy outcomes232 and later outcomes134 are
developing GDM have changed, particularly advancing increasingly reported, and their underlying mechanisms
maternal age at the time of childbearing and increasing are a fertile area for future investigation87,233. Attempts to
understand the molecular mechanisms underlying sex
Box 2 | Key points in the treatment of GDM differences in GDM will inevitably require more studies
of placental factors (such as exosomes and hormones)
Lifestyle intervention with diet and physical activity is the cornerstone of treatment, and their effects on maternal insulin resistance and β-​cell
and only a minority of women with gestational diabetes mellitus (GDM) need function. It is also unclear whether and how the fetal
pharmacotherapy.
pancreatic β-​cell response to maternally derived insulin
Diet secretagogues, such as arginine and other amino acids, is
• Ensure sufficient intake of micronutrients and macronutrients, including >175 g of changed in GDM. Transplacental transfer of these amino
carbohydrates daily acids has not been studied in GDM, although they have
• Minimize glycaemic excursions by intake of low-​glycaemic-index carbohydrates been shown to stimulate insulin release, at least in the
divided over several meals and snacks daily first half of pregnancy, and they are also indirectly
• Provide appropriate level of gestational weight gain involved in regulating the vascular tone of the feto-​
Physical activity placental circulation234. GDM-​associated changes in vas-
• Light exercise, such as walking, swimming and cycling cular tone may result in umbilical blood flow differences
• Regimen appropriate for pregnancy between GDM and non-​GDM pregnancies235.
To date, attempts at prevention of GDM through
Pharmacotherapy lifestyle interventions, probiotics, inositol supplemen-
If the glycaemic goals are not attained within 1–2 weeks of initiating dietary changes tation and medications have not shown consistent bene­
and physical activity fits236, and none could be recommended for routine
• Insulin is the gold-​standard treatment for hyperglycaemia use. The focus of preventive efforts probably needs to
• Metformin may be chosen in selected cases, dependent on convenience and cost shift to the preconception period or very early in gesta­
Treatment after delivery tion to achieve the hoped-​for reductions in both GDM
• Continue lifestyle interventions prevalence and consequent pregnancy complications.
• Encourage breastfeeding Population-​based efforts to reduce the prevalence of obe-
sity and hyperglycaemia across the life course, especially
• Pharmacotherapy can be stopped immediately after delivery
in girls and young women, may ultimately be required to
• Regular screening for the development of diabetes mellitus is recommended
reverse the tide of these dual global epidemics.

NATURE REvIEWS | DiSeASe PRiMeRS | Article citation I­D:­ (2019) 5:47 ­ 15

0123456789();
Primer

Insulin therapy remains the most widely accepted results suggest that maternal glycaemia more strongly
pharmacological intervention for treating GDM. influences offspring glucose metabolic status than pre-
Glibenclamide (glyburide), after decades of wide- viously thought, whereas the reverse seems to be true
spread use in the USA, is increasingly associated with for offspring adiposity. These findings raise the possi-
LGA infants and neonatal hypoglycaemia237. Metformin, bility that targeted interventions based on an improved
despite encouraging immediate pregnancy out- understanding of the underlying pathophysiology of
comes, is associated with greater childhood size and GDM might eventually be able to improve short-​term
adiposity207,238,239, raising questions regarding its wide- and long-​term outcomes in offspring.
spread use. Although oral agents are easier to admin- In conclusion, although much is known about GDM,
ister and are preferred by women with GDM, these evidence gaps and evidence–practice gaps persist at
recent experiences highlight the need for caution and all points in translating molecular understanding of
long-term follow-up before major changes in guidelines GDM to efficacious therapies. A coordinated, well-​
are made. implemented focus on the health of a mother with GDM
The complex interplay between obesity per se, and the health and development of her baby across the
metabolic inflammation and hyperglycaemia (partly life cycle offers the promise of substantial health gains240
driven by obesity), as contributing causes of excess but requires enhanced knowledge at the basic, clinical
fetal growth, other adverse pregnancy outcomes and and implementation science levels and a commitment
later effects in offspring, remains incompletely under- to action on the part of policy-​makers and clinicians.
stood and deserves further attention57. Publications
from the HAPO-​FUS116,128,129 of long-​term follow-​up Published online xx xx xxxx

1. Carrington, E. R., Shuman, C. R. & Reardon, H. S. and risk for gestational diabetes mellitus. Am. J. 31. Bao, W., Tobias, D. K., Hu, F. B., Chavarro, J. E.
Evaluation of the prediabetic state during pregnancy. Obstet. Gynecol. 205, 55.e1–55.e7 (2011). & Zhang, C. Pre-​pregnancy potato consumption and
Obstet. Gynecol. 9, 664–669 (1957). 16. Solomon, C. G. et al. A prospective study of pregravid risk of gestational diabetes mellitus: prospective
2. O’Sullivan, J. B. Gestational diabetes. Unsuspected, determinants of gestational diabetes mellitus. JAMA cohort study. BMJ 352, h6898 (2016).
asymptomatic diabetes in pregnancy. N. Engl. J. Med. 278, 1078–1083 (1997). 32. Bao, W., Tobias, D. K., Olsen, S. F. & Zhang, C.
264, 1082–1085 (1961). 17. Retnakaran, R. et al. Fetal sex and maternal risk of Pre-pregnancy fried food consumption and the
3. O’Sullivan, J. B. & Mahan, C. M. Criteria for the oral gestational diabetes mellitus: the impact of having risk of gestational diabetes mellitus: a prospective
glucose tolerance test in pregnancy. Diabetes 13, a boy. Diabetes Care 38, 844–851 (2015). cohort study. Diabetologia 57, 2485–2491 (2014).
278–285 (1964). 18. Rauh-​Hain, J. A. et al. Risk for developing gestational 33. Bowers, K. et al. A prospective study of prepregnancy
4. Hadden, D. R. Prediabetes and the big baby. diabetes in women with twin pregnancies. J. Matern. dietary iron intake and risk for gestational diabetes
Diabet Med. 25, 1–10 (2008). Fetal Neonatal Med. 22, 293–299 (2009). mellitus. Diabetes Care 34, 1557–1563 (2011).
5. Metzger, B. E. & Coustan, D. R. Summary and 19. Morikawa, M. et al. Prevalence of hyperglycaemia 34. Bowers, K., Tobias, D. K., Yeung, E., Hu, F. B.
recommendations of the Fourth International in singleton versus twin pregnancy. Diabetes Metab. & Zhang, C. A prospective study of prepregnancy
Workshop-​Conference on Gestational Diabetes Res. Rev. 31, 198–203 (2015). dietary fat intake and risk of gestational diabetes.
Mellitus. The Organizing Committee. Diabetes Care 20. Hedderson, M. et al. Racial/ethnic disparities in the Am. J. Clin. Nutr. 95, 446–453 (2012).
21 (Suppl. 2), B161–B167 (1998). prevalence of gestational diabetes mellitus by BMI. 35. Bao, W., Bowers, K., Tobias, D. K., Hu, F. B. & Zhang, C.
6. Menke, A., Casagrande, S. & Cowie, C. C. Contributions Diabetes Care 35, 1492–1498 (2012). Prepregnancy dietary protein intake, major dietary
of A1c, fasting plasma glucose, and 2-hour plasma 21. Anna, V., van der Ploeg, H. P., Cheung, N. W., protein sources, and the risk of gestational diabetes
glucose to prediabetes prevalence: NHANES Huxley, R. R. & Bauman, A. E. Sociodemographic mellitus: a prospective cohort study. Diabetes Care 36,
2011–2014. Ann. Epidemiol. 28, 681–685 (2018). correlates of the increasing trend in prevalence of 2001–2008 (2013).
7. Metzger, B. E. et al. International association of diabetes gestational diabetes mellitus in a large population 36. Bao, W. et al. Prepregnancy low-​carbohydrate dietary
and pregnancy study groups recommendations on the of women between 1995 and 2005. Diabetes Care pattern and risk of gestational diabetes mellitus:
diagnosis and classification of hyperglycemia in 31, 2288–2293 (2008). a prospective cohort study. Am. J. Clin. Nutr. 99,
pregnancy. Diabetes Care 33, 676–682 (2010). 22. Zhang, C. & Ning, Y. Effect of dietary and lifestyle 1378–1384 (2014).
8. World Health Organization. Diagnostic Criteria and factors on the risk of gestational diabetes: review 37. Zhang, C., Liu, S., Solomon, C. G. & Hu, F. B. Dietary
Classification of Hyperglycaemia First Detected in of epidemiologic evidence. Am. J. Clin. Nutr. 94, fiber intake, dietary glycemic load, and the risk for
Pregnancy (WHO Press, Geneva, 2013). 1975S–1979S (2011). gestational diabetes mellitus. Diabetes Care 29,
This paper provides the rationale for, and current 23. Bao, W. et al. Parental smoking during pregnancy 2223–2230 (2006).
basis of, diagnostic recommendations for GDM on and the risk of gestational diabetes in the daughter. 38. Tobias, D. K. et al. Prepregnancy adherence to
a global scale. Int. J. Epidemiol. 45, 160–169 (2016). dietary patterns and lower risk of gestational
9. Omori, Y. & Jovanovic, L. Proposal for the 24. Tobias, D. K., Zhang, C., van Dam, R. M., Bowers, K. diabetes mellitus. Am. J. Clin. Nutr. 96, 289–295
reconsideration of the definition of gestational & Hu, F. B. Physical activity before and during (2012).
diabetes. Diabetes Care 28, 2592–2593 (2005). pregnancy and risk of gestational diabetes mellitus: 39. Zhang, C. et al. Adherence to healthy lifestyle and risk
10. Clausen, T. D. et al. Overweight and the metabolic a meta-​analysis. Diabetes Care 34, 223–229 (2011). of gestational diabetes mellitus: prospective cohort
syndrome in adult offspring of women with diet-​ 25. Zhang, C., Solomon, C. G., Manson, J. E. & Hu, F. B. study. BMJ 349, g5450 (2014).
treated gestational diabetes mellitus or type 1 A prospective study of pregravid physical activity 40. Zalbahar, N., Najman, J., McIntyre, H. D. & Mamun, A.
diabetes. J. Clin. Endocrinol. Metab. 94, 2464–2470 and sedentary behaviors in relation to the risk for Parental pre-​pregnancy obesity and the risk
(2009). gestational diabetes mellitus. Arch. Intern. Med. of offspring weight and body mass index change
11. Clausen, T. D. et al. High prevalence of type 2 diabetes 166, 543–548 (2006). from childhood to adulthood. Clin. Obes. 7, 206–215
and pre-​diabetes in adult offspring of women with 26. Zhang, C. et al. Maternal plasma 25-hydroxyvitamin D (2017).
gestational diabetes mellitus or type 1 diabetes: concentrations and the risk for gestational diabetes 41. Smarr, M. M. et al. Persistent organic pollutants
the role of intrauterine hyperglycemia. Diabetes Care mellitus. PLOS ONE 3, e3753 (2008). and pregnancy complications. Sci. Total Environ.
31, 340–346 (2008). 27. Zhang, C. et al. Maternal plasma ascorbic acid 551–552, 285–291 (2016).
12. Daly, B. et al. Increased risk of ischemic heart disease, (vitamin C) and risk of gestational diabetes mellitus. 42. Zhang, C. et al. A prospective study of prepregnancy
hypertension, and type 2 diabetes in women with Epidemiology 15, 597–604 (2004). serum concentrations of perfluorochemicals and the risk
previous gestational diabetes mellitus, a target group 28. Zhu, Y. et al. A prospective and longitudinal study of gestational diabetes. Fertil. Steril. 103, 184–189
in general practice for preventive interventions: of plasma phospholipid saturated fatty acid profile in (2015).
a population-​based cohort study. PLOS Med. 15, relation to cardiometabolic biomarkers and the risk 43. Hinkle, S. N. et al. A longitudinal study of depression
e1002488 (2018). of gestational diabetes. Am. J. Clin. Nutr. 107, and gestational diabetes in pregnancy and the
13. Zhu, Y. & Zhang, C. Prevalence of gestational diabetes 1017–1026 (2018). postpartum period. Diabetologia 59, 2594–2602
and risk of progression to type 2 diabetes: a global 29. Looman, M. et al. Pre-​pregnancy dietary carbohydrate (2016).
perspective. Curr. Diab. Rep. 16, 7 (2016). quantity and quality, and risk of developing 44. Zhang, C. et al. Genetic variants and the risk of
This article provides a global review of the gestational diabetes: the Australian longitudinal gestational diabetes mellitus: a systematic review.
relationship between GDM and the risk of later study on women’s health. Br. J. Nutr. 120, 435–444 Hum. Reprod. Update 19, 376–390 (2013).
maternal T2DM. (2018). 45. Kwak, S. H. et al. A genome-​wide association study
14. Harville, E. W., Viikari, J. S. & Raitakari, O. T. 30. Chen, L., Hu, F. B., Yeung, E., Willett, W. & Zhang, C. of gestational diabetes mellitus in Korean women.
Preconception cardiovascular risk factors and pregnancy Prospective study of pre-​gravid sugar-​sweetened Diabetes 61, 531–541 (2012).
outcome. Epidemiology 22, 724–730 (2011). beverage consumption and the risk of gestational 46. Ren, J. et al. Genetic variation in MTNR1B is
15. Hedderson, M. M., Darbinian, J. A., Quesenberry, C. P. diabetes mellitus. Diabetes Care 32, 2236–2241 associated with gestational diabetes mellitus and
& Ferrara, A. Pregravid cardiometabolic risk profile (2009). contributes only to the absolute level of beta cell

16 | Article citation ID: (2019) 5:47 www.nature.com/nrdp

0123456789();
Primer

compensation in Mexican Americans. Diabetologia 57, 68. Desoye, G. & Nolan, C. J. The fetal glucose steal: 92. Desoye, G. & van Poppel, M. The feto-​placental
1391–1399 (2014). an underappreciated phenomenon in diabetic dialogue and diabesity. Best Pract. Res. Clin. Obstet.
47. Ding, M. et al. Genetic variants of gestational diabetes pregnancy. Diabetologia 59, 1089–1094 (2016). Gynaecol. 29, 15–23 (2015).
mellitus: a study of 112 SNPs among 8722 women in 69. Haggarty, P., Page, K., Abramovich, D. R., Ashton, J. 93. Eder, M. et al. Sex differences in the association of
two independent populations. Diabetologia 61, & Brown, D. Long-​chain polyunsaturated fatty acid cord blood insulin with subcutaneous adipose tissue
1758–1768 (2018). transport across the perfused human placenta. Placenta in neonates. Int. J. Obes. 40, 538–542 (2016).
48. Catalano, P. M., Tyzbir, E. D. & Sims, E. A. Incidence 18, 635–642 (1997). 94. Houshmand-​Oeregaard, A. et al. DNA methylation
and significance of islet cell antibodies in women with 70. Lewis, R. M., Wadsack, C. & Desoye, G. Placental fatty and gene expression of TXNIP in adult offspring of
previous gestational diabetes. Diabetes Care 13, acid transfer. Curr. Opin. Clin. Nutr. Metab. Care 21, women with diabetes in pregnancy. PLOS ONE 12,
478–482 (1990). 78–82 (2018). e0187038 (2017).
49. Ellard, S., Bellanne-​Chantelot, C. & Hattersley, A. T. 71. Dancis, J., Jansen, V., Kayden, H. J., Schneider, H. 95. Houshmand-​Oeregaard, A. et al. Differential adipokine
Best practice guidelines for the molecular genetic & Levitz, M. Transfer across perfused human placenta. II. DNA methylation and gene expression in subcutaneous
diagnosis of maturity-​onset diabetes of the young. Free fatty acids. Pediatr. Res. 7, 192–197 (1973). adipose tissue from adult offspring of women with
Diabetologia 51, 546–553 (2008). 72. Pagan, A. et al. Materno-​fetal transfer of diabetes in pregnancy. Clin. Epigenetics 9, 37 (2017).
50. Hattersley, A. T. et al. Mutations in the glucokinase docosahexaenoic acid is impaired by gestational 96. Hjort, L. et al. Gestational diabetes and maternal
gene of the fetus result in reduced birth weight. diabetes mellitus. Am. J. Physiol. Endocrinol. Metab. obesity are associated with epigenome-​wide methylation
Nat. Genet. 19, 268–270 (1998). 305, E826–E833 (2013). changes in children. JCI Insight 3, e122572 (2018).
This landmark paper outlines the differing 73. Prieto-​Sanchez, M. T. et al. Placental MFSD2a 97. Pettitt, D. J., Knowler, W. C., Baird, H. R.
phenotypes encountered when maternal and fetal transporter is related to decreased DHA in cord blood & Bennett, P. H. Gestational diabetes: infant and
glucokinase genes are matched or mismatched. of women with treated gestational diabetes. Clin. Nutr. maternal complications of pregnancy in relation to
51. Catalano, P. M. et al. Longitudinal changes in basal 36, 513–521 (2017). third-​trimester glucose tolerance in the Pima Indians.
hepatic glucose production and suppression during 74. Sen, D. K., Kaufmann, P. & Schweikhart, G. Classification Diabetes Care 3, 458–464 (1980).
insulin infusion in normal pregnant women. Am. J. of human placental villi. II. Morphometry. Cell Tissue Res. 98. Beischer, N. A., Wein, P., Sheedy, M. T. & Steffen, B.
Obstet. Gynecol. 167, 913–919 (1992). 200, 425–434 (1979). Identification and treatment of women with
52. Catalano, P. M., Tyzbir, E. D., Roman, N. M., Amini, S. B. 75. Gauster, M., Desoye, G., Totsch, M. & Hiden, U. hyperglycaemia diagnosed during pregnancy can
& Sims, E. A. Longitudinal changes in insulin release The placenta and gestational diabetes mellitus. significantly reduce perinatal mortality rates. Aust. N.
and insulin resistance in nonobese pregnant women. Curr. Diab. Rep. 12, 16–23 (2012). Z. J. Obstet. Gynaecol. 36, 239–247 (1996).
Am. J. Obstet. Gynecol. 165, 1667–1672 (1991). 76. Cvitic, S., Desoye, G. & Hiden, U. Glucose, insulin, and 99. Casey, B. M., Lucas, M. J., McIntire, D. D. & Leveno, K. J.
53. Catalano, P. M., Huston, L., Amini, S. B. & Kalhan, S. C. oxygen interplay in placental hypervascularisation in Pregnancy outcomes in women with gestational
Longitudinal changes in glucose metabolism during diabetes mellitus. Biomed. Res. Int. 2014, 145846 diabetes compared with the general obstetric
pregnancy in obese women with normal glucose (2014). population. Obstet. Gynecol. 90, 869–873 (1997).
tolerance and gestational diabetes mellitus. Am. J. 77. Lassance, L. et al. Hyperinsulinemia stimulates 100. Aberg, A., Rydhstrom, H., Kallen, B. & Kallen, K. Impaired
Obstet. Gynecol. 180, 903–916 (1999). angiogenesis of human fetoplacental endothelial cells: glucose tolerance during pregnancy is associated with
This paper provides detailed insight into the a possible role of insulin in placental hypervascularization increased fetal mortality in preceding sibs. Acta Obstet.
pathophysiology of GDM as compared with in diabetes mellitus. J. Clin. Endocrinol. Metab. 98, Gynecol. Scand. 76, 212–217 (1997).
and contrasted to obesity without diabetes in a E1438–E1447 (2013). 101. Persson, B. & Hanson, U. Neonatal morbidities in
longitudinal cohort of women monitored over the 78. Loegl, J. et al. Hofbauer cells of M2a, M2b and M2c gestational diabetes mellitus. Diabetes Care
course of pregnancy. polarization may regulate feto-​placental angiogenesis. 21(Suppl. 2), B79–B84 (1998).
54. Buchanan, T. A. Pancreatic B cell defects in gestational Reproduction 152, 447–455 (2016). 102. Jensen, D. M. et al. Proposed diagnostic thresholds
diabetes: implications for the pathogenesis and 79. Loegl, J. et al. Pigment epithelium-​derived factor for gestational diabetes mellitus according to a 75-g
prevention of type 2 diabetes. J. Clin. Endocrinol. (PEDF): a novel trophoblast-​derived factor limiting oral glucose tolerance test. Maternal and perinatal
Metab. 86, 989–993 (2001). feto-​placental angiogenesis in late pregnancy. outcomes in 3260 Danish women. Diabet Med. 20,
55. DeFronzo, R. A. et al. Type 2 diabetes mellitus. Angiogenesis 19, 373–388 (2016). 51–57 (2003).
Nat. Rev. Dis. Primers 1, 15019 (2015). 80. Sun, Y. et al. Gestational diabetes mellitus modulates 103. Sermer, M. et al. Impact of increasing carbohydrate
56. Friedman, J. E. et al. Impaired glucose transport and cholesterol homeostasis in human fetoplacental intolerance on maternal-​fetal outcomes in 3637
insulin receptor tyrosine phosphorylation in skeletal endothelium. Biochim. Biophys. Acta Mol. Cell Biol. women without gestational diabetes. The Toronto
muscle from obese women with gestational diabetes. Lipids 1863, 968–979 (2018). Tri-Hospital Gestational Diabetes Project. Am. J.
Diabetes 48, 1807–1814 (1999). 81. Scholler, M. et al. Phospholipid transfer protein is Obstet. Gynecol. 173, 146–156 (1995).
57. McIntyre, Discovery H. D. knowledge, and action-​ differentially expressed in human arterial and venous 104. Sacks, D. A. et al. Toward universal criteria for
diabetes in pregnancy across the translational placental endothelial cells and enhances cholesterol gestational diabetes: the 75-gram glucose tolerance
spectrum: the 2016 Norbert Freinkel Award Lecture. efflux to fetal HDL. J. Clin. Endocrinol. Metab. 97, test in pregnancy. Am. J. Obstet. Gynecol. 172,
Diabetes Care 41, 227–232 (2018). 2466–2474 (2012). 607–614 (1995).
58. Hotamisligil, G. S. et al. IRS-1-mediated inhibition of 82. Desoye, G. The human placenta in diabetes and 105. Moses, R. G. & Calvert, D. Pregnancy outcomes in
insulin receptor tyrosine kinase activity in TNF-​alpha- obesity: friend or foe? The 2017 Norbert Freinkel women without gestational diabetes mellitus related
and obesity-​induced insulin resistance. Science 271, Award Lecture. Diabetes Care 41, 1362–1369 (2018). to the maternal glucose level. Is there a continuum of
665–668 (1996). This article provides a comprehensive overview risk? Diabetes Care 18, 1527–1533 (1995).
59. Kirwan, J. P. et al. Reversal of insulin resistance of placental structure and function in diabetic 106. Jensen, D. M. et al. Clinical impact of mild
postpartum is linked to enhanced skeletal muscle pregnancy. carbohydrate intolerance in pregnancy: a study of
insulin signaling. J. Clin. Endocrinol. Metab. 89, 83. Gauster, M. et al. Dysregulation of placental endothelial 2904 nondiabetic Danish women with risk factors for
4678–4684 (2004). lipase in obese women with gestational diabetes gestational diabetes mellitus. Am. J. Obstet. Gynecol.
60. Waters, T. P. et al. Does maternal insulin sensitivity mellitus. Diabetes 60, 2457–2464 (2011). 185, 413–419 (2001).
improve immediately after delivery or do we need 84. Cvitic, S. et al. The human placental sexome differs 107. Jarrett, R. J. Gestational diabetes: a non-​entity? BMJ
to wait unti six weeks postpartum? Am. J. Obstet. between trophoblast epithelium and villous vessel 306, 37–38 (1993).
Gynecol. 212, S20 (2015). endothelium. PLOS ONE 8, e79233 (2013). 108. Metzger, B. E. et al. Hyperglycemia and adverse
61. Friedman, J. E., Kirwan, J. P., Jing, M., Presley, L. 85. Knabl, J. et al. GDM alters expression of placental pregnancy outcomes. N. Engl. J. Med. 358,
& Catalano, P. M. Increased skeletal muscle tumor estrogen receptor alpha in a cell type and 1991–2002 (2008).
necrosis factor-​alpha and impaired insulin signaling gender-specific manner. Reprod. Sci. 22, 1488–1495 This landmark epidemiological study defines, on a
persist in obese women with gestational diabetes (2015). worldwide basis, the relationship between impaired
mellitus 1 year postpartum. Diabetes 57, 606–613 86. Sedlmeier, E. M. et al. Human placental transcriptome glucose metabolism that is less severe than overt
(2008). shows sexually dimorphic gene expression and diabetes and pregnancy complications.
62. Berggren, E. K., Presley, L., Amini, S. B., responsiveness to maternal dietary n-3 long-​chain 109. Crowther, C. A. et al. Effect of treatment of gestational
Hauguel-de Mouzon, S. & Catalano, P. M. Are the polyunsaturated fatty acid intervention during diabetes mellitus on pregnancy outcomes. N. Engl. J.
metabolic changes of pregnancy reversible in the first pregnancy. BMC Genomics 15, 941 (2014). Med. 352, 2477–2486 (2005).
year postpartum? Diabetologia 58, 1561–1568 (2015). 87. Strutz, J. et al. Gestational diabetes alters microRNA This landmark randomized controlled trial
63. Peters, R. K., Kjos, S. L., Xiang, A. & Buchanan, T. A. signatures in human feto-​placental endothelial cells demonstrates that treatment of GDM is effective
Long-​term diabetogenic effect of single pregnancy in depending on fetal sex. Clin. Sci. 132, 2437–2449 in reducing immediate pregnancy complications,
women with previous gestational diabetes mellitus. (2018). including excess fetal growth and pregnancy-​related
Lancet 347, 227–230 (1996). 88. Desoye, G. & Hauguel-​de Mouzon, S. The human hypertension.
64. Lewis, R. M. et al. The placental exposome: placental placenta in gestational diabetes mellitus. The insulin 110. Landon, M. B. et al. A multicenter, randomized trial of
determinants of fetal adiposity and postnatal body and cytokine network. Diabetes Care 30 (Suppl. 2), treatment for mild gestational diabetes. N. Engl. J. Med.
composition. Ann. Nutr. Metab. 63, 208–215 (2013). S120–S126 (2007). 361, 1339–1348 (2009).
65. Desoye, G. & Shafrir, E. Placental metabolism and its 89. Watson, A. L., Skepper, J. N., Jauniaux, E. This USA-​based randomized controlled trial
regulation in health and diabetes. Mol. Aspects Med. & Burton, G. J. Changes in concentration, localization confirms that treatment of GDM reduces excess
15, 505–682 (1994). and activity of catalase within the human placenta fetal growth, excess fetal adiposity and pregnancy-​
66. Hauguel, S., Desmaizieres, V. & Challier, J. C. Glucose during early gestation. Placenta 19, 27–34 (1998). related hypertension.
uptake, utilization, and transfer by the human placenta 90. Lappas, M. et al. The role of oxidative stress in the 111. Bellamy, L., Casas, J. P., Hingorani, A. D. & Williams, D.
as functions of maternal glucose concentration. pathophysiology of gestational diabetes mellitus. Type 2 diabetes mellitus after gestational diabetes:
Pediatr. Res. 20, 269–273 (1986). Antioxid. Redox Signal. 15, 3061–3100 (2011). a systematic review and meta-​analysis. Lancet 373,
67. Osmond, D. T., Nolan, C. J., King, R. G., Brennecke, S. P. 91. Hoch, D., Gauster, M., Hauguel-​de Mouzon, S. & 1773–1779 (2009).
& Gude, N. M. Effects of gestational diabetes on human Desoye, G. Diabesity-​associated oxidative and 112. Cheung, N. W. & Byth, K. Population health significance
placental glucose uptake, transfer, and utilisation. inflammatory stress signalling in the early human of gestational diabetes. Diabetes Care 26, 2005–2009
Diabetologia 43, 576–582 (2000). placenta. Mol. Aspects Med. 66, 21–30 (2018). (2003).

NATURE REvIEWS | DiSeASe PRiMeRS | Article citation I­D:­ (2019) 5:47 ­ 17

0123456789();
Primer

113. Lauenborg, J. et al. Increasing incidence of diabetes 137. Zhang, C., Rawal, S. & Chong, Y. S. Risk factors for 159. Tang, T., Lord, J. M., Norman, R. J., Yasmin, E. &
after gestational diabetes: a long-​term follow-​up in a gestational diabetes: is prevention possible? Balen, A. H. Insulin-​sensitising drugs (metformin,
Danish population. Diabetes Care 27, 1194–1199 Diabetologia 59, 1385–1390 (2016). rosiglitazone, pioglitazone, D-​chiro-inositol) for women
(2004). 138. Song, C. et al. Long-​term risk of diabetes in women with polycystic ovary syndrome, oligo amenorrhoea
114. Damm, P. Future risk of diabetes in mother and child at varying durations after gestational diabetes: and subfertility. Cochrane Database Syst. Rev. 5,
after gestational diabetes mellitus. Int. J. Gynaecol. a systematic review and meta-​analysis with more than 2 CD003053 (2012).
Obstet. 104 (Suppl. 1), S25–S26 (2009). million women. Obes. Rev. 19, 421–429 (2018). 160. Carpenter, M. W. & Coustan, D. R. Criteria for screening
115. Damm, P., Kuhl, C., Bertelsen, A. & Molsted-​Pedersen, L. 139. Russo, L. M., Nobles, C., Ertel, K. A., Chasan-​Taber, L. tests for gestational diabetes. Am. J. Obstet. Gynecol.
Predictive factors for the development of diabetes in & Whitcomb, B. W. Physical activity interventions in 144, 768–773 (1982).
women with previous gestational diabetes mellitus. pregnancy and risk of gestational diabetes mellitus: 161. Berggren, E. K., Boggess, K. A., Stuebe, A. M. &
Am. J. Obstet. Gynecol. 167, 607–616 (1992). a systematic review and meta-​analysis. Obstet. Gynecol. Jonsson Funk, M. National Diabetes Data Group
116. Lowe, W. L. Jr. et al. Association of gestational diabetes 125, 576–582 (2015). versus Carpenter-​Coustan criteria to diagnose
with maternal disorders of glucose metabolism and 140. Sanabria-​Martinez, G. et al. Effectiveness of physical gestational diabetes. Am. J. Obstet. Gynecol. 205,
childhood adiposity. JAMA 320, 1005–1016 (2018). activity interventions on preventing gestational 253.e1–253.e7 (2011).
This follow-​up study from the landmark HAPO diabetes mellitus and excessive maternal weight gain: 162. ACOG Committee on Practice Bulletins. Practice
study demonstrates that GDM diagnosed using the a meta-​analysis. BJOG 122, 1167–1174 (2015). Bulletin No. 137: gestational diabetes mellitus.
IADPSG criteria is associated with long-​term risk of 141. Song, C., Li, J., Leng, J., Ma, R. C. & Yang, X. Lifestyle Obstet. Gynecol. 122, 406–416 (2013).
T2DM in the mother and with risk of obesity in intervention can reduce the risk of gestational diabetes: 163. Metzger, B. E. et al. The hyperglycemia and adverse
offspring (WHO 2013). a meta-​analysis of randomized controlled trials. pregnancy outcome (HAPO) Study. Int. J. Gynaecol.
117. Lauenborg, J. et al. The prevalence of the metabolic Obes. Rev. 17, 960–969 (2016). Obstet. 78, 69–77 (2002).
syndrome in a Danish population of women with 142. Han, S., Crowther, C. A., Middleton, P. & Heatley, E. 164. Metzger, B. E. et al. Hyperglycemia and adverse
previous gestational diabetes mellitus is three-​fold Different types of dietary advice for women with pregnancy outcome study: neonatal glycemia. Pediatrics
higher than in the general population. J. Clin. gestational diabetes mellitus. Cochrane Database 126, e1545–e1552 (2010).
Endocrinol. Metab. 90, 4004–4010 (2005). Syst. Rev. 3, CD009275 (2013). 165. Sacks, D. A. et al. Frequency of gestational diabetes
118. Retnakaran, R. Hyperglycemia in pregnancy and its 143. Bain, E. et al. Diet and exercise interventions for mellitus at collaborating centers based on IADPSG
implications for a woman’s future risk of cardiovascular preventing gestational diabetes mellitus. Cochrane consensus panel-​recommended criteria: the
disease. Diabetes Res. Clin. Pract. 145, 193–199 Database Syst. Rev. 4, CD010443 (2015). hyperglycemia and adverse pregnancy outcome
(2018). 144. Celentano, C. et al. Myo-​Inositol supplementation to (HAPO) study. Diabetes Care. 35, 526–528 (2012).
119. Retnakaran, R. & Shah, B. R. Role of type 2 diabetes prevent gestational diabetes mellitus. Curr. Diab Rep. 166. Diabetes, I. Ao. et al. International association of
in determining retinal, renal, and cardiovascular 16, 30 (2016). diabetes and pregnancy study groups recommendations
outcomes in women with previous gestational diabetes 145. Santamaria, A. et al. Myo-​inositol may prevent on the diagnosis and classification of hyperglycemia in
mellitus. Diabetes Care 40, 101–108 (2017). gestational diabetes onset in overweight women: pregnancy. Diabetes Care 33, 676–682 (2010).
120. Kramer, C. K., Campbell, S. & Retnakaran, R. a randomized, controlled trial. J. Matern. Fetal 167. American Diabetes Association. Classification
Gestational diabetes and the risk of cardiovascular Neonatal Med. 29, 3234–3237 (2016). and diagnosis of diabetes: standards of medical care in
disease in women: a systematic review and meta-​ 146. Farren, M. et al. The prevention of gestational diabetes diabetes-2019. Diabetes Care 42, S13–S28 (2019).
analysis. Diabetologia 62, 905–914 (2019). mellitus with antenatal oral inositol supplementation: 168. Blumer, I. et al. Diabetes and pregnancy: an endocrine
121. Harder, T. et al. Pancreatic islet transplantation in a randomized controlled trial. Diabetes Care 40, society clinical practice guideline. J. Clin. Endocrinol.
diabetic pregnant rats prevents acquired malformation 759–763 (2017). Metab. 98, 4227–4249 (2013).
of the ventromedial hypothalamic nucleus in their 147. Crawford, T. J., Crowther, C. A., Alsweiler, J. & Brown, J. 169. Nankervis, A., McIntyre, H. D., Moses, R. G., Ross, G. P.
offspring. Neurosci. Lett. 299, 85–88 (2001). Antenatal dietary supplementation with myo-​inositol in & Callaway, L. K. Testing for gestational diabetes
122. Aerts, L. & Van Assche, F. A. Animal evidence for the women during pregnancy for preventing gestational mellitus in Australia. Diabetes Care 36, e64 (2013).
transgenerational development of diabetes mellitus. diabetes. Cochrane Database Syst. Rev. 12, CD011507 170. Ikenoue, S. et al. Clinical impact of women with
Int. J. Biochem. Cell Biol. 38, 894–903 (2006). (2015). gestational diabetes mellitus by the new consensus
123. Silverman, B. L. et al. Long-​term prospective 148. Luoto, R., Laitinen, K., Nermes, M. & Isolauri, E. criteria: two year experience in a single institution in
evaluation of offspring of diabetic mothers. Diabetes Impact of maternal probiotic-​supplemented dietary Japan. Endocr. J. 61, 353–358 (2014).
40 (Suppl. 2), 121–125 (1991). counselling on pregnancy outcome and prenatal and 171. Kautzky-​Willer, A. et al. Gestational diabetes mellitus
124. Dabelea, D. et al. Intrauterine exposure to diabetes postnatal growth: a double-​blind, placebo-​controlled [German]. Wien. Klin. Wochenschr. 124 (Suppl. 2),
conveys risks for type 2 diabetes and obesity: a study study. Br. J. Nutr. 103, 1792–1799 (2010). 58–65 (2012).
of discordant sibships. Diabetes 49, 2208–2211 (2000). 149. Callaway, L. K. et al. Probiotics for the prevention of 172. Hod, M. et al. The International Federation of
125. Fraser, A. & Lawlor, D. A. Long-​term health outcomes gestational diabetes mellitus in overweight and obese Gynecology and Obstetrics (FIGO) Initiative on
in offspring born to women with diabetes in pregnancy. women: findings from the SPRING double-​blind gestational diabetes mellitus: a pragmatic guide for
Curr. Diab. Rep. 14, 489 (2014). randomized controlled trial. Diabetes Care 42, diagnosis, management, and care. Int. J. Gynaecol.
126. Kelstrup, L. et al. Insulin resistance and impaired 364–371 (2019). Obstet. 131, S173–S211 (2015).
pancreatic beta-​cell function in adult offspring of 150. Lindsay, K. L. et al. Probiotics in obese pregnancy do 173. Diehl, K. et al. German gynecologists’ experience with a
women with diabetes in pregnancy. J. Clin. Endocrinol. not reduce maternal fasting glucose: a double-​blind, universal screening for gestational diabetes mellitus in
Metab. 98, 3793–3801 (2013). placebo-​controlled, randomized trial (probiotics in daily practice: a qualitative study. J. Turk. Ger. Gynecol.
127. Grunnet, L. G. et al. Adiposity, dysmetabolic traits, and pregnancy study). Am. J. Clin. Nutr. 99, 1432–1439 Assoc. 17, 10–15 (2016).
earlier onset of female puberty in adolescent offspring (2014). 174. Rowan, J. et al. New Zealand National GDM Guidelines:
of women with gestational diabetes mellitus: a clinical 151. Lindsay, K. L., Walsh, C. A., Brennan, L. an alternative view of some good practice points.
study within the Danish National Birth Cohort. & McAuliffe, F. M. Probiotics in pregnancy and Aust. N. Z. J. Obstet. Gynaecol. 55, 17–20 (2015).
Diabetes Care 40, 1746–1755 (2017). maternal outcomes: a systematic review. J. Matern. 175. Benhalima, K. et al. A proposal for the use of uniform
128. Lowe, W. L. Jr. et al. Hyperglycemia and adverse Fetal Neonatal Med. 26, 772–778 (2013). diagnostic criteria for gestational diabetes in Europe:
pregnancy outcome follow-​up study (HAPO FUS): 152. Pellonpera, O. et al. Efficacy of fish oil and/or probiotic an opinion paper by the European Board & College of
maternal gestational diabetes and childhood glucose intervention on the incidence of gestational diabetes Obstetrics and Gynaecology (EBCOG). Diabetologia
metabolism. Diabetes Care 42, 372–380 (2019). mellitus in an at-​risk group of overweight and obese 58, 1422–1429 (2015).
129. Lowe, W. L. Jr. et al. Maternal glucose levels during women: a randomized, placebo-​controlled, double-​blind 176. Sadikot, S., Purandare, C. N., Cho, N. H. & Hod, M.
pregnancy and childhood adiposity in the Hyperglycemia clinical trial. Diabetes Care 42, 1009–1017 (2019). FIGO-​IDF joint statement and declaration on
and Adverse Pregnancy Outcome Follow-​up Study. 153. Jarde, A. et al. Pregnancy outcomes in women taking hyperglycemia in pregnancy. Diabetes Res. Clin. Pract.
Diabetologia 62, 598–610 (2019). probiotics or prebiotics: a systematic review and meta-​ 145, 1–4 (2018).
130. Clausen, T. D. et al. Cognitive function in adult analysis. BMC Pregnancy Childbirth 18, 14 (2018). 177. Committee on Practice Bulletins—Obstetrics. ACOG
offspring of women with gestational diabetes—the role 154. Zheng, J., Feng, Q., Zheng, S. & Xiao, X. The effects Practice Bulletin No. 190: gestational diabetes
of glucose and other factors. PLOS ONE 8, e67107 of probiotics supplementation on metabolic health in mellitus. Obstet. Gynecol. 131, e49–e64 (2018).
(2013). pregnant women: an evidence based meta-​analysis. 178. Feig, D. S. et al. Diabetes and pregnancy. Can. J.
131. Xiang, A. H. et al. Association of maternal diabetes with PLOS ONE 13, e0197771 (2018). Diabetes 42 (Suppl. 1), 255–282 (2018).
autism in offspring. Jama 313, 1425–1434 (2015). 155. Barrett, H. L., Dekker Nitert, M., Conwell, L. S. & 179. National Institute for Health and Care Excellence.
132. Xiang, A. H. et al. Maternal gestational diabetes Callaway, L. K. Probiotics for preventing gestational Diabetes in Pregnancy: Management from
mellitus, type 1 diabetes, and type 2 diabetes during diabetes. Cochrane Database Syst. Rev. 2, CD009951 Preconception to the Postnatal Period (NICE, 2015).
pregnancy and risk of ADHD in offspring. Diabetes Care (2014). 180. van Leeuwen, M. et al. Glucose challenge test for
41, 2502–2508 (2018). 156. Chiswick, C. et al. Effect of metformin on maternal detecting gestational diabetes mellitus: a systematic
133. Gillman, M. W. et al. Effect of treatment of gestational and fetal outcomes in obese pregnant women review. BJOG 119, 393–401 (2012).
diabetes mellitus on obesity in the next generation. (EMPOWaR): a randomised, double-​blind, placebo-​ 181. Saccone, G., Khalifeh, A., Al-​Kouatly, H. B., Sendek, K. &
Diabetes Care 33, 964–968 (2010). controlled trial. Lancet Diabetes Endocrinol. 3, Berghella, V. Screening for gestational diabetes mellitus:
134. Landon, M. B. et al. Mild gestational diabetes mellitus 778–786 (2015). one step versus two step approach. A meta-analysis of
and long-​term child health. Diabetes Care 38, 445–452 157. Dodd, J. M. et al. Effect of metformin in addition to randomized trials. J. Matern. Fetal Neonatal Med.
(2015). dietary and lifestyle advice for pregnant women who https://doi.org/10.1080/14767058.2018.1519543
135. Luoto, R. et al. Primary prevention of gestational are overweight or obese: the GRoW randomised, (2018).
diabetes mellitus and large-​for-gestational-​age double-​blind, placebo-​controlled trial. Lancet Diabetes 182. Sievenpiper, J. L., McDonald, S. D., Grey, V. &
newborns by lifestyle counseling: a cluster-​randomized Endocrinol. 7, 15–24 (2019). Don-Wauchope, A. C. Missed follow-​up opportunities
controlled trial. PLOS Med. 8, e1001036 (2011). 158. Glueck, C. J., Pranikoff, J., Aregawi, D. & Wang, P. using a two-​step screening approach for gestational
136. Dodd, J. M. et al. Antenatal lifestyle advice for women Prevention of gestational diabetes by metformin plus diabetes. Diabetes Res. Clin. Pract. 96, e43–e46 (2012).
who are overweight or obese: LIMIT randomised trial. diet in patients with polycystic ovary syndrome. 183. Donovan, L. E., Savu, A., Edwards, A. L., Johnson, J. A.
BMJ 348, g1285 (2014). Fertil. Steril. 89, 625–634 (2008). & Kaul, P. Prevalence and timing of screening and

18 | Article citation ID: (2019) 5:47 www.nature.com/nrdp

0123456789();
Primer

diagnostic testing for gestational diabetes mellitus: 207. Rowan, J. A. et al. Metformin in gestational diabetes: 228. Kim, C. Maternal outcomes and follow-​up after
a population-​based study in Alberta, Canada. the offspring follow-​up (MiG TOFU): body composition gestational diabetes mellitus. Diabet. Med. 31,
Diabetes Care 39, 55–60 (2016). and metabolic outcomes at 7–9 years of age. 292–301 (2014).
184. McIntyre, H. D. Diagnosing gestational diabetes BMJ Open Diabetes Res. Care 6, e000456 (2018). 229. Wexler, D. J. et al. Research gaps in gestational
mellitus: rationed or rationally related to risk? 208. Ijas, H., Vaarasmaki, M., Saarela, T., Keravuo, R. diabetes mellitus: executive summary of a National
Diabetes Care 36, 2879–2880 (2013). & Raudaskoski, T. A follow-​up of a randomised study Institute of Diabetes and Digestive and Kidney Diseases
185. McIntyre, H. D. et al. Gestational diabetes mellitus: of metformin and insulin in gestational diabetes Workshop. Obstet. Gynecol. 132, 496–505 (2018).
does one size fit all? A challenge to uniform worldwide mellitus: growth and development of the children at This paper provides a summary of the areas
diagnostic thresholds. Diabetes Care 41, 1339–1342 the age of 18 months. BJOG 122, 994–1000 (2015). requiring further research in GDM, which provides
(2018). 209. Langer, O., Conway, D. L., Berkus, M. D., Xenakis, E. M. an outline of current evidence gaps and a blueprint
186. American Diabetes Association. 14. Management of & Gonzales, O. A comparison of glyburide and insulin for future research priorities.
diabetes in pregnancy: standards of medical care in women with gestational diabetes mellitus. N. Engl. 230. Poon, L. C., McIntyre, H. D., Hyett, J. A.,
in diabetes — 2019. Diabetes Care 42, S165–S172 J. Med. 343, 1134–1138 (2000). da Fonseca, E. B. & Hod, M. The first-​trimester of
(2019). 210. Schwartz, R. A., Rosenn, B., Aleksa, K. & Koren, G. pregnancy – a window of opportunity for prediction
187. Gunderson, E. P. Gestational diabetes and nutritional Glyburide transport across the human placenta. and prevention of pregnancy complications and future
recommendations. Curr. Diab. Rep. 4, 377–386 (2004). Obstet. Gynecol. 125, 583–588 (2015). life. Diabetes Res. Clin. Pract. 145, 20–30 (2018).
188. Viana, L. V., Gross, J. L. & Azevedo, M. J. Dietary 211. Hebert, M. F. et al. Are we optimizing gestational 231. Ghosh, P. et al. Plasma glycated CD59, a novel
intervention in patients with gestational diabetes diabetes treatment with glyburide? The pharmacologic biomarker for detection of pregnancy-​induced glucose
mellitus: a systematic review and meta-​analysis of basis for better clinical practice. Clin. Pharmacol. Ther. intolerance. Diabetes Care 40, 981–984 (2017).
randomized clinical trials on maternal and newborn 85, 607–614 (2009). 232. Bahado-​Singh, R. O. et al. Fetal male gender and the
outcomes. Diabetes Care 37, 3345–3355 (2014). 212. Camelo Castillo, W. et al. Association of adverse benefits of treatment of mild gestational diabetes
189. Rasmussen, K. M., Catalano, P. M. & Yaktine, A. L. pregnancy outcomes with glyburide versus insulin in mellitus. Am. J. Obstet. Gynecol. 206, 422.e1–422.e5
New guidelines for weight gain during pregnancy: women with gestational diabetes. JAMA Pediatr. 169, (2012).
what obstetrician/gynecologists should know. 452–458 (2015). 233. van Poppel, M. N., Eder, M., Lang, U. & Desoye, G.
Curr. Opin. Obstet. Gynecol. 21, 521–526 (2009). 213. Tobias, D. K. et al. Healthful dietary patterns and type 2 Sex-​specific associations of insulin-​like peptides in
190. Rizzo, T., Metzger, B. E., Burns, W. J. & Burns, K. diabetes mellitus risk among women with a history of cord blood with size at birth. Clin. Endocrinol. 89,
Correlations between antepartum maternal metabolism gestational diabetes mellitus. Arch. Intern. Med. 172, 187–193 (2018).
and intelligence of offspring. N. Engl. J. Med. 325, 1566–1572 (2012). 234. Guzman-​Gutierrez, E. et al. Role of insulin and
911–916 (1991). 214. Bao, W. et al. Physical activity and sedentary behaviors adenosine in the human placenta microvascular and
191. American Diabetes Association. 13. Management associated with risk of progression from gestational macrovascular endothelial cell dysfunction in gestational
of diabetes in pregnancy: standards of medical care diabetes mellitus to type 2 diabetes mellitus: a diabetes mellitus. Microcirculation 21, 26–37 (2014).
in diabetes — 2018. Diabetes Care 41, S137–S143 prospective cohort study. JAMA Intern. Med. 174, 235. Li, J. et al. The impact of umbilical blood flow regulation
(2018). 1047–1055 (2014). on fetal development differs in diabetic and non-​diabetic
192. Ruchat, S. M. et al. Effectiveness of exercise 215. Bao, W. et al. Long-​term risk of type 2 diabetes mellitus pregnancy. Kidney Blood Press. Res. 39, 369–377
interventions in the prevention of excessive gestational in relation to BMI and weight change among women (2014).
weight gain and postpartum weight retention: with a history of gestational diabetes mellitus: 236. Catalano, P. & deMouzon, S. H. Maternal obesity
a systematic review and meta-​analysis. Br. J. Sports Med. a prospective cohort study. Diabetologia 58, and metabolic risk to the offspring: why lifestyle
52, 1347–1356 (2018). 1212–1219 (2015). interventions may have not achieved the desired
193. Vaarasmaki, M. Is it worth treating gestational diabetes: 216. Gunderson, E. P. et al. Lactation and progression to outcomes. Int. J. Obes. 39, 642–649 (2015).
if so, when and how? Diabetologia 59, 1391–1395 type 2 diabetes mellitus after gestational diabetes 237. Finneran, M. M. & Landon, M. B. Oral agents for the
(2016). mellitus: a prospective cohort study. Ann. Intern. Med. treatment of gestational diabetes. Curr. Diab. Rep. 18,
194. Kurtzhals, L. L. et al. The impact of restricted gestational 163, 889–898 (2015). 119 (2018).
weight gain by dietary intervention on fetal growth in 217. Buchanan, T. A. et al. Preservation of pancreatic 238. Hanem, L. G. E. et al. Intrauterine metformin
women with gestational diabetes mellitus. Diabetologia beta-​cell function and prevention of type 2 diabetes exposure and offspring cardiometabolic risk factors
61, 2528–2538 (2018). by pharmacological treatment of insulin resistance in (PedMet study): a 5–10 year follow-​up of the
195. Kjos, S. L. & Schaefer-​Graf, U. M. Modified therapy for high-risk hispanic women. Diabetes 51, 2796–2803 PregMet randomised controlled trial. Lancet Child
gestational diabetes using high-​risk and low-​risk fetal (2002). Adolesc. Health 3, 166–174 (2019).
abdominal circumference growth to select strict versus 218. Aroda, V. R. et al. The effect of lifestyle intervention 239. Hanem, L. G. E. et al. Metformin use in PCOS
relaxed maternal glycemic targets. Diabetes Care 30 and metformin on preventing or delaying diabetes pregnancies increases the risk of offspring overweight at
(Suppl. 2), 200–205 (2007). among women with and without gestational diabetes: 4 years of age: follow-​up of two RCTs. J. Clin. Endocrinol.
196. Charlton, R. A. et al. Prescribing of antidiabetic the Diabetes Prevention Program outcomes study Metab. 103, 1612–1621 (2018).
medicines before, during and after pregnancy: a study 10-year follow-​up. J. Clin. Endocrinol. Metab. 100, 240. Hod, M., Kapur, A. & McIntyre, H. D. Evidence in support
in seven European regions. PLOS ONE 11, e0155737 1646–1653 (2015). of the international association of diabetes in pregnancy
(2016). 219. Diabetes Prevention Program Research Group. Long-​ study groups’ criteria for diagnosing gesta­tional
197. Lepercq, J. et al. Meta-​analysis of maternal and term effects of metformin on diabetes prevention: diabetes worldwide in 2019. Am. J. Obstet. Gynecol.
neonatal outcomes associated with the use of insulin identification of subgroups that benefited most in the https://doi.org/10.1016/j.ajog.2019.01.206 (2019).
glargine versus NPH insulin during pregnancy. Obstet. Diabetes Prevention Program and Diabetes Prevention
Gynecol. Int. 2012, 649070 (2012). Program outcomes study. Diabetes Care 42, 601–608 Acknowledgements
198. Mikkelsen, M. R., Nielsen, S. B., Stage, E., (2019). C.Z. is supported by the intramural research program of the
Mathiesen, E. R. & Damm, P. High maternal HbA1c is 220. Wilkinson, S. A., McCray, S., Beckmann, M., Parry, A. Eunice Kennedy Shriver National Institute of Child Health and
associated with overweight in neonates. Dan. Med. Bull. & McIntyre, H. D. Barriers and enablers to translating Human Development, NIH.
58, A4309 (2011). gestational diabetes guidelines into practice.
199. Herrera, K. M. et al. Randomized controlled trial Pract. Diabetes 31, 67–72a (2014). Author contributions
of insulin detemir versus NPH for the treatment of 221. Dalfra, M. G., Nicolucci, A., Bisson, T., Bonsembiante, B. Introduction (H.D.M.); Epidemiology (C.Z.); Mechanisms/
pregnant women with diabetes. Am. J. Obstet. Gynecol. & Lapolla, A. Quality of life in pregnancy and post-​ pathophysiology (P.C. and G.D.); Diagnosis, screening
213, 426.e1–426.e7 (2015). partum: a study in diabetic patients. Qual. Life Res. and prevention (H.D.M., C.Z. and E.R.M.); Management
200. Jia, Y., Lao, Y., Zhu, H., Li, N. & Leung, S. W. 21, 291–298 (2012). (P.C. and E.R.M.); Quality of life (P.D.); Outlook (H.D.M., P.C.,
Is metformin still the most efficacious first-​line oral 222. Danyliv, A. et al. Health related quality of life two C.Z., G.D. and P.D.); Overview of the Primer (H.D.M.).
hypoglycaemic drug in treating type 2 diabetes? to five years after gestational diabetes mellitus:
A network meta-​analysis of randomized controlled cross-sectional comparative study in the ATLANTIC Competing interests
trials. Obes. Rev. 20, 1–12 (2019). DIP cohort. BMC Pregnancy Childbirth 15, 274 P.D. is participating in multi-​centre and multinational clinical
201. Rowan, J. A., Hague, W. M., Gao, W., Battin, M. R. & (2015). studies on the use of insulin in pregnant women with pre-​
Moore, M. P. Metformin versus insulin for the treatment 223. Marchetti, D., Carrozzino, D., Fraticelli, F., Fulcheri, M. existing diabetes in collaboration with Novo Nordisk; no per-
of gestational diabetes. N. Engl. J. Med. 358, & Vitacolonna, E. Quality of life in women with sonal honorarium is involved. E.R.M. is participating in
2003–2015 (2008). gestational diabetes mellitus: a systematic review. multinational clinical studies on the use of insulin in pregnant
202. Ijas, H. et al. Metformin should be considered in the J. Diabetes Res. 2017, 7058082 (2017). women with pre-​existing diabetes in collaboration with Novo
treatment of gestational diabetes: a prospective 224. Pace, R., Rahme, E., Da Costa, D. & Dasgupta, K. Nordisk. E.R.M. has given talks for Novo Nordisk, AstraZeneca,
randomised study. BJOG 118, 880–885 (2011). Association between gestational diabetes mellitus and Lilly and Sanofi Aventis and has received a salary for these tasks.
203. Balsells, M. et al. Glibenclamide, metformin, and insulin depression in parents: a retrospective cohort study. H.D.M., C.Z., P.C. and G.D. declare no competing interests.
for the treatment of gestational diabetes: a systematic Clin. Epidemiol. 10, 1827–1838 (2018).
review and meta-​analysis. BMJ 350, h102 (2015). 225. Clark, C. E., Rasgon, N. L., Reed, D. E. 2nd Peer review information
204. Barbour, L. A. & Feig, D. S. Metformin for gestational & Robakis, T. K. Depression precedes, but does not Nature Reviews Disease Primers thanks D. Coustan,
diabetes mellitus: progeny, perspective, and a follow, gestational diabetes. Acta Psychiatr. Scand. T. Buchanan and L. Sobrevia for their contribution to the
personalized approach. Diabetes Care 42, 396–399 139, 311–321 (2019). peer review of this work.
(2019). 226. Egan, A. M. et al. Epidemiology of gestational
205. Rowan, J. A. et al. Metformin in gestational diabetes: diabetes mellitus according to IADPSG/WHO 2013 Publisher’s note
the offspring follow-​up (MiG TOFU): body composition at criteria among obese pregnant women in Europe. Springer Nature remains neutral with regard to jurisdictional
2 years of age. Diabetes Care 34, 2279–2284 (2011). Diabetologia 60, (1913–1921 (2017). claims in published maps and institutional affiliations.
206. Battin, M. R. et al. Blood pressure measurement at 227. Goveia, P. et al. Lifestyle intervention for the
two years in offspring of women randomized to a trial prevention of diabetes in women with previous Related Links
of metformin for GDM: follow up data from the MiG gestational diabetes mellitus: a systematic review WHO: http://www.who.int/about/regions/en
trial. BMC Pediatr. 15, 54 (2015). and meta-​analysis. Front. Endocrinol. 9, 583 (2018).

NATURE REvIEWS | DiSeASe PRiMeRS | Article citation I­D:­ (2019) 5:47 ­ 19

View publication stats 0123456789();

You might also like