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

PRIMER

Rickets
Thomas O. Carpenter1, Nick J. Shaw2,3, Anthony A. Portale4, Leanne M. Ward5,
Steven A. Abrams6 and John M. Pettifor7
Abstract | Rickets is a bone disease associated with abnormal serum calcium and phosphate levels.
The clinical presentation is heterogeneous and depends on the age of onset and pathogenesis but
includes bowing deformities of the legs, short stature and widening of joints. The disorder can be
caused by nutritional deficiencies or genetic defects. Mutations in genes encoding proteins involved in
vitamin D metabolism or action, fibroblast growth factor 23 (FGF23) production or degradation, renal
phosphate handling or bone mineralization have been identified. The prevalence of nutritional rickets
has substantially declined compared with the prevalence 200 years ago, but the condition has been
re-emerging even in some well-resourced countries; prematurely born infants or breastfed infants who
have dark skin types are particularly at risk. Diagnosis is usually established by medical history, physical
examination, biochemical tests and radiography. Prevention is possible only for nutritional rickets and
includes supplementation or food fortification with calcium and vitamin D either alone or in
combination with sunlight exposure. Treatment of typical nutritional rickets includes calcium and/or
vitamin D supplementation, although instances infrequently occur in which phosphate repletion may
be necessary. Management of heritable types of rickets associated with defects in vitamin D
metabolism or activation involves the administration of vitamin D metabolites. Oral phosphate
supplementation is usually indicated for FGF23‑independent phosphopenic rickets, whereas the
conventional treatment of FGF23‑dependent types of rickets includes a combination of phosphate and
activated vitamin D; an anti‑FGF23 antibody has shown promising results and is under further study.

Rickets is a bone disease that is associated with decreased Rickets was almost universal among young children in
serum calcium and/or phosphate levels in the blood, large towns and cities in Europe and North America in the
leading primarily to widening and delay of mineraliza- late 19th and early 20th centuries3,4. However, following
tion of growth plates in bones1 (FIG. 1). Rickets is also the discovery of the role of cod liver oil (which contains
1
Department of Pediatrics
associated with osteomalacia, which is characterized by a high levels of vitamin D) and ultraviolet (UV) irradiation
(Endocrinology), School of
Medicine, Yale University, delay in the mineralization of bone matrix (FIG. 2). Rickets (involved in the formation of vitamin D in the skin) in
PO Box 208064, New Haven, can be caused by deficiencies of vitamin D, calcium or the prevention and management of vitamin D deficiency,
Connecticut 06520–8064, phosphate attributable to nutritional or environmental nutritional rickets was almost eradicated in several
USA. causes (that is, nutritional rickets) or by mutations in countries. Despite these early successes, the prevalence
7
South African Medical
Research Council–Wits
genes encoding proteins involved in vitamin D activa- of nutritional rickets has risen in at-risk communities
Developmental Pathways tion and function, phosphate homeostasis and/or bone in Europe and North America over the past 40 years5,6.
for Health Research Unit, mineralization (that is, heritable rickets); rickets can also A high prevalence of rickets has also been observed in
Department of Paediatrics, be the consequence of acquired defects in vitamin D children in the Middle East7, the Indian subcontinent, the
Faculty of Health Sciences,
metabolism (such as in severe liver disease) or renal northern provinces of China and in Mongolia8. A global
University of the
Witwatersrand, tubular handling of minerals (which may occur with a estimate of the risk of developing rickets is limited by
7 York Road, Parktown, number of drugs). Nutritional rickets is the most com- the paucity of basic data such as calcium and vitamin D
Johannesburg 2194, mon type of rickets globally 2. Although nutritional rick- intake or serum 25‑hydroxyvitamin D (25(OH)D; an
South Africa. ets can rarely be caused by dietary phosphate deficiency, inactive vitamin D metabolite) concentrations among
Correspondence to T.O.C. this form of rickets is most often the consequence of an children in non-industrialized countries.
and J.M.P.
inability to maintain serum calcium levels as a result of Although nutritional rickets remains the most com-
thomas.carpenter@yale.edu;
john.pettifor@wits.ac.za inadequate dietary calcium intake, vitamin D deficiency mon type of rickets globally, with the decline in its prev-
(either due to insufficient dietary intake or insufficient alence from the early 20th century, the heritable forms
Article number: 17101
doi:10.1038/nrdp.2017.101 sunlight exposure, which is required for the activation of rickets have become comparatively more common.
Published online 21 Dec 2017 of vitamin D) or a combination of both. The development of molecular genetic techniques has

NATURE REVIEWS | DISEASE PRIMERS VOLUME 4 | ARTICLE NUMBER 17101 | 1


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

Author addresses milk, nutritional rickets in the United States was virtu-
ally eliminated by the 1950s. Nevertheless, the disorder
1
Department of Pediatrics (Endocrinology), School of Medicine, Yale University, still occurs in the United States, Europe and many other
PO Box 208064, New Haven, Connecticut 06520–8064, USA. countries globally, as consistently evidenced by limited
2
Department of Endocrinology & Diabetes, Birmingham Children’s Hospital, case series. Case rate estimates of 2.9–27 per 100,000
Birmingham, UK.
individuals have been reported in the United States and
3
Institute of Metabolism and Systems Research, University of Birmingham,
Birmingham, UK. Europe, and dark-skinned infants who are breastfed for
4
Department of Pediatrics, University of California, San Francisco, California, USA. >6 months are particularly at risk5. Nutritional rickets
5
Department of Pediatrics, University of Ottawa, Ottawa, Ontario, Canada. seems to be less common in Australia and New Zealand,
6
Department of Pediatrics, Dell Medical School at the University of Texas-Austin, Austin, the Americas and Europe than in the Indian subconti-
Texas, USA. nent, Africa and the Middle East 2,11–14. The prevalence
7
South African Medical Research Council–Wits Developmental Pathways for Health in Turkey in 2008 was reported to be 100 per 100,000
Research Unit, Department of Paediatrics, Faculty of Health Sciences, University of the individuals2,15. As recently as the late 1900s, the rates of
Witwatersrand, 7 York Road, Parktown, Johannesburg 2194, South Africa. rickets in infants and young children were reported to
be extremely high in Tibet and Mongolia (>50%) and in
made it possible to identify the genetic cause of many a few African and Middle Eastern countries (>10%)16.
of these uncommon types of rickets. Through these dis- Globally, the disorder predominantly affects infants
coveries, the past 40 years have seen a rapid expansion and preschool-aged children11–14. In well-resourced
in our understanding of the causes and pathogenesis of countries, the prevalence is highest among immigrant
rickets in children. populations, those with dark skin and those with min-
In this Primer, we highlight the epidemiology, imal sunshine exposure. Breastfed infants are particu-
pathophysiology, diagnosis and management of ­rickets larly at risk because of the high prevalence of vitamin D
with a focus on advances in the understanding of the deficiency in many mothers and the low content of
disease. ­vitamin D and its metabolites in breast milk17.
In the United States5 and the United Kingdom6,
Epidemiology reports indicate that nutritional rickets is becoming
Nutritional rickets more common. Incidence estimates of rickets were
Rickets associated with vitamin D deficiency was wide- 24.1 per 100,000 children <3 years of age in the early
spread 120–200 years ago9, with a prevalence of 25% 2000s compared with 0 per 100,000 in the early 1970s
among infants and young children in parts of the United and 2.2 per 100,000 in the early 1980s in Minnesota,
States reported in 1870 (REF. 10). Definitive treatment of United States5. In the United Kingdom, 7.5 per 100,000
nutritional rickets with cod liver oil became common children <5 years of age were diagnosed with rick-
around the 1930s, and with vitamin D fortification in ets18. These observations have been attributed to an

a b
a c
Epiphysis

Resting zone
Growth Growth
plate Proliferating zone plate
Hypertrophic zone

Metaphysis

Figure 1 | Morphology of the growth plate in rickets. a,b | Morphology apoptotic chondrocytes. Apoptosis ofNature Reviews |chondrocytes
hypertrophic Disease Primers is
of a healthy, human growth plate (physis). The growth plate is induced by extracellular phosphate via phosphorylation of
characterized by maturation of chondrocytes (cartilage cells) occurring mitogen-activated protein kinase (MAPK) pathway intermediates and
progressively from the epiphysis to the metaphysis. The border between downstream inhibition of the caspase 9‑dependent mitochondrial
the metaphysis and the growth plate is marked by a provisional zone of apoptotic pathway43. Thus, reduction in ambient phosphate availability to
calcification of the cartilage matrix (red-pink staining) that undergoes the chondrocyte, which is common to all forms of rickets, seems to be
resorption and replacement with mineralized bone (turquoise staining). central to the impaired apoptosis. The ligand 1,25‑dihydroxyvitamin D
c | A rachitic growth plate showing the marked increase in longitudinal (1,25(OH) 2D) and its receptor might be involved as well 194. Finally,
width marked by the persistence of the zone of hypertrophic expansion of the hypertrophic chondrocyte zone can be induced by
chondrocytes that lose their columnar arrangement. The growth plate impaired vascularization, influenced by vascular endothelial growth
abnormalities are the consequence of impaired chondrocyte apoptosis factor, which is regulated by MAPK pathway intermediates195. Parts b and
and impaired mineralization of the cartilage matrix surrounding the c courtesy of F. H. Glorieux.

2 | ARTICLE NUMBER 17101 | VOLUME 4 www.nature.com/nrdp


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

a b c

Figure 2 | Histological characteristics of osteomalacia. Goldner’s trichrome staining of undecalcified iliac


Nature Reviews crest bone
| Disease Primers
samples showing mineralized bone in turquoise and unmineralized bone matrix (osteoid) in red-pink. Compared with a
control sample (part a), a sample obtained from a child with X‑linked hypophosphataemic rickets (part b) shows a marked
increase in the amount and thickness of osteoid (arrows) covering mineralized bone in the bone cortex or outer shell (*) and
the trabecular area, which is a typical feature of osteomalacia. Part c depicts abnormal accumulation of osteoid surrounding
the osteocytes (the intercommunicating osteogenic lineage cells that are embedded throughout the bone matrix), which is
another characteristic of osteomalacia in patients with X‑linked hypophosphataemic rickets known as peri-osteocytic
lesions. Parts a and b courtesy of F. H. Glorieux. Part c is from REF. 128, Macmillan Publishers Limited.

increasing number of dark-skinned immigrants as Rickets in adolescents


well as an increase in the use of sunscreen and other Most of the literature pertaining to the epidemiology
skin-protective barriers19. of rickets is focused on infants and young children
because rickets is identified most frequently in the first
Rickets in infants and young children few years of life2. Nevertheless, rickets will persist into
Currently, nutritional rickets in prepubertal children adolescence in those who are at risk of vitamin D defi-
is most common in prematurely born infants5,20 and ciency and in those with early-onset genetic forms, such
in infants who are breastfed for >6–12 months, have as X‑linked hypophosphataemia (XLH); genetic forms
dark skin and have inadequate vitamin D, calcium or, rarely first come to medical attention during teenage
rarely, phosphate intake21. Rickets associated with die- years27. Genetic forms of rickets are less common than
tary calcium deficiency was first described in an infant nutritional rickets. Indeed, XLH is the most common
with suspected food allergies who had been placed on a heritable form of hypophosphataemia, but the incidence
milk-free diet with very low calcium content 22. Rickets is only 1 in 20,000 individuals28.
has also been described in several infants who were fed Nutritional rickets may also present during ado-
macrobiotic or vegan diets that were low in calcium23. lescence. In such cases, the manifestations tend to be
Preterm birth is associated with an increased risk of different than those of infants and children, with iso-
developing rickets, especially when the birthweight is lated bone pain and proximal muscle weakness being
<1,000 g. In premature infants, rickets is primarily caused common presenting features29,30. Regional reports of
by insufficient calcium and phosphate intake, but not nutritional rickets in adolescents suggest that the con-
vitamin D deficiency 24. Data on the prevalence of rick- dition is highly prevalent in several regions, including
ets in preterm infants are limited. In one study, ~15% of the Middle East and north India, particularly among
infants in the United States with a birthweight of <1,000 g girls. In one study, 25% of children aged 10–13 years
showed radiological evidence of rickets20. This number is presenting to orthopaedic clinics in north India with a
lower than estimates in the early 1980s, which revealed a myriad of musculoskeletal complaints had rickets, with
presence of rickets in up to 50% of preterm infants25; this knee pain at night followed by leg deformity being the
improvement reflects the introduction of modern nutri- most common complaints31. In Saudi Arabia, 40% of
tional practices. In the United States, ~20,000 babies are adolescents with increased alkaline phosphatase levels
born each year weighing between 500 g and 1,000 g, and (a biochemical marker of rickets and osteomalacia)
6,000 babies are born weighing <500 g. Extrapolation of were found to have nutritional rickets; of these, 39%
these data based on the incidence estimate of rickets presented with bone pain, 18% with limb deformity,
of 15% and taking into account that some of these babies, 14% with pathological fractures and 29% were asymp-
especially those born <500 g, will not survive suggests tomatic29. A female preponderance was reported both
that ~3,000 babies are at risk of developing rickets annu- in India and in Saudi Arabia29,31. Globally, girls in the
ally in the United States, which makes rickets an impor- Middle East seem to have the highest rates of vitamin D
tant cause of neonatal morbidity 26. Extremely preterm deficiency, an observation that has been attributed to
infants (born at <28 weeks gestation) are now surviving cultural dress codes combined with inadequate calcium
with multiple complications and considerable treatment and vitamin D intake32–35. The sex difference is concern-
requirements, some of which increase the risk of rickets, ing in light of the implications for future offspring. Risk
for example, treatment with corticosteroids to quicken factors for vitamin D deficiency have been studied in
lung maturation. Medical conditions such as necrotizing Ethiopian adolescents, almost half of whom were vita-
enterocolitis, which is observed in 2–4% of infants with a min D deficient, and included living in urban settings,
birthweight of <1,500 g, and other conditions might add lower maternal socioeconomic and education status and
to the b
­ urden of rickets. ownership of televisions or computers35.

NATURE REVIEWS | DISEASE PRIMERS VOLUME 4 | ARTICLE NUMBER 17101 | 3


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

7-Dehydrocholesterol Dietary intake of vitamin D2 the mitochondrial enzyme 25‑­hydroxyvitamin D-1α‑­


(some plants and fortified foods) hydroxylase (further referred to as 1α‑hydroxylase;
7-Dehydrocholesterol encoded by CYP27B1). Activity of 1α‑hydroxylase in
reductase
the kidney is tightly regulated by PTH, calcium, phos-
Vitamin D3
phate, FGF23 and 1,25(OH)2D. Metabolic inactivation of
1,25(OH)2D and 25(OH)D occurs, primarily in the kidney
and intestine, via hydroxylation at the 24 position by the
Vitamin D
enzyme 25‑hydroxyvitamin D-24‑hydroxylase (further
referred to as 24‑hydroxylase; encoded by CYP24A1)36.
Classically, 1,25(OH)2D acts on the intestine to
stimulate calcium and phosphate absorption, thereby
25-Hydroxylase maintaining plasma concentrations of these ions at
levels sufficient for normal bone mineralization (FIG. 3).
25(OH)D
Such action depends on binding of 1,25(OH)2D to the
vitamin D receptor (VDR).

FGF23. FGF23 — a phosphaturic hormone produced by


1α-Hydroxylase PTH osteocytes — decreases the threshold for renal tubular
reabsorption of phosphate by decreasing the abundance
Degradation 1,25(OH)2D
24-Hydroxylase of the main renal phosphate transporters, sodium-­
dependent phosphate transport protein 2A (NPT2A;
also known as NaPi-IIa) and NPT2C (also known as
FGF23 ↑ Bone ↑ Calcium and ↑ Phosphate NaPi-IIc), on the apical surface of proximal renal tubular
resorption phosphate absorption excretion cells37,38 (FIG. 4). Thus, when FGF23 levels are increased,
serum phosphate is maintained at considerably lower
levels than normal.
In addition, FGF23 regulates vitamin D metabolism
by increasing the expression of CYP24A1 and down-
regulating the expression of CYP27B1 (FIGS 3,4). Thus,
Figure 3 | Regulation of calcium and phosphate homeostasis. Serum calcium abnormally increased FGF23 levels result in lower cir-
Nature Reviews | Diseaseand
Primers
phosphate levels are regulated by the concerted action of 1,25‑dihydroxyvitamin D culating levels of the 1,25(OH)2D metabolite than would
(1,25(OH)2D), fibroblast growth factor 23 (FGF23) and parathyroid hormone (PTH) on the normally be expected given the hypophosphataemia39.
intestine, bone and kidney. 25(OH)D, 25‑hydroxyvitamin D. The relative deficiency of 1,25(OH)2D may contribute to
the low serum phosphate level, as intestinal phosphate
absorption is positively regulated by 1,25(OH)2D (FIG. 3)
Mechanisms/pathophysiology via NPT2B (also known as NaPi-IIb), which is expressed
Calcium and phosphate homeostasis in the intestine and mediates ~50% of intestinal phos-
Serum calcium and phosphate levels are regulated by phate transport 40. Indeed, intestinal phosphate absorp-
vitamin D, fibroblast growth factor 23 (FGF23) and tion is decreased in a mouse model of XLH (Hyp mice)
parathyroid hormone (PTH) through actions on the compared with control mice41. Nevertheless, the primary
intestine, kidney and bone (FIG. 3). The development disruption incurred by excessive FGF23 action must be
of rickets, which is associated with abnormal serum in renal phosphate reabsorption, as the tubule is unable
calcium and/or phosphate levels, is often associated to compensate for any limitations in intestinal phosphate
with abnormalities in vitamin D, PTH and FGF23 absorption, which would normally occur in primary
­metabolism or action. ­disorders of nutritional phosphate deprivation.

Vitamin D. Vitamin D, a biologically inactive prohor- Calciopenic rickets


mone, exists as either ergocalciferol (vitamin D2) pro- Although various types of rickets have different causes
duced by fungi and yeast or as cholecalciferol (vitamin D3) and pathogenetic mechanisms42, the characteristic
produced by near-UV irradiation (290–315 nm) from growth plate abnormalities of rickets are similar, inde-
7‑dehydrocholesterol in human skin (FIG. 3). Vitamin D pendent of the cause (FIG. 1). BOX 1 highlights the three
is hydroxylated at the 25 position in the liver by one or broad categories of the causes of rickets, which are
more enzymes, including the microsomal enzyme vita- based on the primary defect in the pathogenesis of the
min D 25‑hydroxylase (further referred to as 25‑hydrox- disease: calcium-related disorders (calciopenic rickets),
ylase; encoded by CYP2R1). The activity of hepatic phosphate-related disorders (phosphopenic rickets) and
25‑­hydroxylase is not tightly controlled and, conse- those disorders associated with the direct impairment
quently, circulating concentrations of 25(OH)D are deter- of mineralization. Although the terms ‘calciopenic’ and
mined primarily by dietary intake and skin ­synthesis of ‘phosphopenic’ refer to the initiating clinical insults
vitamin D. Next, 25(OH)D undergoes ­1α‑­hydroxylation that eventually lead to rickets (that is, hypocalcaemia
(primarily in the kidney) to its hormonally active or hypophosphataemia), both mineral deficits will gen-
form, 1,25‑dihydroxyvitamin  D (1,25(OH) 2D), by erate hypophosphataemia either through secondary

4 | ARTICLE NUMBER 17101 | VOLUME 4 www.nature.com/nrdp


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

hyperparathyroidism in the case of calciopenic rickets In both forms of nutritional rickets, the absolute
or through mechanisms unrelated to secondary hyper- intestinal calcium absorption is inadequate, leading to
parathyroidism (that is, primary renal tubular defect a fall in ionized calcium levels and secondary hyper-
or elevated FGF23 levels) in the case of phosphopenic parathyroidism, with the consequent development of
rickets. Reduced ambient phosphate supply seems to increased renal phosphate loss and hypophosphataemia.
be the common pathway in the development of the Biochemically, rickets associated with vitamin D defi-
growth plate abnormalities observed in rickets, in both ciency and dietary calcium deficiency differ from each
­calciopenic and phosphopenic forms43. other, in that serum 25(OH)D values are typically higher
(>25–30 nmol per litre) and serum 1,25(OH)2D levels
Nutritional and environmental causes. The central are increased in dietary calcium deficiency compared
abnormality common to the two major causes of nutri- with vitamin D deficiency, which is associated with low
tional rickets (that is, dietary calcium deficiency and 25(OH)D and variable 1,25(OH)2D levels48.
vitamin D deficiency) is an inability to absorb suffi- FGF23 might also play a part in rickets owing to low
cient dietary calcium to meet the requirements of the dietary calcium intake. Indeed, an intriguing finding in
growing skeleton44 (FIG. 5). In dietary calcium deficiency, children with rachitic-like bone lesions and suspected
the poor calcium content or its limited bioavailability dietary calcium deficiency in the Gambia has been the
in the diet prevents sufficient absorption to meet the presence of markedly elevated plasma FGF23 levels,
growing child’s needs even though fractional intestinal which were inversely correlated with serum phosphate
calcium absorption may be maximized45 through ele- levels49. A potential explanation could be that low serum
vated 1,25(OH)2D levels46. In the case of vitamin D defi- calcium levels lead to increased PTH and 1,25(OH)2D
ciency, intestinal calcium absorption is impaired owing levels, which in turn lead to increased FGF23 secretion
to inadequate levels of 1,25(OH)2D as a consequence of (FIG. 3). An inverse relationship between elevated FGF23
insufficient substrate (25(OH)D) (FIG. 5). The combina- levels and haemoglobin values was also observed in this
tion of low calcium intake and poor vitamin D status study, which was marked in the children with rachitic-­
exacerbates the development and severity of rickets47. like bone deformities50. Whether FGF23 plays a part in
rickets associated with dietary calcium deficiency
in other communities will need further investigation.
Phosphate Hereditary defects in vitamin D synthesis, degrad­
ation or action can also give rise to hypocalcaemia,
Urine hypophosphataemia and rickets in children (BOX 1).
NPT2A
These types of rickets are rare, especially when com-
Proximal renal tubular cell pared with nutritional rickets. The disorders now known
NHERF1 NPT2C
P ↓1,25(OH)2D to be due to 1α‑hydroxylase deficiency and hereditary
1,25(OH)2D resistance were originally labelled ‘vita-
min D‑dependent’ rickets in reference to the obser-
SGK1 ↑ CYP24A1 vation that rickets recurred in these disorders when
↓ CYP27B1 pharmacologic doses of vitamin D were discontinued.
This finding contrasts with rickets caused by vitamin D
ERK1
ERK2 deficiency, in which disease recurrence can be prevented
Lysosome Nucleus with ­physiological doses of vitamin D.

25‑Hydroxylase deficiency. Rickets associated with low


Blood α-KLOTHO
to undetectable 25(OH)D levels can be caused by inac-
FGFR1 tivating mutations in the hepatic microsomal enzyme
FGF23
CYP2R1, the predominant 25‑hydroxylase in humans51,52
Figure 4 | Regulation of renal phosphate transport and vitamin D metabolism by
(FIG. 3). Individuals with 25‑hydroxylase deficiency pres-
FGF23. Fibroblast growth factor 23 (FGF23) specificallyNature Reviews
recognizes | Disease
the FGF Primers
receptor 1
(FGFR1) associated with the kidney-produced co-receptor α‑KLOTHO on the basolateral ent in a manner similar to those with rickets associ-
surface of renal tubular cells196. Downstream signalling through extracellular-signal-­ ated with vitamin D deficiency. Two different CYP2R1
regulated kinase 1 (ERK1), ERK2 and serine/threonine protein kinase SGK1 leads to mutations were identified in Nigerian children with
phosphorylation of sodium-hydrogen exchanger regulatory factor 1 (NHERF1) and familial rickets; both mutations dramatically reduced
translocation of the sodium-dependent phosphate transport protein 2A (NPT2A) 25‑­hydroxylase activity in vitro53,54. Further, two Saudi
and NPT2C from the apical membrane107. NPT2A is then degraded in the lysosomal Arabian siblings with severe vitamin D deficiency were
compartment, whereas NTP2C might be recycled to the membrane or degraded. found to be compound heterozygotes (presence of a dif-
Furthermore, FGF23 has been shown to downregulate transcription of NTP2A, providing ferent mutation in each allele) for CYP2R1 mutations that
another mechanism by which renal phosphate transport can be affected197. Parathyroid are predicted to encode null alleles55. Response to treat-
hormone contributes to the regulation of renal tubular phosphate reabsorption by its
ment differs between individuals who are homozygous
effect on NTP2A and NPT2C transporter removal from the apical membrane and may
interact with FGF23 in regulating membrane transporter abundance107,198. In addition and those who are heterozygous for CYP2R1 mutations.
to regulation of renal phosphate transport, FGFR1 activation inhibits CYP27B1
expression (encoding 1α‑hydroxylase) and stimulates CYP24A1 expression (encoding 1α‑Hydroxylase deficiency. 1α‑Hydroxylase deficiency
24‑hydroxylase), resulting in impaired production and increased degradation of causes rickets owing to loss‑of‑function mutations in
1,25‑dihydroxyvitamin D (1,25(OH)2D). CYP27B1 (FIG. 3), which often manifests within the first

NATURE REVIEWS | DISEASE PRIMERS VOLUME 4 | ARTICLE NUMBER 17101 | 5


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

Box 1 | Classification of rickets* 24 months of life. Rickets associated with 1­ α‑hydroxylase


deficiency is also known as vitamin  D‑dependent
Calciopenic rickets ­r ickets type  1, hereditary pseudo-vitamin  D defi-
Calciopenic rickets is the consequence of the inability to maintain serum calcium ciency ­r ­­ickets type 1 or vitamin D dependency 56,57.
concentrations in the normal range, leading to secondary hyperparathyroidism and In patients with 1α‑hydroxylase deficiency, serum con-
hypophosphataemia. The hypocalcaemia can be caused by vitamin D or dietary centrations of 25(OH)D are not decreased, which dis-
calcium deficiencies, and abnormalities of vitamin D metabolism and action.
tinguishes patients with 1α‑hydroxylase deficiency from
Vitamin D‑associated causes of rickets: those with nutritional vitamin D deficiency in whom
• Rickets associated with vitamin D deficiency due to insufficient vitamin D intake, serum 25(OH)D is reduced. Levels of 1,25(OH)2D are
insufficient epidermal ultraviolet radiation exposure or intestinal malabsorption ­consistently very low in 1α‑hydroxylase deficiency.
syndromes (such as obstructive jaundice)
Although 1α‑hydroxylase deficiency is rare in most
• 25‑Hydroxylase deficiency (also known as vitamin D‑dependent rickets type 1B): populations, it is particularly common among French
impaired 25‑hydroxylation of vitamin D owing to mutations in CYP2R1
Canadians, with an apparent carrier rate of 1 per 26 indi-
• 1α‑Hydroxylase deficiency (the inherited form also known as vitamin D‑dependent viduals58; inheritance is autosomal recessive. As of 2017,
rickets type 1A): impaired 1α‑hydroxylation of 25(OH)D owing to renal disease or ~60 different CYP27B1 mutations have been identified
mutations in CYP27B1
in >100 patients since the first described mutations in
• Hereditary 1,25(OH)2D-resistant rickets due to mutations in the VDR (also known as 1997 (REF. 59). All the described frameshift and nonsense
vitamin D‑dependent rickets type 2) or to overexpression of a vitamin D response
(premature translation arrest) mutations eliminate the
element (VDRE)-binding ribonucleoprotein (also known as vitamin D‑dependent
rickets type 2b) haeme-binding site of the 1α‑hydroxylase, resulting in a
protein devoid of enzymatic activity. Of the many mis-
• Increased degradation of vitamin D due to use of anticonvulsants (such as
phenobarbital, carbamazepine and phenytoin) or antimicrobials (such as rifampicin)
sense mutations reported, the majority 60,61 also eliminate
enzyme activity when assayed in vitro. Some patients
Rickets as a consequence of dietary calcium deficiency due to insufficient dietary
with mild clinical manifestations have CYP27B1 muta-
calcium intake or gastrointestinal disorders resulting in calcium malabsorption
tions that result in a 1α‑hydroxylase protein with partial
Phosphopenic rickets enzyme activity 60,61.
Phosphopenic rickets results from a primary inability to maintain serum phosphate
concentrations in the normal range due to impaired renal phosphate reabsorption or Hereditary 1,25(OH)2D-resistant rickets. Hereditary
reduced gastrointestinal phosphate absorption.
1,25(OH)2D-resistant rickets, which is caused by end-
Rickets as a consequence of dietary phosphate deficiency or impaired bioavailability: organ resistance to the action of 1,25(OH)2D usually
• Breastfed very low birthweight infants due to mutations in VDR, was first described in 1978
• Use of elemental or hypoallergenic formula diet or parental nutrition by Brooks et al.62 and is characterized by hypocalcaemia,
• Excessive use of phosphate binders hyperparathyroidism, normal serum concentrations of
• Gastrointestinal surgery or disorders 25(OH)D and substantially increased (3‑fold to 30‑fold)
Rickets caused by increased renal phosphate loss: serum 1,25(OH)2D levels. This disorder, initially termed
• Fibroblast growth factor 23 (FGF23)‑dependent types of rickets are associated with vitamin D‑dependent rickets type 2, is also known as
increased FGF23 levels owing to overproduction or reduced degradation of FGF23: pseudo-vitamin D deficiency type 2, calcitriol-resistant
-- X‑Linked hypophosphataemia (XLH): associated with mutations in PHEX rickets, hypocalcaemic vitamin D‑resistant rickets and
-- Autosomal recessive hypophosphataemic rickets (ARHR): associated with mutations hereditary 1,25(OH)2D-resistant rickets. Although most
in DMP1 (ARHR type 1) or ENPP1 (ARHR type 2) of the clinical, laboratory and radiographic findings in
-- Autosomal dominant hypophosphataemic rickets (ADHR): associated with patients with hereditary 1,25(OH)2D-resistant rickets62–65
mutations in FGF23 are similar to those in patients with 1α‑hydroxylase
-- Raine syndrome: associated with mutations in FAM20C deficiency or nutritional vitamin D deficiency, a strik-
-- Others: tumour-induced osteomalacia or polyostotic fibrous dysplasia,
ing difference is the finding of either sparse body hair
overexpression of α‑KLOTHO
or total alopecia in the majority of affected patients64–68.
• FGF23‑independent types of rickets:
Patients with alopecia, which can be present at birth
-- Due to mutations in SLC34A1 (encoding sodium-dependent phosphate transport
or develop within the first year of life, seem to have
protein 2A (NPT2A))
-- Due to mutations in SLC34A3 (encoding NTP2C), causing hereditary an earlier age of onset of rickets and greater resistance
hypophosphataemic rickets with hypercalciuria to 1,25(OH)2D treatment than those with hereditary
-- Due to mutations in NHERF1 1,25(OH)2D‑resistant rickets without alopecia68.
-- Due to disorders such as Fanconi syndrome or distal renal tubular acidosis The VDR is a member of the family of nuclear
Rickets associated with a direct inhibition of mineralization transcription factors69 and is composed of an amino-­
Direct impairment of the process of mineralization at the growth plate
terminal DNA binding domain and a carboxy-­terminal
ligand-binding domain. The VDR heterodimerizes
• Hereditary hypophosphatasia
preferentially with the retinoid‑X receptor (RXR),
• Treatment with etidronate (first-generation bisphosphonate)
enabling optimal binding to vitamin D response ele-
• Fluoride excess ments (VDREs), DNA motifs that are critical for VDR-
• Aluminium excess dependent gene regulation. Approximately 50 different
*This classification provides a rational division of the numerous causes of rickets into three mutations in VDR have been described70; they have been
major categories based on pathogenesis, biochemical abnormalities and methods of treatment categorized into those that affect the ligand-­binding
and, consequently, provides a straightforward approach to diagnosis and management. domain, which interrupt binding of 1,25(OH)2D to
Please note that this list is incomplete, as rare causes of rickets have been omitted.
VDR, and those that affect the DNA-binding domain,

6 | ARTICLE NUMBER 17101 | VOLUME 4 www.nature.com/nrdp


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

Vitamin D deficiency Dietary calcium deficiency

↓ Conversion to vitamin D3 in the skin (UV) Inadequate dietary calcium intake


↓ Dietary intake of vitamin D2 or vitamin D3

Inadequate intestinal ↑ Fractional intestinal


↓ Serum 1,25(OH)2D calcium absorption calcium absorption
VDR

↓ Serum calcium
25-Hydroxylase
↑ Serum PTH

1α-Hydroxylase ↓ Serum phosphate ↑ Serum 1,25(OH)2D

↓ Chondrocyte apoptosis
↓ Serum 25(OH)D ↑ Delayed endochondral calcification

Rickets Impaired
Osteomalacia mineralization

Figure 5 | The pathogenesis of rickets associated with nutritional vitamin D or dietary calcium
Nature deficiency.
Reviews | Disease Primers
Vitamin D deficiency (left panel) leads to a fall in 25‑hydroxyvitamin D (25(OH)D) levels, resulting in a reduction in
1,25‑dihydroxyvitamin D (1,25(OH)2D) levels and consequent impairment of intestinal calcium absorption and fall in serum
calcium. Secondary hyperparathyroidism results in an increase in renal phosphate excretion and hypophosphataemia and
stimulation of 1α‑hydroxylase activity that fails to increase 1,25(OH)2D levels owing to a lack of substrate (25(OH)D). The
resultant hypophosphataemia is primarily responsible for the development of osteomalacia and rickets. Low dietary
calcium intake (right panel) results in a fall in serum calcium levels and an increase in parathyroid hormone (PTH) levels,
which promotes renal phosphate loss and increases 1α‑hydroxylase activity. The consequent increase of 1,25(OH)2D levels
increases the fractional intestinal calcium absorption, which might partially correct serum calcium levels with slightly
elevated PTH levels and mild hypophosphataemia. However, if calcium intake decreases further or calcium demand
increases, intestinal calcium transport can no longer compensate and normocalcaemia cannot be maintained; secondary
hyperparathyroidism, elevated 1,25(OH)2D levels and hypophosphataemia are established with consequent development
of rickets and osteomalacia. UV, ultraviolet; VDR, vitamin D receptor.

which interfere with binding of the 1,25(OH)2D–VDR calcium deficiency or vitamin D deficiency. However, in
complex to the VDREs in DNA. Of interest, alopecia is premature infants, phosphate deficiency can dominate
observed in patients with mutations in the DNA-binding the clinical picture of rickets, particularly in breastfed
domain or in regions that inhibit hetero­dimerization preterm infants; these infants may require high levels of
with RXR or in those with mutations that cause trun- phosphate supplementation via oral or parenteral routes
cation of the VDR. By contrast, patients with muta- to meet requirements80,81. A recent report has identified
tions in the ligand-binding domain or co-activator that the use of amino acid-based elemental formulas,
binding regions without defective heterodimerization such as Neocate (Nutricia), can result in hypophospha-
do not develop alopecia71–73. Studies in Vdr-null mice taemic rickets, particularly in infants with coexistent
have helped to establish the critical role of the VDR gastrointestinal diseases. Although such formulas have
in hair follicle development 74–76, which presumably appropriate phosphate content, the bioavailability seems
involves WNT signalling via β‑catenin and Hedgehog to be impaired in certain clinical settings82. Similarly,
signalling pathways77. excessive use of antacids that bind dietary phosphate can
Several patients with apparent resistance to result in symptomatic hypophosphataemia and bone dis-
1,25(OH)2D but with no identified VDR mutation have ease83, a complication that has been reported in all ages.
been described. One such individual, with the biochem- Parenteral nutrition and gastrointestinal surgery may
ical and clinical hallmarks of hereditary 1,25(OH)2D- also limit adequate phosphate delivery to infants or older
resistant rickets including alopecia, has been found to patients. In these cases, phosphate supplementation gen-
overexpress a heterogenous nuclear ribonucleoprotein erally corrects the problem. Finally, congenital impair-
that competes with VDR for VDRE binding, illustrative ment of combined intestinal phosphate malabsorption
of yet another mechanism for vitamin D resistance78. and renal reabsorption has been reported84.

Phosphopenic rickets FGF23‑mediated causes. This group of disorders is


Nutritional cause. Phosphate is required together with caused by either overproduction of FGF23 or reduced
calcium for mineralization of bones. Thus, severe phos- degradation of intact FGF23, thereby leading to excess
phate deficiency can lead to osteomalacia and rickets79. FGF23 levels and hypophosphataemia (BOX 1; FIG. 3).
However, phosphate-deficient diets are rare, and the clin- As shown in FIG. 6a, excess circulating FGF23 ­levels may
ical picture of nutritional rickets is usually that of dietary result from abnormalities in genes encoding proteins

NATURE REVIEWS | DISEASE PRIMERS VOLUME 4 | ARTICLE NUMBER 17101 | 7


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

a Physiological b Fibrous dysplasia


regulators
• ↑ Phosphate
Osteocyte • ↑ 1,25(OH)2D
• ↓ Iron Gain-of-function
Genetic regulators • ↑ α-KLOTHO mutations in GNAS
Loss-of-function
mutation in:
• PHEX
• DMP1 ↑ Protein kinase A
• ENPP1 signalling
• FAM20C

↑ FGF23
Gain of function
of FGFR1 ↑ FGF23

c Tumour cell associated with d Epidermal nevus


tumour-induced osteomalacia syndrome

Fusion proteins
(FN1–FGFR1 or
FN1–FGF1 fusions) Gain-of-function
mutations in:
• HRAS
• NRAS
Increased FGFR1
signalling

↑ FGF23 ↑ FGF23

Figure 6 | FGF23: tissue sources and mechanisms of regulation. a | The regulation of the production or secretion of
fibroblast growth factor 23 (FGF23) in osteocytes is not completely understood, but phosphate Natureitself and | Disease Primers
Reviews
1,25‑dihydroxyvitamin D (1,25(OH)2D) stimulate production of FGF23 (REF. 199). Iron deficiency results in increased FGF23
secretion200 and, paradoxically, administration of certain iron salts (that is, ferric carboxymaltose) will increase FGF23
levels in humans201,202. As evidenced by the presence of rickets associated with loss of function of PHEX, DMP1 and ENPP1,
proteins encoded by these genes regulate FGF23 production, but the mechanisms are unknown201–203. Furthermore, FGF23
production and secretion by osteocytes depends on the FGF receptor 1 (FGFR1). b–d | Excessive production of FGF23
occurs in various pathological conditions, such as the skeletal fibrous dysplasia in McCune–Albright syndrome (part b),
which is due to activating mutations in GNAS. Signalling through FGFR1 can stimulate FGF23 production, a process
thought to be amplified by somatic mutations resulting in FGFR1 fusion proteins in tumours associated with
tumour-induced osteomalacia (part c). Finally, activating mutations in HRAS and NRAS cause epidermal nevus syndrome,
which is associated with increased FGF23 levels owing to unknown mechanisms. FN, fibronectin.

involved in the regulation of FGF23 expression or of FGF23, which, in turn, enables increased glycosylation
abnormalities in regulators of FGF23 metabolism85. at nearby critical residues in FGF23, such that cleavage of
XLH is caused by mutations in PHEX and autosomal FGF23 is limited88,89. Ultimately, the mutation can result
recessive hypophosphataemic rickets (ARHR) is caused in FGF23‑mediated hypophosphataemia with variable
by mutations in DMP1 or ENPP1. All these mutations occurrence of rickets or osteomalacia90.
are associated with increased FGF23 levels, but the exact Other mechanisms by which FGF23 secretion seems
mechanisms are unknown. FGF23 is active as a full- to be regulated include activation of the protein kinase A
length molecule and is enzymatically cleaved into inac- pathway through activating mutations in GNAS, as
tive amino-terminal and carboxy-terminal fragments86. occurs in the fibrous dysplasia lesions observed in
The prototype disorder of defective degradation of intact McCune–Albright syndrome91, and exposure to high lev-
FGF23, autosomal dominant hypophosphataemic rick- els of circulating α-KLOTHO92,93 (FIG. 6a,b). Excess FGF23
ets (ADHR), results from specific missense mutations secretion in the pathological setting of tumour-induced
resulting in amino acid substitutions at residues 176–179, osteomalacia has been speculated to result from a
a protease recognition site. Such mutations disrupt the somatic rearrangement, resulting in a fusion protein
RXXR motif of this recognition site, thereby rendering consisting of the amino terminus of fibronectin and
the molecule resistant to proteolytic cleavage, resulting in the FGF receptor 1 (FGFR1) in a substantial proportion
excess accumulation of active, intact FGF23 (REF. 87). of tumours causing this syndrome (FIG. 6c). Germline
In addition, loss‑of‑function mutations of the extracellu- mutations in FGFR1 have been identified as causing
lar serine/threonine protein kinase FAM20C, as observed osteoglophonic dysplasia, a particular skeletal dysplasia
in Raine syndrome, can interfere with phosphorylation with FGF23‑mediated hypophosphataemia, pronounced

8 | ARTICLE NUMBER 17101 | VOLUME 4 www.nature.com/nrdp


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

short stature and characteristic facial features94. More normal to low. Long-term oral administration of neutral
recently, another rearrangement generating a fibronectin phosphate salts alone is associated with improvements
and FGF1 fusion protein has been described95. Another in the hypophosphataemia, bone pain, muscle strength,
unusual disorder, epidermal nevus syndrome, also growth rate and healing of rickets. Hypophosphataemia
known as cutaneous skeletal hypophosphataemia syn- and rickets also can occur in children as part of renal
drome, is characterized by somatic activating mutations Fanconi syndrome associated with, for example, nephro-
in HRAS or NRAS in both skin and bone and can lead to pathic cystinosis, oculocerebrorenal syndrome (also
excess FGF23 production by unknown mechanisms96,97 known as Lowe syndrome) or Dent disease.
(FIG. 6d). Finally, FGF23 can be stimulated by adminis- Mutations in the gene encoding the sodium-­hydrogen
tration of specific iron products, such as ferric carboxy­ regulatory factor 1 (NHERF1) have been identified in
maltose that is used to treat refractory iron deficiency, patients with hypophosphataemia due to renal phosphate
resulting in considerable hypophosphataemia that may wasting; these patients can also present with nephrolithia-
masqerade as tumour-induced osteomalacia98. sis (renal stone formation), osteomalacia and normal cir-
culating levels of PTH105,106. Interaction of NHERF1 with
FGF23‑independent causes. Hypophosphataemia disor- NPT2A is important for proper expression of NPT2A
ders with renal phosphate wasting attributable to primary at the plasma membrane (FIG. 4). In addition, NHERF1
renal defects may be genetic or acquired. The genetic is involved in PTH signalling by modulating the PTH-
causes are most often due to mutations in genes encoding induced generation of cAMP107. Mutations located in the
NPT2A (that is, SLC34A1) or NPT2C (that is, SLC34A3). PDZ2 domain or the inter-region between PDZ1 and
Homozygous, loss‑of‑function mutations in SLC34A1 PDZ2 of NHERF1 had no effect on phosphate uptake
were identified in two adult siblings from a consanguin- when expressed in vitro but rather stimulated PTH-
eous family with a form of hypophosphataemic rickets induced generation of cAMP and, consequently, exac-
and Fanconi syndrome (a syndrome of impaired renal erbated the inhibition of phosphate transport by PTH105.
solute reabsorption in the proximal tubules caused by By contrast, mutations in the PDZ1 domain were asso-
diseases or toxicity) inherited in an autosomal recessive ciated with reduced abundance of NPT2A at the plasma
pattern99. As children, both siblings had hypophospha- membrane but had no effect on PTH-induced cAMP
taemic rickets, severe bone deformities, fractures and production106. It seems that the PTH-dependent pathway
stunted growth100, Fanconi syndrome without metabolic that stimulates NPT2A movement from the plasma mem-
acidosis, high serum levels of 1,25(OH)2D and hyper- brane converges with the FGF23‑dependent pathway at
calciuria. As adults, the hypophosphataemia persisted, NHERF1; activation of either pathway can lead to inhibi-
but serum 25(OH)D and 1,25(OH)2D levels were low, tion ­of phosphate transport through NHERF1 (REF. 107).
plasma FGF23 levels were low to normal and glomerular
filtration rate was reduced99. The mutant NPT2A protein, Direct inhibition of mineralization
when expressed in Xenopus laevis oocytes and opossum Although ambient hypophosphataemia is considered
kidney cells, failed to transport phosphate owing to fail- the major cause of rickets in XLH, other factors likely
ure of the transporter to reach the plasma membrane99. contribute to the bone abnormalities. Indeed, transgenic
Mutations in SLC34A1 were also identified in 16 patients mouse models designed to rescue XLH by overexpressing
with a form of idiopathic infantile hypercalcaemia with- PHEX in osteoblasts of Hyp mice do not entirely remedi-
out mutations in CYP24A1 (REF. 101). These patients, ate the underlying osteomalacic bone disease of the Hyp
four from consanguineous families and twelve unrelated, mice108–110. In humans and in mouse models, phosphate
presented in early infancy with hypophosphataemia due supplementation does not entirely correct the bone dis-
to renal phosphate wasting, failure to thrive, polyuria, ease111,112. Whether the inability to completely rescue the
symptomatic hypercalcaemia, high serum concentrations phenotype of Hyp mice can be overcome with more-­
of 1,25(OH)2D and medullary nephrocalcinosis (renal sustained physiological exposure to phosphate or with
parenchymal calcification). The clinical phenotype is inhibition of FGF23 remains to be seen.
attributed to hypophosphataemia owing to primary renal Although a number of explanations might exist for
phosphate wasting and suppression of FGF23 expression, these observations, it is possible that FGF23 or the PHEX
leading to inappropriate activation of 1,25(OH)2D with mutation itself have direct effects on mineralization inde-
subsequent hypercalcaemia and hypercalciuria. pendent of phosphate. Along those lines, the presence of
Inactivating mutations of SLC34A3, which encodes α-KLOTHO has recently been identified in bone cells,
NPT2C, give rise to the autosomal recessive disorder albeit at relatively low levels, which potentially allows for
hereditary hypophosphataemic rickets with hyper­ a targeted role for FGF23 in bone113. Pyrophosphate, a
calciuria (HHRH)102–104. HHRH is characterized by mineralization inhibitor, has been shown to accumulate
early childhood onset of hypophosphataemia second- in the bone matrix of Hyp mice114. Finally, fragments of
ary to renal phosphate wasting, rickets or osteomala- the family of small integrin-binding ligand, N‑linked
cia, lower limb deformities, muscle weakness and bone glycoprotein (SIBLING) proteins, that is, acidic serine
pain. In contrast to XLH, serum levels of 1,25(OH)2D in aspartate-rich MEPE-associated motif (ASARM) pep-
HHRH are elevated and, presumably as a consequence, tides115,116, have been shown to directly inhibit minerali-
intestinal calcium absorption is increased, resulting in a zation in vitro and to accumulate in the absence of PHEX.
compensatory increase in calcium excretion in the urine. Other causes of direct inhibition of mineralization are
Circulating PTH102,103 and FGF23 (REF. 104) levels are mentioned in BOX 1.

NATURE REVIEWS | DISEASE PRIMERS VOLUME 4 | ARTICLE NUMBER 17101 | 9


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

a b of the anterior fontanelle, craniosynostosis (premature


closure of the skull sutures) and delayed eruption of the
teeth with enamel hypoplasia. In XLH, dental abscesses
associated with defects in the consolidated mineraliza-
tion of globular dentin may come to the attention of the
family before overt evidence of skeletal disease. Chest
deformities include a rachitic rosary (expansion of the
costochondral junctions) and Harrison sulcus (groove
in the lower thoracic cage in which the diaphragmatic
attachments pull the softened bone inward). Associated
features that may occur in rarer forms of rickets include
features of McCune–Albright syndrome, early-onset hear-
ing loss120 and melanocytic nevi (pigmented mole-like
lesions)96. Complications of XLH that present in adult-
hood include enthesopathy (mineralizing abnormalities
of tendon and ligament insertion sites), osteophytes and
arthritic-like degeneration of joints, none of which are
well understood121.

Investigations
Figure 7 | Clinical manifestations of rickets. a | A childNature
with X‑linked
Reviews | Disease Primers The diagnosis of rickets is made on the basis of medical
hypophosphataemia with a windswept deformity of the legs characterized by a history, physical examination and biochemical testing
valgus deformity (genu valgus) of the left leg (inward bowing of the knee) and a
and is confirmed by radiography. The use of molecular
mild varus deformity (genu varus) of the right leg (an outward bowing of the knee).
b | The typical widened wrist in active rickets. This bone deformity is frequently more
genetic techniques enables identification of the mutations
apparent visually than on palpation and involves abnormalities in the distal metaphyses causing heritable types of rickets. Management of the dif-
and growth plates of the radius and ulna. The deformity may take some time to resolve ferent types of rickets differs markedly; thus, a specific
after appropriate treatment of the rickets. In severe rickets, widening may also occur at diagnosis for each patient with rickets is imperative not
the knees and distal tibias. only to inform parents of a possible genetic aetiology but
also to provide appropriate information on prevention,
prognosis and management.
Diagnosis, screening and prevention
Clinical presentation Medical history. Important enquiries when taking the
The clinical presentation of rickets is heterogeneous, medical history of a child suspected to have rickets include
involves skeletal and non-skeletal manifestations and environmental risk factors for nutritional rickets (BOX 2), an
depends on the age of presentation. Many children who assessment of dietary calcium intake and a family history,
present with rickets in early childhood will have progres- which might suggest an inherited form of rickets. A sim-
sive bowing deformities of the legs (genu varum) (FIG. 7a) ple food intake assessment can be used to provide relia-
often associated with a characteristic waddling gait. ble information on dietary calcium intake122. An intake
In children who develop rickets in the first year of life, the of <300 mg of calcium per day in a child aged >12 months
ability to walk is typically delayed (>18 months of age). is consistent with dietary calcium deficiency, whereas
When rickets develops in older children (aged >3 years), infants aged 0–6 months and 6–12 months should have a
a knock-knee deformity (genu valgum) or a combina- daily dietary intake of 200 mg and 260 mg, respectively 15.
tion of both genu varum and genu valgum is common117
(FIG. 7a). The age dependency of the presentation as genu Biochemical investigations. The typical biochemical pic-
varum or genu valgum reflects the normal growth pat- ture of a child with calciopenic rickets is low or normal
tern of the lower limb in the second year of life when the serum calcium levels with increased serum PTH levels
tibio-femoral angle changes from varus to valgus. Older leading to decreased renal tubular reabsorption of phos-
children with rickets may complain of bone pain and phate and low serum phosphate levels. By contrast, in
fatigue. Fractures may be a presenting feature in a child typical phosphopenic rickets, normal serum calcium,
with nutritional rickets but only in the presence of clear low serum phosphate and normal serum levels of PTH
radiological features of rickets15. Loss of height and short and 25(OH)D are observed. Serum alkaline phosphatase
stature might occur as a consequence of the progressive levels are usually markedly raised in calciopenic rickets
leg deformities. Infants with rickets associated with vita- owing to the high PTH causing increased bone turno-
min D deficiency or abnormalities of vitamin D metabo- ver but are often only modestly raised in phosphopenic
lism may also develop seizures owing to hypocalcaemia118 ­rickets (especially in XLH) (TABLE 1).
and life-threatening cardiomyopathy 119; the presence of Measurement of serum PTH is essential to distinguish
vitamin D deficiency rickets should always be investigated between calciopenic and phosphopenic rickets. In calcio­
when children present with these conditions. penic rickets, serum PTH levels will be considerably
Additional clinical signs seen in an infant with rickets increased, whereas serum PTH levels are often in the
include craniotabes (softening of the skull bones), widen- normal range or only slightly increased in phosphopenic
ing of the wrists (FIG. 7b), frontal bossing, delayed closure rickets.

10 | ARTICLE NUMBER 17101 | VOLUME 4 www.nature.com/nrdp


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

Serum phosphate levels should be interpreted with exclude more generalized tubular defects by measuring,
care as reference phosphate values are higher in young for example, pH or glucose and protein levels. Calcium
children and infants (1.3–2.1 mmol per litre) than in excretion is measured as a ratio of urine calcium to urine
adults (0.7–1.3 mmol per litre). Thus, it is important to creatinine levels (careful note of the units used to meas-
ensure that a reference range for adults is not quoted for ure the concentrations needs to be taken to interpret the
serum phosphate levels in children; a serum phosphate results correctly), whereas assessment of renal excretion
level of <1.0 mmol per litre is always abnormal in a child. of phosphate is best performed by calculating the tubular
Although serum 25(OH)D concentrations are con- reabsorption of phosphate and the tubular maximum
sidered to reflect vitamin D nutritional status, the serum for phosphate reabsorption127 using an equation that
level below which the clinical consequences of vitamin D includes serum phosphate and creatinine.
deficiency occur has been difficult to define and remains
a contentious issue. The US Institute of Medicine (IOM) Radiography. Radiographs of the wrist and knee enable
has defined vitamin D deficiency as serum 25(OH)D the diagnosis of rickets (FIG. 8). Typical changes seen in
levels of <30 nmol per litre; this cut-off reflects the level the metaphyseal region are cupping, fraying and lateral
below which the risk of developing rickets and meta- widening with expansion of the growth plate. Bones
bolic bone disease increases123. However, a task group of with metaphyseal abnormalities may show evidence of
the Endocrine Society defined vitamin D deficiency as osteopenia in calciopenic rickets owing to bone dem-
serum 25(OH)D levels of <50 nmol per litre124, although ineralization as a consequence of the high PTH levels.
the criteria used to define vitamin D deficiency were By contrast, in XLH, the bones may become somewhat
different from those used by the IOM. More recently, sclerotic by adulthood, and dense vertebrae can be a
a report from a scientific panel in the United Kingdom particular feature of ARHR type 1, even in childhood.
suggested that the risk of rickets increases at serum XLH and ARHR type 1 can be further distinguished
25(OH)D levels <25 nmol per litre125. Several factors from other forms of hypophosphataemia by the pres-
likely influence whether or not clinical or biochemical ence of characteristic peri-osteocytic lesions on bone
abnormalities occur at low levels of 25(OH)D; among biopsy samples128,129 (FIG. 2c). Other forms of hypophos-
these are the severity and duration of low 25(OH)D lev- phataemic rickets may be associated with severe loss of
els, habitual dietary calcium intake and its bioavailability, trabecular and cortical bone, which is well described in
the rate of bone growth and mineralization and genetic HHRH and tumour-induced osteomalacia. Occasionally,
polymorphisms that might influence the total levels of fractures of long bones may be seen in both calciopenic
25(OH)D and its circulating available component 126. and phosphopenic (excluding XLH) forms of rick-
It is worth collecting an additional serum sample ets in conjunction with osteopenic bone and typical
before starting any treatment for subsequent analysis for metaphyseal changes.
1,25(OH)2D or FGF23 if the diagnosis is not explained
by the initial investigations. A urine sample, ideally in Differential diagnosis
the fasting state, is an important additional investigation A number of conditions exist of which radiographic and
to assess renal excretion of calcium and phosphate and to clinical features can be mistaken for rickets. In meta-
physeal chondrodysplasia (typically Schmid type), meta­
physeal changes are seen on radiographs in conjunction
Box 2 | Environmental risk factors for nutritional rickets with short stature and a waddling gait130 (FIG. 9). However,
serum biochemistry is normal in this condition. Jansen
Certain environmental risk factors can predispose a child to develop nutritional rickets type metaphyseal chondrodysplasia due to an activating
and should be taken into account when rickets is suspected. The relative importance of
mutation of PTHR1 (encoding the PTH1 receptor) can
these factors will vary among different childhood populations around the world15.
also have a radiographic appearance similar to that of
Vitamin D deficiency rickets131. Bone turnover is increased with an increased
• Dark skin pigmentation serum alkaline phosphatase level, and the serum calcium
• Living at a high latitude during winter or spring with lack of ultraviolet B radiation level is often raised with PTH levels suppressed.
exposure as a consequence Infants with osteopetrosis (a condition in which
• Restricted sun exposure, for example, due to cultural skin covering or pollution bones become dense and brittle) due to a defect in
• Lack of vitamin D supplementation during infancy osteo­clastic bone resorption can have metaphyseal
• Prolonged exclusive breastfeeding changes that resemble rickets termed osteopetro­
• Maternal vitamin D deficiency rickets (REF. 132). However, the long bones are strikingly
dense in this condition. Hypocalcaemia and secondary
Dietary calcium deficiency
­hyperparathyroidism may be evident in severe cases.
• Limited dairy product intake, for example, due to cow milk intolerance or lack of Another important but rare condition is the infantile
access to milk
form of hereditary hypophosphatasia, in which the radio­
• Unusual diets, for example, veganism in infants graphic appearance may resemble rickets133. This condi-
• Malabsorption, for example, cholestatic liver disease* tion, due to inactivating mutations in TNSALP (encoding
Heritable forms of rickets are more likely to occur if there is a known family history (for example, tissue-nonspecific alkaline phosphatase), may present
X‑linked hypophosphataemic rickets) or in children born to consanguineous parents (for in the perinatal period to the first 6 months of life with
example, 1α‑hydroxylase deficiency). *Cholestatic liver disease is also associated with impaired poor feeding and failure to thrive due to hypercalcaemia.
absorption of vitamin D.
Respiratory insufficiency due to the small and poorly

NATURE REVIEWS | DISEASE PRIMERS VOLUME 4 | ARTICLE NUMBER 17101 | 11


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

mineralized ribcage is often evident. The key biochemi- particularly of infants, has been the main method used
cal marker that distinguishes infantile hypo­phosphatasia to prevent nutritional rickets, with varying evidence of
from true rickets is a low, rather than increased, serum success. Current consensus opinion, supported by a ran-
alkaline phosphatase level. Additional investigations, domized clinical trial involving different vitamin D doses
such as the measurement of the urinary phospho­ in breastfed infants137, is that a daily dose of 400 interna-
ethanolamine to creatinine ratio and plasma pyridoxal tional units (IU) of vitamin D given to all infants from
phosphate, will help to confirm the diagnosis as they birth to 12 months of age, regardless of mode of feeding,
are typically increased in children with infantile hypo­ is adequate to prevent rickets15. Examples of successful
phosphatasia. Early diagnosis of this condition, which public health strategies using 400 IU vitamin D daily exist
has a high mortality rate, is important as treatment with that have reduced the incidence of rickets and sympto-
recombinant alkaline phosphatase is now available and matic vitamin D deficiency 138,139. However, ensuring
can heal skeletal manifestations and improve survival134. adequate daily supplementation to infants is difficult 140.
For example, the annual rate of vitamin D supplement
Prevention use was only 17.9% despite a policy of free vitamin D
Prevention is possible only for nutritional rickets, not supplements for all breastfed infants141. Alternative strat­
for heritable forms. Adequate dietary calcium intake is egies to increase effective use of vitamin D supplements
necessary to prevent nutritional rickets due to calcium need to be considered. A randomized clinical trial of
deficiency, with a recommended intake of 500 mg daily daily vitamin D (400 IU) versus a bolus dose of 50,000
in children aged 1–3 years. In communities in which die- IU of vitamin D3 in babies born to vitamin D‑deficient
tary calcium intakes are habitually low, calcium intakes mothers in Melbourne, Australia, demonstrated that the
can be increased effectively with calcium fortification bolus dose could safely achieve similar 25(OH)D levels
of, for example, cereals or laddus (an Indian sweet ball at 3–4 months as the daily dose142. Maternal supplemen-
made with flour, ghee and sugar)135. In Nigeria, both tation with high doses of vitamin D (such as 6,400 IU
powdered limestone and ground dried fish have been per day)143 was safe and effective at maintaining suffi-
used to increase dietary calcium intakes by their addition cient 25(OH)D levels, but some concerns have been
to porridge136. expressed about the possible risk of complications related
To ensure an adequate vitamin D status, three strat- to the high dose of vitamin D given to the mother, which
egies have been proposed that can be used alone or passes to the infant in breastmilk. Caution about its rou-
together: vitamin D supplementation, food fortification tine acceptance has been raised until more large-scale
and sunlight exposure. Vitamin D supplementation, clinical trials have been conducted15.

Table 1 | Serum and urinary biochemistry associated with different types of rickets
Condition Associated gene sCa sP sPTH s25(OH)D s1,25(OH)2D sALP sFGF23 uP uCa
mutation
Calciopenic rickets
Dietary calcium NA = or ↓ ↓ ↑ = ↑ ↑ ? Variable ↓
deficiency
Vitamin D NA = or ↓ ↓ ↑ ↓ Variable ↑ = Variable ↓
deficiency
25‑Hydroxylase CYP2R1 = or ↓ ↓ ↑ ↓ ↓ ↑ ? Variable ↓
deficiency
1α‑Hydroxylase CYP27B1 ↓ ↓ ↑ = ↓ ↑ ? Variable ↓
deficiency
Hereditary VDR ↓ ↓ ↑ = ↑ ↑ ? Variable ↓
1,25(OH)2D-
resistant rickets
Phosphopenic rickets
Phosphate NA = or ↑ ↓ = = ↑ or = ↑ ↓ or = ↓ ?
deficiency
XLH PHEX = ↓ = = ↓ or = ↑ ↑ ↑ ↓ or =
ADHR FGF23 = ↓ = = ↓ ↑ ↑ ↑ ?
ARHR type 1 DMP1 = ↓ = = ↓ ↑ ↑ ↑ ?
ARHR type 2 ENPP1 = ↓ = = ↓ ↑ ↑ ↑ ?
HHRH SLC34A3 = ↓ = = ↑ or = ↑ ↓ or = ↑ ↑
25(OH)D, 25‑hydroxyvitamin D; 1,25(OH)2D, 1,25‑dihydroxyvitamin D; ADHR, autosomal dominant hypophosphataemic rickets; ALP, alkaline phosphatase;
ARHR, autosomal recessive hypophosphataemic rickets; Ca, calcium; FGF23, fibroblast growth factor 23; HHRH, hypophosphataemic rickets with hypercalciuria;
NA, not applicable; P, phosphate; PTH, parathyroid hormone; s, serum; u, urinary; XLH, X‑linked hypophosphataemia; ↑, increased levels; ↓, decreased levels;
=, normal levels; ?, not known.

12 | ARTICLE NUMBER 17101 | VOLUME 4 www.nature.com/nrdp


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

a b between sunlight exposure and 25(OH)D levels149,150,


Femur but these studies have usually been performed in light-
skinned children and may not be appropriate for those
with darker skin types. A study in adults using whole-
body irradiation chambers to simulate summer sunlight
exposure showed that 30 minutes of sunlight three times
per week for 6 weeks would achieve a 25(OH)D level of
>50 nmol per litre in >90% of white adults151. However,
no south Asian adult exposed to the same level of UV
radiation could achieve such a concentration. A dose–
response study in south Asian adults demonstrated that
exposures equivalent to 45 minutes of sunlight three
times per week would achieve 25(OH)D concentrations
of >25 nmol per litre in 94% of individuals, but only 12%
could achieve levels >50 nmol per litre152.

Management
Radius Ulna Tibia Calciopenic rickets
Nutritional rickets. If dietary calcium deficiency is con-
Figure 8 | Radiographic characteristics of rickets. TheNature
radiographic
Reviews changes of rickets
| Disease Primers sidered the primary cause of rickets, calcium supplements
are most evident at the rapidly growing ends of long bones. Thus, the most appropriate
of ≥1,000 mg per day are recommended, as they ensure a
sites for radiography when rickets is suspected are the distal radius and ulna at the wrist
(part a) and the distal femur and proximal tibia at the knee (part b). Typical radiographic more rapid biochemical and radiographic response than
changes in nutritional rickets are widening of the growth plates (arrows), loss of the 500 mg per day 153. Calcium supplements in children with
provisional zone of calcification at the distal metaphyses with fraying of the junction dietary calcium deficiency rickets may need to be contin-
between the growth plate and the metaphysis and splaying of the metaphyses. The ued for >6 months to ensure radiographic healing of the
metaphyses of the long bones typically show coarsening of the trabecular pattern with growth plate153. Nutritional rickets associated primarily
apparent increase in radiolucency. Note that identifying osteopenia or low bone mineral with vitamin D deficiency is managed by correcting defi-
content is difficult and subjective, as it is dependent on the quality and contrast of the ciencies in both vitamin D and calcium; thus, recommen-
radiograph. Typically, mineralization of the carpal bones and development of epiphyses dations include both vitamin D supplementation (TABLE 2)
are delayed.
and ensuring an adequate diet­ary calcium intake15. The
vitamin D and calcium status of children with nutritional
Fortification of staple foods with vitamin D is another rickets should be maintained after treatment by ensur-
preventive strategy that has been used to v­ arying degrees ing sufficient vitamin D and calcium intake and/or by
around the world. A systematic review and meta-­analysis increasing sunshine exposure.
of trials in adults showed that food fortification can safely Vitamin D supplements may be given as a single bolus
reduce the prevalence of vitamin D deficiency 144. A ran­ dose or as daily therapy, as outlined in TABLE 2. There is
domized controlled trial of vitamin D fortification of milk no evidence of the clinical superiority of supplemen-
in Indian schoolchildren showed that those consuming tation with vitamin D3 over vitamin D2 when admin-
milk that contained vitamin D were able to achieve istered daily 154, but when administered as bolus doses,
satisfactory 25(OH)D levels of >50 nmol per litre145. vitamin D3 is more effective than D2 in raising 25(OH)
Fortification of milk and margarine in Finland was estim­ D levels154. Single-day high-dose vitamin D therapy
ated to increase daily vitamin D intakes from 176 IU (stoss therapy) of up to 600,000 IU has been suggested,
to 360 IU and to raise 25(OH)D levels by 10 nmol per but owing to the risk of hypercalcaemia155,156, an inter-
litre146. There is considerable variation worldwide in food national consensus group recommends more prudent
fortification practice, with cow milk and some fruit juices doses that have been shown to be as effective in healing
being fortified with vitamin D in the United States and rickets without the risk of vitamin D toxicity 15. Although
Canada but with fortification limited to margarine, some daily therapy provides a more stable 25(OH)D level
breakfast cereals and yoghurts in the United Kingdom147. during the supplementation period than stoss therapy,
In considering the fortification of staple foods within a non-adherence to the treatment regimen may result in a
country, careful attention must be paid to the variation in failure of or delay in radiographic healing of the growth
staple food consumption across targeted communities to plate. Therapy should be continued for 8–12 weeks but
ensure the widest possible consumption, while ­preventing may need to be continued longer if radio­graphic healing
toxic effects in some members of a community. is not complete or serum ­biochemistry has not returned
To provide consistent simple advice regarding sun- completely to normal157.
light exposure to prevent rickets in children of different
ages, geographical locations and skin types is impossible. 25‑Hydroxylase deficiency. The optimal dose of vita-
In fact, achieving satisfactory vitamin D status should min D supplementation seems to vary in patients with
not depend on sunlight exposure alone148. The concept 25‑hydroxylase deficiency depending on whether one or
of ‘safe sunlight exposure’ is often provided in vita- both alleles of CYP2R1 are affected by loss‑of‑function
min D guidelines but is rarely precisely defined. Some mutations, although these observations are based on
studies have provided information on the relationship few patients54,55,158,159. Patients who are heterozygous are

NATURE REVIEWS | DISEASE PRIMERS VOLUME 4 | ARTICLE NUMBER 17101 | 13


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

oral administration of physiological replacement doses


of 1,25(OH)2D (10–400 ng kg–1 per day) or alfacalcidol
(80–100 ng kg–1 per day)161. Response to 1,25(OH)2D
or alfacalcidol is rapid, with healing of rickets and nor-
malization of the biochemistry within 7–9 weeks162. The
maintenance dose is lower than that needed to initiate
healing of rickets, but therapy is lifelong and adherence
might be an issue163,164. Monitoring of serum calcium
and PTH levels and urinary calcium levels is required
regularly to ensure that values are maintained within the
normal range. The disease shows an autosomal recessive
inheritance, and thus genetic counselling of the parents
and the offspring needs to be undertaken.

Hereditary 1,25(OH)2D-resistant rickets. Success of


the management of autosomal recessive hereditary
1,25(OH)2D-resistant rickets varies depending on the
VDR mutation165. Mutations in the DNA binding domain
cause almost total 1,25(OH)2D resistance and alope-
cia, whereas mutations in the ligand-binding domain
cause variable resistance to 1,25(OH)2D. Patients who
have mutations that interfere with heterodimerization
of the VDR also have alopecia. Patients without alo-
pecia are generally more responsive to treatment than
patients with alopecia68. Some patients without alopecia
responded to treatment with 25(OH)D3 in doses rang-
ing from 20 to 200 μg per day and 1,25(OH)2D in doses
of 17–20 μg per day 63; the latter dose is approximately
20‑fold that needed to reverse hypocalcaemia in patients
Figure 9 | Radiograph of metaphyseal
Nature Reviewschondrodysplasia.
| Disease Primers with 1α‑hydroxylase deficiency. For patients who are
Radiograph of the femur of a child with metaphyseal refractory to treatment with activated vitamin D, intra-
chondrodysplasia, a rare genetic condition; these venous administration of large amounts of elemental
abnormalities resemble those of rickets.
calcium (400–1400 mg/m2 per day) for periods of up to
3.8 years was associated with resolution of bone pain,
able to increase circulating 25(OH)D levels in response normalization of biochemical findings of rickets166–170,
to a vitamin D bolus, but the achieved levels are lower radiographic and histologic healing of skeletal lesions166
than those achieved in controls. Patients who are homo­ and increased growth velocity 166,167. Oral administration
zygous for mutations in CYP2R1 show a very limited or of calcium also has been effective in such patients171,172.
no increase in 25(OH)D levels in response to the same The alopecia, if present, does not respond to treatment
vitamin D bolus54. Consequently, clinical, biochemical with calcium or vitamin D metabolites.
and radiographic changes might normalize in patients
who are heterozygous when receiving 5,000–10,000 IU Phosphopenic rickets
of vitamin D per day, whereas a homozygous patient The primary goal of therapy is to improve rickets.
might need a 600,000 IU dose every 3 months. It is of Initially, improvement of growth plate abnormal-
interest that affected children can respond radiograph- ities should be aimed for, which will set the stage for
ically to calcium supplements alone, despite very little the chronic correction of bow deformities of the bone
change in the perturbed biochemistry 54. Studies using and growth enhancement. Conventional treatment of
25(OH)D in the management of this genetic disorder FGF23‑dependent types of rickets includes a combina-
have not been reported. tion of phosphate and activated vitamin D. As XLH is
the most frequent type of FGF23‑dependent rickets, we
1α‑Hydroxylase deficiency. Symptoms and signs of focus on this condition. In phosphopenic rickets inde-
1α‑hydroxylase deficiency generally appear in the second pendent of FGF23, 1,25(OH)2D production is usually
6 months of life160. A high index of suspicion should be not impaired owing to primary defects specific to renal
raised if an infant with features of vitamin D deficiency phosphate reabsorption, and treatment usually does not
does not respond rapidly to vitamin D therapy. Before include activated forms of vitamin D.
starting definitive therapy with activated vitamin D
(that is, 1,25(OH)2D or the monohydroxylated analogue Children with XLH. Current management of XLH
alfacalcidol (1α(OH)D)), diagnosis of 1α‑hydroxylase involves the combination of phosphate supplements
deficiency should be confirmed by measuring serum and active vitamin D. Although this approach has been
levels of 25(OH)D and 1,25(OH)2D (TABLE 1). Currently a mainstay for the treatment of XLH since the early
recommended therapy for 1α‑hydroxylase deficiency is 1980s, it is cumbersome, does not completely correct the

14 | ARTICLE NUMBER 17101 | VOLUME 4 www.nature.com/nrdp


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

skeletal abnormalities and is associated with considerable serum and urinary calcium175. Parathyroidectomy may
toxicity 173. Even with optimal dosage of phosphate and be required if hyper­parathyroidism with hypercalcaemia
activated vitamin D, many patients require orthopaedic is persistent and unresponsive to medical therapy.
surgical intervention, and an experienced p ­ aediatric Surgery is indicated for severe bowing or tibial tor-
orthopaedic surgeon is invaluable to the care team. sion unlikely to improve with medical management
The primary goal of XLH therapy is to improve skele­ alone176. Corrective osteotomies (a surgical procedure
tal abnormalities and growth, not serum phosphate in which a bone is cut to straighten it) are not usually
levels. Moreover, it is not necessary to increase serum performed in children aged <6 years, as medical therapy
phosphorus levels into the normal range to ensure often improves bow deformities at this age. Osteotomies
healing of rickets; to do so requires administration of are generally performed when growth has nearly ceased,
large doses of phosphate, increasing the risk of hyper­ but severe deformities may require earlier intervention.
parathyroidism and nephrocalcinosis and m ­ ineralization Less-invasive approaches include epiphysiodesis, a pro-
of other soft tissues. cedure that induces corrective differential growth of the
Typical dosages of 1,25(OH)2D are 20–30 ng kg–1 per growth plate177. Many children are prone to recurrent
day given in 2–3 individual doses together with phos- dental abscesses178. Frequent brushing and regular ­dental
phate (20–40 mg kg–1 per day)173. The two medications hygienist visits are recommended.
must be administered at balanced dosages and moni-
tored carefully; relative excess of 1,25(OH)2D may lead Adults with XLH. Symptomatic adults with XLH may
to hypercalcaemia and hypercalciuria, whereas relative benefit from therapy; the primary goal of treatment is
excess of phosphate may result in secondary hyper- reducing pain. Patients are selected for treatment on
parathyroidism or, if persistent, in parathyroid gland the basis of the presence of spontaneous insufficiency
hyperplasia. Dose adjustments are needed when these fractures, disabling skeletal pain and/or biochemical
abnormalities are encountered. Consequently, hyper- evidence of osteomalacia (such as increased serum
calcaemia or hypercalciuria are indications to reduce alkaline phosphatase levels). Therapy in adulthood
the 1,25(OH)2D dose. Increased PTH levels may be also uses oral phosphate salts and activated vitamin D
corrected by either increasing the 1,25(OH)2D dose or metabolites. The therapy does not seem to influence
reducing the phosphate dose. The optimal dosages of (for better or worse) the commonly occurring entheso­
phosphate and 1,25(OH)2D are determined by monitor­ pathy (disorder involving the mineralization of the
ing treatment effectiveness. Height, degree of skeletal attachments of tendons or ligaments to bone) or osteo­
deformity and radiographic healing of the growth plate phyte formation (bony spurs or outgrowths localized on
regions necessitate periodic adjustments in the dose. the joint margins)179. Doses of 1,25(OH)2D are usually
Radiographs of the distal femurs and proximal tibias started at 0.75–1 μg per day administered once or twice
should be performed at least every 2 years. Renal ultra- daily. Phosphate salts are added with gradual increase to
sonography may be intermittently performed to moni- approximately 750–1,000 mg of elemental phosphorus
tor the development or progression of nephrocalcinosis, daily given in 3–4 doses. As with children, doses may
which would warrant decreasing treatment dosages. change abruptly during the course of therapy. Patients
If persistent hyperparathyroidism is not success­fully must be agreeable to follow‑up evaluations and moni­
managed with careful adjustments of the dosing regi­ tored for toxicity. Orthopaedic surgery is likely to
men, the phosphate or both, medications may need more effectively heal when the skeleton is optimally
to be discontinued. Further approaches have included mineral­ized than in the presence of osteomalacia. Thus,
the administration of cinacalcet (a calcimimetic) after planning for elective orthopaedic procedures should
discontinuation of 1,25(OH) 2D and phosphate 174. include the consideration of an ~6‑month window of
Cinacalcet has been variably tolerated in our experience, medical treatment before surgery. Further details of this
and rebound hyperparathyroidism following discontinu- ­treatment strategy are described elsewhere173.
ation of the drug may occur. We have also used paricalci-
tol (a 1,25(OH)2D analogue) in the setting of intolerance Emerging strategies. The identification of excess FGF23
or unresponsiveness to cinacalcet with monitoring of as the cause of renal phosphate wasting in XLH has prov­
ided a sound basis for novel therapeutic strategies aimed
at lowering FGF23 levels180. Monthly injections of an
Table 2 | Treatment doses of vitamin D for nutritional rickets anti‑FGF23 monoclonal antibody (burosumab) in adults
Age Daily dose for Single dose (IU) Maintenance with XLH effectively restore serum phosphate and
90 days (IU)* daily dose (IU) 1,25(OH)2D levels181, and biweekly injections in children
<3 months 2,000 NA 400 have induced similar biochemical findings, improved
radiographic parameters of rickets and a favourable safety
3–12 months 2,000 50,000 400
profile182. A randomized placebo-controlled study (pre-
1–12 years 3,000–6,000 150,000 600 sented as a meeting abstract) in adults with XLH has con-
>12 years 6,000 300,000 600 firmed the effects on serum phosphate and 1,25(OH)2D
IU, international units; NA; not available. Ensure a daily calcium intake of at least 500 mg. levels with improved rates of fracture healing evident in
*Reassess to verify decreasing serum alkaline phosphatase level and/or radiographic evidence the burosumab-treated group compared with placebo183.
of healing rickets after 90 days, as further treatment may be required. Republished with
permission of the Endocrine Society, from Munns, C. F. et al. J. Clin. Endocrinol. Metab. 101, At the time of this writing, ­burosumab is not yet approved
394–415 (2016). for widespread clinical use.

NATURE REVIEWS | DISEASE PRIMERS VOLUME 4 | ARTICLE NUMBER 17101 | 15


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

Other FGF23‑dependent types of hypophosphataemic methods, 84% of adults with XLH were found to have
rickets. FGF23‑dependent types of rickets other than impaired QOL188. Increasing age, the presence of struc-
XLH (BOX 1) are generally treated in a manner similar tural lesions (for example, enthesopathy, osteo­arthritis
to that for XLH. An important consideration has arisen or insufficiency fractures) and dental defects were the
in the context of ADHR184. As iron deficiency stimu- strongest predictors of reduced QOL. Interestingly,
lates FGF23 production185, individuals with ADHR are QOL was more significantly affected in women than in
susceptible to worsening disease in the setting of iron men with XLH, which is an X‑linked dominant condi-
deficiency owing to their inherent limitation to degrade tion, a finding that was not specifically explored in this
FGF23. Thus, identifying and correcting iron defi- study. Treatment with phosphate supplementation and
ciency in patients with ADHR are important manage­ activated vitamin D was significantly associated with
ment considerations. Tumour-induced osteomalacia improved mental health but not physical functioning
best responds to removal of the tumour if it can be scores188. Compared with patients with axial spondylo­
approached surgically; if surgery cannot be performed, arthritis, who also have structural bone lesions, those
therapy with phosphate and activated vitamin D is rec- with XLH had lower QOL scores, likely because 75%
ommended. Recent studies have also indicated that of the patients with axial spondyloarthritis received
burosumab may be an alternative strategy for those ­effective treatment for their disorder 188.
with inoperable or non-localized tumours186, although The current standard of care (phosphate and activated
it is currently available only as an investigational agent. vitamin D therapy) only partially improves XLH190, but
In epidermal nevus syndrome, skin resection of affected data from clinical trials suggest that burosumab180,181
lesions has had variable transient results, but burosumab might improve QOL in XLH. Following a monthly
may also be useful in this condition, as well as activated dose of burosumab for 4 months, the patients reported
vitamin D and phosphate. Curiously, hypophospha­ improved perception of their chronic functional impair-
taemia seems to wane somewhat during early adulthood ments related to poor physical health, physical function-
in patients with epidermal nevus syndrome. ing and stiffness scores189. Controlled trials investigating
the long-term effects of burosumab on QOL are ongoing.
FGF23‑independent types of hypophosphataemic rickets. The ultimate proof of treatment efficacy in any setting
­HHRH is treated with phosphate alone, in view of the is improved QOL; as such, these preliminary data for
normal‑to‑high serum levels of 1,25(OH)2D in this patients with XLH are encouraging and serve as a model
condition103. Treatment with phosphate is targeted at for studying QOL outcomes in other types of rickets.
decreasing the 1,25(OH)2D levels to effect a reduction
in intestinal calcium absorption and hypercalciuria Outlook
while also improving skeletal mineralization. To further Nutritional deficiencies of vitamin D and dietary cal-
decrease the risk of nephrocalcinosis and nephrolithiasis, cium are the most common causes of rickets worldwide.
generous oral hydration to lower urine calcium concen- These disorders still account for considerable morbid-
tration is recommended and diets with a high sodium ity, particularly in poorly resourced countries, where the
content should be avoided. In phosphopenic rickets sec- disease often presents late with established deformities
ondary to tubulopathies, Fanconi syndrome, Dent dis- unlikely to be reversed by therapy. Thus, effective ways
ease or other systemic diseases, specific treatment of the to implement preventive strategies are key. With increas-
underlying disease, in addition to cautious phosphate and ing use of prepackaged food, dietary fortification holds
active vitamin D supplementation, is important. great promise in many countries and should be an inter-
national priority. However, the diversity of staple foods
Quality of life between different countries makes it difficult to establish
Limb deformities and pain can considerably impair universal recommendations.
quality of life (QOL) of children with rickets and adults At the same time, the chronic nature of heritable types
with osteomalacia. Ultimately, rickets has the poten- of rickets requires attention. Lifelong management should
tial to impede activities of daily living owing to bone be optimized, as relapse often results in consider­able
pain, muscle weakness and, in the case of phosphopenic morbidity. Discovery of the mutations causing herit­able
rickets, pain and reduction of physical function due to types of rickets has led to detailed understanding of the
enthesopathy, arthritis and spinal stenosis187. Bone pain mechanisms of these disorders and continues to provide
is typically the worst in those with profound lower limb a basis for the development of new therapies. A number
deformity. Dental abnormalities can also negatively of issues remain poorly understood, including the mech-
affect QOL187. The burden to the individual and soci- anisms by which loss of function of PHEX (in XLH) gives
ety remains greatest in developing countries, where the rise to excess production of FGF23 by osteocytes. It is
incidence of nutritional rickets remains particularly high not known whether the mineralization defect in XLH is
and where fewer social networks exist to support those due primarily to FGF23‑mediated hypophosphataemia
who struggle with physical disabilities2. and suppression of 1,25(OH)2D production or whether
There are limited studies that have specifically accumu­lation of ASARM peptides, pyrophosphate or
quanti­fied the impact of rickets on QOL beyond reports other factors have a critical role. In addition, the basis for
of the typical clinical manifestations of the disease. the development of many musculoskeletal complications
Those that have done so arise from studies in adults with of XLH, such as enthesopathy, osteophyte formation and
XLH188,189. In one study that used validated QOL scoring arthritis, remains poorly understood.

16 | ARTICLE NUMBER 17101 | VOLUME 4 www.nature.com/nrdp


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

With the understanding that excess FGF23 is the degradation192 and use of the carboxy-­terminal FGF23
humoral basis for renal phosphate wasting in FGF23‑ fragment to inhibit binding of the active full-length
dependent types of rickets, attention has focused molecule193. As our understanding of the pathogenesis
on newer strategies that directly inhibit the renal of rickets and osteomalacia in the hypophospha­taemic
actions of FGF23, such as a recombinant human mono­ syndromes increases, more appropriate therapies
clonal antibody that binds to FGF23, thereby block- will likely become available. However, as with the
ing its biological activity. Other potential approaches nutritional deficiencies, it is important that these
explored in mouse models include inhibition of down- therapies be available globally, including in poorly
stream FGFR signalling 191, acceleration of FGF23 resourced countries.

1. Shore, R. M. & Chesney, R. W. Rickets: Part I. Pediatr. 16. Prentice, A. Nutritional rickets around the world. vitamin D deficiency (25(OH)D levels of <50 nmol
Radiol. 43, 140–151 (2012). J. Steroid Biochem. Mol. Biol. 136, 201–206 (2013). per litre). Vitamin D deficiency was higher in urban
2. Creo, A. L., Thacher, T. D., Pettifor, J. M., 17. við Streym, S. et al. Vitamin D content in human settings, and lack of sunlight, higher body fat,
Strand, M. A. & Fischer, P. R. Nutritional rickets breast milk: a 9‑mo follow‑up study. Am. J. Clin. Nutr. availability of television or computers at home,
around the world: an update. Paediatr. Int. Child 103, 107–114 (2016). maternal education and low socioeconomic status
Health 37, 84–98 (2016). 18. Callaghan, A. L., Moy, R. J. D., Booth, I. W., were key predictors.
This paper discusses the global epidemiology of Debelle, G. & Shaw, N. J. Incidence of symptomatic 36. Bikle, D. D. Vitamin D metabolism, mechanism of
rickets, including the key risk factors for rickets in vitamin D deficiency. Arch. Dis. Child. 91, 606–607 action, and clinical applications. Chem. Biol. 21,
different parts of the world. The key relationship (2005). 319–329 (2014).
between calcium deficiency and vitamin D 19. Abrams, S. A. Nutritional rickets: an old disease 37. Tenenhouse, H. S. et al. Renal Na+-phosphate
deficiency in the aetiology of rickets and how the returns. Nutr. Rev. 60, 111–115 (2002). cotransport in murine X‑linked hypophosphatemic
relative importance of these aetiologies varies by 20. Mitchell, S. M. et al. High frequencies of elevated rickets. Molecular characterization. J. Clin. Invest. 93,
location and age are emphasized. alkaline phosphatase activity and rickets exist in 671–676 (1994).
3. Chick, H. Study of rickets in Vienna 1919–1922. extremely low birth weight infants despite current 38. Carpenter, T. O. in www.endotext.org (eds DeGroot, L.
Med. Hist. 20, 41–51 (1976). nutritional support. BMC Pediatr. 9, 47 (2009). & Singer, F.) (MDText.com, Inc., 2014).
4. O’Riordan, J. L. H. & Bijvoet, O. L. M. Rickets before 21. Wagner, C. L. & Greer, F. R., American Academy of 39. Perwad, F., Zhang, M. Y. H., Tenenhouse, H. S.
the discovery of vitamin D. Bonekey Rep. 3, 478 Pediatrics Section on Breastfeeding & American & Portale, A. A. Fibroblast growth factor 23 impairs
(2014). Academy of Pediatrics Committee on Nutrition. phosphorus and vitamin D metabolism in vivo and
5. Thacher, T. D. et al. Increasing incidence of nutritional Prevention of rickets and vitamin D deficiency in suppresses 25‑hydroxyvitamin D-1 -hydroxylase
rickets: a population-based study in Olmsted County. infants, children, and adolescents. Pediatrics 122, expression in vitro. Am. J. Physiol. Renal Physiol.
Minnesota. Mayo Clin. Proc. 88, 176–183 (2013). 1142–1152 (2008). 293, F1577–F1583 (2007).
This epidemiological study demonstrates that 22. Kooh, S. W. et al. Rickets due to calcium deficiency. 40. Sabbagh, Y. et al. Intestinal npt2b plays a major role
rickets has increased in frequency over the past N. Engl. J. Med. 297, 1264–1266 (1977). in phosphate absorption and homeostasis. J. Am. Soc.
35 years and that incidences are greatest in 23. Dagnelie, P. C. et al. High prevalence of rickets in Nephrol. 20, 2348–2358 (2009).
breastfed, African-American infants and low infants on macrobiotic diets. Am. J. Clin. Nutr. 51, 41. Meyer, R. A., Meyer, M. H., Gray, R. W. & Bruns, M. E.
birthweight infants. 202–208 (1990). Evidence that low plasma 1,25‑dihydroxyvitamin D
6. Goldacre, M., Hall, N. & Yeates, D. G. R. 24. Abrams, S. A. Calcium and vitamin D requirements causes intestinal malabsorption of calcium and
Hospitalisation for children with rickets in England: of enterally fed preterm infants. Pediatrics 131, phosphate in juvenile X‑linked hypophosphatemic
a historical perspective. Lancet 383, 597–598 (2014). e1676–e1683 (2013). mice. J. Bone Miner. Res. 2, 67–82 (1987).
This paper highlights the rising rate of hospital 25. McIntosh, N., Livesey, A. & Brooke, O. G. Plasma 42. Fukumoto, S. et al. Pathogenesis and diagnostic
admissions for rickets in England over the past 25‑hydroxyvitamin D and rickets in infants of criteria for rickets and osteomalacia — proposal by an
20 years. Although global data providing a extremely low birthweight. Arch. Dis. Child. 57, expert panel supported by Ministry of Health, Labour
population-based rate are limited, those that are 848–850 (1982). and Welfare, Japan, The Japanese Society for Bone
available support the contention of an increasing 26. Martin, J. A., Hamilton, B. E., Osterman, M. J. K., and Mineral Research and The Japan Endocrine
prevalence of rickets in a number of countries over Driscoll, A. K. & Mathews, T. J. National Vital Society. Endocr. J. 62, 665–671 (2015).
the past few decades. Statistics Reports, Volume 66, Number 1 (CDC, 43. Sabbagh, Y., Carpenter, T. O. & Demay, M. B.
7. Bener, A. & Hoffmann, G. Nutritional rickets among 2017). Hypophosphatemia leads to rickets by impairing
children in a sun rich country. Int. J. Pediatr. 27. Holm, I. A. et al. Mutational analysis and genotype– caspase-mediated apoptosis of hypertrophic
Endocrinol. 2010, 410502 (2010). phenotype correlation of the PHEX gene in X‑linked chondrocytes. Proc. Natl Acad. Sci. USA 102,
8. Uush, T. Prevalence of classic signs and symptoms hypophosphatemic rickets. J. Clin. Endocrinol. Metab. 9637–9642 (2005).
of rickets and vitamin D deficiency in Mongolian 86, 3889–3899 (2001). 44. Carmeliet, G., Dermauw, V. & Bouillon, R. Vitamin D
children and women. J. Steroid Biochem. Mol. Biol. 28. Endo, I. et al. Nationwide survey of fibroblast growth signaling in calcium and bone homeostasis: a delicate
136, 207–210 (2013). factor 23 (FGF23)-related hypophosphatemic diseases balance. Best Pract. Res. Clin. Endocrinol. Metab. 29,
9. Holick, M. F. Resurrection of vitamin D deficiency and in Japan: prevalence, biochemical data and treatment. 621–631 (2015).
rickets. J. Clin. Invest. 116, 2062–2072 (2006). Endocr. J. 62, 811–816 (2015). 45. Oramasionwu, G. E., Thacher, T. D., Pam, S. D.,
10. Rajakumar, K., Greenspan, S. L., Thomas, S. B. & 29. Hazzazi, M., Alzeer, I., Tamimi, W., Al Atawi, M. Pettifor, J. M. & Abrams, S. A. Adaptation of calcium
Holick, M. F. SOLAR ultraviolet radiation and & Al Alwan, I. Clinical presentation and etiology of absorption during treatment of nutritional rickets in
vitamin D. Am. J. Public Health 97, 1746–1754 osteomalacia/rickets in adolescents. Saudi J. Kidney Nigerian children. Br. J. Nutr. 100, 387–392 (2008).
(2007). Dis. Transplant. 24, 938–941 (2013). 46. Thacher, T. D., Obadofin, M. O., O’Brien, K. O. &
11. Ward, L. M., Gaboury, I., Ladhani, M. & Zlotkin, S. 30. Ward, K. A. et al. A randomized, controlled trial of Abrams, S. A. The effect of vitamin D2 and vitamin D3
Vitamin D‑deficiency rickets among children in vitamin D supplementation upon musculoskeletal on intestinal calcium absorption in Nigerian children
Canada. Can. Med. Assoc. J. 177, 161–166 (2007). health in postmenarchal females. J. Clin. Endocrinol. with rickets. J. Clin. Endocrinol. Metab. 94,
12. Wheeler, B. J., Dickson, N. P., Houghton, L. A., Metab. 95, 4643–4651 (2010). 3314–3321 (2009).
Ward, L. M. & Taylor, B. J. Incidence and 31. Agarwal, A. & Gulati, D. Early adolescent nutritional 47. Aggarwal, V. et al. Role of calcium deficiency in
characteristics of vitamin D deficiency rickets in New rickets. J. Orthop. Surg. (Hong Kong) 17, 340–345 development of nutritional rickets in Indian children:
Zealand children: a New Zealand Paediatric (2009). a case control study. J. Clin. Endocrinol. Metab. 97,
Surveillance Unit study. Aust. N. Z. J. Public Health 32. Palacios, C. & Gonzalez, L. Is vitamin D deficiency 3461–3466 (2012).
39, 380–383 (2015). a major global public health problem? J. Steroid 48. Pettifor, J. M. Calcium and vitamin D metabolism in
13. Munns, C. F. et al. Incidence of vitamin D deficiency Biochem. Mol. Biol. 144, 138–145 (2014). children in developing countries. Ann. Nutr. Metab.
rickets among Australian children: an Australian 33. Habibesadat, S., Ali, K., Shabnam, J. M. & Arash, A. 64, 15–22 (2014).
Paediatric Surveillance Unit study. Med. J. Aust. 196, Prevalence of vitamin D deficiency and its related 49. Prentice, A., Ceesay, M., Nigdikar, S., Allen, S. J.
466–468 (2012). factors in children and adolescents living in North & Pettifor, J. M. FGF23 is elevated in Gambian
14. Beck-Nielsen, S. S., Brock-Jacobsen, B., Gram, J., Khorasan, Iran. J. Pediatr. Endocrinol. Metab. 27, children with rickets. Bone 42, 788–797 (2008).
Brixen, K. & Jensen, T. K. Incidence and prevalence 431–436 (2014). 50. Braithwaite, V., Jarjou, L. M. A., Goldberg, G. R.
of nutritional and hereditary rickets in southern 34. Dahifar, H., Faraji, A., Ghorbani, A. & Yassobi, S. & Prentice, A. Iron status and fibroblast growth
Denmark. Eur. J. Endocrinol. 160, 491–497 (2008). Impact of dietary and lifestyle on vitamin D in healthy factor‑23 in Gambian children. Bone 50, 1351–1356
15. Munns, C. F. et al. Global consensus recommendations student girls aged 11–15 years. J. Med. Investig. 53, (2012).
on prevention and management of nutritional rickets. 204–208 (2006). 51. Cheng, J. B., Motola, D. L., Mangelsdorf, D. J.
J. Clin. Endocrinol. Metab. 101, 394–415 (2016). 35. Wakayo, T., Belachew, T., Vatanparast, H. & Russell, D. W. De‑orphanization of cytochrome
This paper is a consensus document drawn up by & Whiting, S. J. Vitamin D deficiency and its predictors P450 2R1. J. Biol. Chem. 278, 38084–38093
experts in the field of nutritional rickets on the in a country with thirteen months of sunshine: the (2003).
prevention and management of nutritional rickets case of school children in central Ethiopia. PLoS ONE 52. Zhu, J. G., Ochalek, J. T., Kaufmann, M., Jones, G.
globally. Particular attention is paid to making the 10, e0120963 (2015). & DeLuca, H. F. CYP2R1 is a major, but not exclusive,
recommendations appropriate and relevant for This study shows that almost half of adolescents contributor to 25‑hydroxyvitamin D production in vivo.
poorly resourced countries. aged 11–18 years living in Ethiopia had prevalent Proc. Natl Acad. Sci. USA 110, 15650–15655 (2013).

NATURE REVIEWS | DISEASE PRIMERS VOLUME 4 | ARTICLE NUMBER 17101 | 17


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

53. Cheng, J. B., Levine, M. A., Bell, N. H., follicle cycling deficiency in Vdr knockout mice. 101. Schlingmann, K. P. et al. Mutations in CYP24A1 and
Mangelsdorf, D. J. & Russell, D. W. Genetic evidence J. Cell. Physiol. 225, 482–489 (2010). idiopathic infantile hypercalcemia. N. Engl. J. Med.
that the human CYP2R1 enzyme is a key vitamin D 77. Bikle, D. D. Vitamin D and the skin. J. Bone Miner. 365, 410–421 (2011).
25‑hydroxylase. Proc. Natl Acad. Sci. USA 101, Metab. 28, 117–130 (2010). 102. Tieder, M. et al. Hereditary hypophosphatemic rickets
7711–7715 (2004). 78. Chen, H., Hewison, M. & Adams, J. S. Functional with hypercalciuria. N. Engl. J. Med. 312, 611–617
54. Thacher, T. D., Fischer, P. R., Singh, R. J., Roizen, J. characterization of heterogeneous nuclear ribonuclear (1985).
& Levine, M. A. CYP2R1 mutations impair generation protein C1/C2 in vitamin D resistance: a novel 103. Bergwitz, C. et al. SLC34A3 mutations in patients
of 25‑hydroxyvitamin D and cause an atypical form of response element-binding protein. J. Biol. Chem. 281, with hereditary hypophosphatemic rickets with
vitamin D deficiency. J. Clin. Endocrinol. Metab. 100, 39114–39120 (2006). hypercalciuria predict a key role for the sodium–
E1005–E1013 (2015). 79. Baylink, D., Wergedal, J. & Stauffer, M. phosphate cotransporter NaPi-IIc in maintaining
55. Al Mutair, A. N., Nasrat, G. H. & Russell, D. W. Formation, mineralization, and resorption of bone phosphate homeostasis. Am. J. Hum. Genet. 78,
Mutation of the CYP2R1 vitamin D 25‑hydroxylase in hypophosphatemic rats. J. Clin. Invest. 50, 179–192 (2006).
in a Saudi Arabian family with severe vitamin D 2519–2530 (1971). 104. Lorenz-Depiereux, B. et al. Hereditary
deficiency. J. Clin. Endocrinol. Metab. 97, 80. Schanler, R. J., Abrams, S. A. & Garza, C. hypophosphatemic rickets with hypercalciuria is
E2022–E2025 (2012). Bioavailability of calcium and phosphorus in human caused by mutations in the sodium–phosphate
56. Prader, A., Illig, R. & Heierli, E. Eine besondere form milk fortifiers and formula for very low birth weight cotransporter gene SLC34A3. Am. J. Hum. Genet. 78,
des primare vitamin-D‑resistenten rachitis mit infants. J. Pediatr. 113, 95–100 (1988). 193–201 (2006).
hypocalcamie und autosomal-dominanten erbgang: 81. Abrams, S. A. In utero physiology: role in nutrient 105. Karim, Z. et al. NHERF1 mutations and
die hereditare pseudomangelrachitis [German]. delivery and fetal development for calcium, responsiveness of renal parathyroid hormone. N. Engl.
Helv. Paediatr. Acta 16, 452–468 (1961). phosphorus, and vitamin D. Am. J. Clin. Nutr. 85, J. Med. 359, 1128–1135 (2008).
57. Scriver, C. R. Vitamin D dependency. Pediatrics 45, 604S–607S (2007). 106. Courbebaisse, M. et al. A new human NHERF1
361–363 (1970). 82. Gonzalez Ballesteros, L. F. et al. Unexpected mutation decreases renal phosphate transporter
58. De Braekeleer, M. & Larochelle, J. Population genetics widespread hypophosphatemia and bone disease NPT2a expression by a PTH-independent mechanism.
of vitamin D‑dependent rickets in northeastern associated with elemental formula use in infants PLoS ONE 7, e34764 (2012).
Quebec. Ann. Hum. Genet. 55, 283–290 (1991). and children. Bone 97, 287–292 (2017). 107. Sneddon, W. B. et al. Convergent signaling pathways
59. Miller, W. L. Genetic disorders of Vitamin D 83. Insogna, K. L., Bordley, D. R., Caro, J. F. regulate parathyroid hormone and fibroblast growth
biosynthesis and degradation. J. Steroid Biochem. & Lockwood, D. H. Osteomalacia and weakness from factor‑23 action on NPT2A‑mediated phosphate
Mol. Biol. 165, 101–108 (2017). excessive antacid ingestion. JAMA 244, 2544–2546 transport. J. Biol. Chem. 291, 18632–18642 (2016).
60. Alzahrani, A. S. et al. A novel G102E mutation of (1980). 108. Bai, X. et al. Partial rescue of the Hyp phenotype by
CYP27B1 in a large family with vitamin D‑dependent 84. Kagitani, K. et al. Hypophosphatemic rickets osteoblast-targeted PHEX (phosphate-regulating gene
rickets type 1. J. Clin. Endocrinol. Metab. 95, accompanying congenital microvillous atrophy. with homologies to endopeptidases on the X
4176–4183 (2010). J. Bone Miner. Res. 13, 1946–1952 (1998). chromosome) expression. Mol. Endocrinol. 16,
61. Wang, X. Novel gene mutations in patients with 85. Bonewald, L. F. & Wacker, M. J. FGF23 production 2913–2925 (2002).
1 -hydroxylase deficiency that confer partial enzyme by osteocytes. Pediatr. Nephrol. 28, 563–568 (2012). 109. Liu, S., Guo, R., Tu, Q. & Quarles, L. D.
activity in vitro. J. Clin. Endocrinol. Metab. 87, 86. Wolf, M. & White, K. E. Coupling fibroblast growth Overexpression of Phex in osteoblasts fails to rescue
2424–2430 (2002). factor 23 production and cleavage. Curr. Opin. the Hyp mouse phenotype. J. Biol. Chem. 277,
62. Brooks, M. H., Stern, P. H. & Bell, N. H. Vitamin Nephrol. Hypertens. 23, 411–419 (2014). 3686–3697 (2001).
D‑dependent rickets type II. N. Engl. J. Med. 302, 87. White, K. E. et al. Autosomal-dominant 110. Boskey, A. et al. The PHEX transgene corrects
810 (1980). hypophosphatemic rickets (ADHR) mutations stabilize mineralization defects in 9‑month-old
63. Marx, S. J. et al. A familial syndrome of decrease FGF‑23. Kidney Int. 60, 2079–2086 (2001). hypophosphatemic mice. Calcif. Tissue Int. 84,
in sensitivity to 1,25‑dihydroxyvitamin D. J. Clin. 88. Tagliabracci, V. S. et al. Dynamic regulation of FGF23 126–137 (2008).
Endocrinol. Metab. 47, 1303–1310 (1978). by Fam20C phosphorylation, GalNAc‑T3 glycosylation, 111. Marie, P. J., Travers, R. & Glorieux, F. H. Bone
64. Beer, S. et al. Vitamin D resistant rickets with and furin proteolysis. Proc. Natl Acad. Sci. USA 111, response to phosphate and vitamin D metabolites in
alopecia: a form of end organ resistance to 1,25 5520–5525 (2014). the hypophosphatemic male mouse. Calcif. Tissue Int.
dihydroxy vitamin D. Clin. Endocrinol. (Oxf.) 14, 89. Rafaelsen, S. H. et al. Exome sequencing reveals 34, 158–164 (1982).
395–402 (1981). FAM20c mutations associated with fibroblast growth 112. Marie, P. J. & Glorieux, F. H. Stimulation of cortical
65. Rosen, J. F., Fleischman, A. R., Finberg, L., factor 23‑related hypophosphatemia, dental bone mineralization and remodeling by phosphate
Hamstra, A. & DeLuca, H. F. Rickets with alopecia: anomalies, and ectopic calcification. J. Bone Miner. and 1,25‑dihydroxyvitamin D in vitamin D‑resistant
an inborn error of vitamin D metabolism. J. Pediatr. Res. 28, 1378–1385 (2013). rickets. Metab. Bone Dis. Relat. Res. 3, 159–164
94, 729–735 (1979). 90. Takeyari, S. et al. Hypophosphatemic osteomalacia and (1981).
66. Hochberg, Z. et al. 1,25‑dihydroxyvitamin D bone sclerosis caused by a novel homozygous mutation 113. Kaludjerovic, J. et al. Klotho expression in long bones
resistance, rickets, and alopecia. Am. J. Med. 77, of the FAM20C gene in an elderly man with a mild regulates FGF23 production during renal failure.
805–811 (1984). variant of Raine syndrome. Bone 67, 56–62 (2014). FASEB J. 31, 2050–2064 (2017).
67. Hochberg, Z. Calcitriol-resistant rickets with alopecia. 91. Riminucci, M. et al. FGF‑23 in fibrous dysplasia of 114. Murali, S. K., Andrukhova, O., Clinkenbeard, E. L.,
Arch. Dermatol. 121, 646–647 (1985). bone and its relationship to renal phosphate wasting. White, K. E. & Erben, R. G. Excessive osteocytic Fgf23
68. Marx, S. J., Bliziotes, M. M. & Nanes, M. Analysis J. Clin. Invest. 112, 683–692 (2003). secretion contributes to pyrophosphate accumulation
of the relation between alopecia and resistance to 92. Smith, R. C. et al. Circulating αKlotho influences and mineralization defect in Hyp mice. PLoS Biol. 14,
1,25‑dihydroxyvitamin D. Clin. Endocrinol. (Oxf.) 25, phosphate handling by controlling FGF23 production. e1002427 (2016).
373–381 (1986). J. Clin. Invest. 122, 4710–4715 (2012). 115. Addison, W. N., Nakano, Y., Loisel, T., Crine, P.
69. Haussler, M. R. et al. The nuclear vitamin D receptor: 93. Brownstein, C. A. et al. A translocation causing & McKee, M. D. MEPE-ASARM peptides control
biological and molecular regulatory properties increased alpha-klotho level results in extracellular matrix mineralization by binding to
revealed. J. Bone Miner. Res. 13, 325–349 (1998). hypophosphatemic rickets and hyperparathyroidism. hydroxyapatite: an inhibition regulated by PHEX
70. Bouillon, R. et al. Vitamin D and human health: Proc. Natl Acad. Sci. USA 105, 3455–3460 (2008). cleavage of ASARM. J. Bone Miner. Res. 23,
lessons from vitamin D receptor null mice. Endocr. 94. Lee, J.‑C. et al. Characterization of FN1–FGFR1 and 1638–1649 (2008).
Rev. 29, 726–776 (2008). novel FN1–FGF1 fusion genes in a large series of 116. Martin, A. et al. Degradation of MEPE, DMP1, and
71. Malloy, P. J., Xu, R., Peng, L., Clark, P. A. phosphaturic mesenchymal tumors. Mod. Pathol. 29, release of SIBLING ASARM-peptides (minhibins):
& Feldman, D. A. Novel mutation in helix 12 of the 1335–1346 (2016). ASARM-Peptide(s) are directly responsible for
vitamin D receptor impairs coactivator interaction and 95. White, K. E. et al. Mutations that cause osteoglophonic defective mineralization in HYP. Endocrinology 149,
causes hereditary 1,25‑dihydroxyvitamin D‑resistant dysplasia define novel roles for FGFR1 in bone 1757–1772 (2008).
rickets without alopecia. Mol. Endocrinol. 16, elongation. Am. J. Hum. Genet. 76, 361–367 (2005). 117. Thacher, T. D. in Vitamin D and Rickets Vol. 6 (ed
2538–2546 (2002). 96. Lim, Y. H., Ovejero, D., Derrick, K. M., Collins, M. T. Hochberg, Z.) 105–125 (Karger, 2003).
72. Malloy, P. J., Xu, R., Cattani, A., Reyes, M. L. & Choate, K. A. Cutaneous skeletal hypophosphatemia 118. Basatemur, E. & Sutcliffe, A. Incidence of hypocalcemic
& Feldman, D. A. Unique insertion/substitution in syndrome (CSHS) is a multilineage somatic mosaic seizures due to vitamin D deficiency in children in the
helix H1 of the vitamin D receptor ligand binding RASopathy. J. Am. Acad. Dermatol. 75, 420–427 United Kingdom and Ireland. J. Clin. Endocrinol.
domain in a patient with hereditary (2016). Metab. 100, E91–E95 (2015).
1,25‑dihydroxyvitamin D‑resistant rickets. J. Bone 97. Lim, Y. H. et al. Multilineage somatic activating 119. Maiya, S. et al. Hypocalcaemia and vitamin D
Miner. Res. 19, 1018–1024 (2004). mutations in HRAS and NRAS cause mosaic deficiency: an important, but preventable, cause of
73. Malloy, P. J. et al. Hereditary 1,25‑dihydroxyvitamin D cutaneous and skeletal lesions, elevated FGF23 and life-threatening infant heart failure. Heart 94,
resistant rickets due to a mutation causing multiple hypophosphatemia. Hum. Mol. Genet. 23, 397–407 581–584 (2008).
defects in vitamin D receptor function. Endocrinology (2014). 120. Steichen-Gersdorf, E., Lorenz-Depiereux, B.,
145, 5106–5114 (2004). 98. Shimizu, Y. et al. Hypophosphatemia induced by Strom, T. M. & Shaw, N. J. Early onset hearing loss in
74. Hsieh, J.‑C. et al. Analysis of hairless corepressor intravenous administration of saccharated ferric autosomal recessive hypophosphatemic rickets caused
mutants to characterize molecular cooperation with oxide: another form of FGF23‑related by loss of function mutation in ENPP1. J. Pediatr.
the vitamin D receptor in promoting the mammalian hypophosphatemia. Bone 45, 814–816 (2009). Endocrinol. Metab. 28, 967–970 (2015).
hair cycle. J. Cell. Biochem. 110, 671–686 (2010). 99. Magen, D. et al. A loss‑of‑function mutation in NaPi- 121. Liang, G., Katz, L. D., Insogna, K. L., Carpenter, T. O.
75. Cianferotti, L., Cox, M., Skorija, K. & Demay, M. B. IIa and renal Fanconi’s syndrome. N. Engl. J. Med. & Macica, C. M. Survey of the enthesopathy of
Vitamin D receptor is essential for normal keratinocyte 362, 1102–1109 (2010). X‑linked hypophosphatemia and its characterization
stem cell function. Proc. Natl Acad. Sci. USA 104, 100. Tieder, M. et al. Elevated serum in Hyp mice. Calcif. Tissue Int. 85, 235–246 (2009).
9428–9433 (2007). 1,25‑dihydroxyvitamin D concentrations in siblings 122. Nordblad, M., Graham, F., Mughal, M. Z. &
76. Teichert, A., Elalieh, H. & Bikle, D. Disruption of the with primary Fanconi’s syndrome. N. Engl. J. Med. Padidela, R. Rapid assessment of dietary calcium
hedgehog signaling pathway contributes to the hair 319, 845–849 (1988). intake. Arch. Dis. Child. 101, 634–636 (2015).

18 | ARTICLE NUMBER 17101 | VOLUME 4 www.nature.com/nrdp


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

123. Institute of Medicine. Dietary Reference Intakes for improve vitamin D status in northern European 1,25‑dihydroxyvitamin D‑resistant rickets. J. Clin.
Calcium and Vitamin D (The National Academies children: exploring the merits of vitamin D fortification Endocrinol. Metab. 76, 509–512 (1993).
Press, 2011). and supplementation. J. Nutr. 136, 1130–1134 169. Ma, N. S. et al. Hereditary vitamin D resistant rickets:
124. Holick, M. F. et al. Evaluation, treatment, and (2006). identification of a novel splice site mutation in the
prevention of vitamin D deficiency: an Endocrine 147. Allen, R. E., Dangour, A. D., Tedstone, A. E. vitamin D receptor gene and successful treatment
Society clinical practice guideline. J. Clin. Endocrinol. & Chalabi, Z. Does fortification of staple foods improve with oral calcium therapy. Bone 45, 743–746
Metab. 96, 1911–1930 (2011). vitamin D intakes and status of groups at risk of (2009).
125. Public Health England. SACN vitamin D and health deficiency? A United Kingdom modeling study. 170. Kanakamani, J., Tomar, N., Kaushal, E., Tandon, N.
report. GOV.UK https://www.gov.uk/government/ Am. J. Clin. Nutr. 102, 338–344 (2015). & Goswami, R. Presence of a deletion mutation
publications/sacn-vitamin-d-and-health-report 148. Saraff, V. & Shaw, N. Sunshine and vitamin D. (c.716delA) in the ligand binding domain of the
(2016). Arch. Dis. Child. 101, 190–192 (2015). vitamin D receptor in an Indian patient with
126. Wang, T. J. et al. Common genetic determinants of 149. Specker, B. L., Valanis, B., Hertzberg, V., Edwards, N. vitamin D‑dependent rickets type II. Calcif. Tissue Int.
vitamin D insufficiency: a genome-wide association & Tsang, R. C. Sunshine exposure and serum 86, 33–41 (2009).
study. Lancet 376, 180–188 (2010). 25‑hydroxyvitamin D concentrations in exclusively 171. Tamura, M. et al. Detection of hereditary
127. Walton, R. J. & Bijvoet, O. L. M. Nomogram for breast-fed infants. J. Pediatr. 107, 372–376 (1985). 1,25‑hydroxyvitamin D‑resistant rickets caused by
derivation of renal threshold phosphate concentration. 150. Meena, P. et al. Sunlight exposure and vitamin D uniparental disomy of chromosome 12 using genome-
Lancet 306, 309–310 (1975). status in breastfed infants. Indian Pediatr. 54, wide single nucleotide polymorphism array. PLoS ONE
128. Feng, J. Q. et al. Loss of DMP1 causes rickets and 105–111 (2016). 10, e0131157 (2015).
osteomalacia and identifies a role for osteocytes in 151. Farrar, M. D. et al. Recommended summer sunlight 172. Kinoshita, Y., Ito, N., Makita, N., Nangaku, M.
mineral metabolism. Nat. Genet. 38, 1310–1315 exposure amounts fail to produce sufficient vitamin D & Fukumoto, S. Changes in bone metabolic
(2006). status in UK adults of South Asian origin. Am. J. Clin. parameters following oral calcium supplementation
129. Marie, P. J. & Glorieux, F. H. Relation between Nutr. 94, 1219–1224 (2011). in an adult patient with vitamin D‑dependent rickets
hypomineralized periosteocytic lesions and bone 152. Farrar, M. D. et al. Efficacy of a dose range of type 2A. Endocr. J. 64, 589–596 (2017).
mineralization in vitamin D‑resistant rickets. simulated sunlight exposures in raising vitamin D 173. Carpenter, T. O., Imel, E. A., Holm, I. A., Jan de
Calcif. Tissue Int. 35, 443–448 (1983). status in South Asian adults: implications for targeted Beur, S. M. & Insogna, K. L. A clinician’s guide to
130. Stickler, G. B. et al. Familial bone disease resembling guidance on sun exposure. Am. J. Clin. Nutr. 97, X‑linked hypophosphatemia. J. Bone Miner. Res. 26,
rickets (hereditary metaphysial dysostosis). Pediatrics 1210–1216 (2013). 1381–1388 (2011).
29, 996–1004 (1962). 153. Thacher, T. D. et al. Optimal dose of calcium for This paper written by adult and paediatric
131. Jüppner, H. Jansen’s metaphyseal chondrodysplasia. treatment of nutritional rickets: a randomized clinicians provides the rationale and details of
Trends Endocrinol. Metab. 7, 157–162 (1996). controlled trial. J. Bone Miner. Res. 31, 2024–2031 current treatment approaches for XLH.
132. Kaur, S. & Kulkarni, K. Osteopetrorickets: a twin (2016). 174. Alon, U. S. et al. Calcimimetics as an adjuvant
paradox. BMJ Case Rep. http://dx.doi.org/10.1136/ 154. Tripkovic, L. et al. Comparison of vitamin D2 and treatment for familial hypophosphatemic rickets.
bcr.06.2009.2012 (2010). vitamin D3 supplementation in raising serum Clin. J. Am. Soc. Nephrol. 3, 658–664 (2008).
133. Whyte, M. P. Hypophosphatasia: enzyme replacement 25‑hydroxyvitamin D status: a systematic review and 175. Carpenter, T. O. et al. Effect of paricalcitol on
therapy brings new opportunities and new challenges. meta-analysis. Am. J. Clin. Nutr. 95, 1357–1364 circulating parathyroid hormone in X‑linked
J. Bone Miner. Res. 32, 667–675 (2017). (2012). hypophosphatemia: a randomized, double-blind,
134. Whyte, M. P. et al. Asfotase alfa treatment improves 155. Cesur, Y., Çaksen, H., Gündem, A., Kirimi, E. & placebo-controlled study. J. Clin. Endocrinol. Metab.
survival for perinatal and infantile hypophosphatasia. Odabas¸, D. Comparison of low and high dose of 99, 3103–3111 (2014).
J. Clin. Endocrinol. Metab. 101, 334–342 (2016). vitamin D treatment in nutritional vitamin D deficiency 176. Sharkey, M. S., Grunseich, K. & Carpenter, T. O.
135. Ekbote, V. H. et al. A pilot randomized controlled trial rickets. J. Pediatr. Endocrinol. Metab. 16, 1105–1109 Contemporary medical and surgical management of
of oral calcium and vitamin D supplementation using (2003). X‑linked hypophosphatemic rickets. J. Am. Acad.
fortified laddoos in underprivileged Indian toddlers. 156. Mittal, H. et al. 300,000 IU or 600,000 IU of oral Orthop. Surg. 23, 433–442 (2015).
Eur. J. Clin. Nutr. 65, 440–446 (2011). vitamin D3 for treatment of nutritional rickets: 177. Novais, E. & Stevens, P. M. Hypophosphatemic rickets.
136. Thacher, T. D., Bommersbach, T. J., Pettifor, J. M., a randomized controlled trial. Indian Pediatr. 51, J. Pediatr. Orthop. 26, 238–244 (2006).
Isichei, C. O. & Fischer, P. R. Comparison of limestone 265–272 (2014). 178. Chaussain-Miller, C. et al. Dental abnormalities in
and ground fish for treatment of nutritional rickets in 157. Pearce, S. H. & Cheetham, T. D. Diagnosis and patients with familial hypophosphatemic
children in Nigeria. J. Pediatr. 167, 148–154.e1 management of vitamin D deficiency. BMJ 340, vitamin D‑resistant rickets: prevention by early
(2015). b5664 (2010). treatment with 1‑hydroxyvitamin D. J. Pediatr. 142,
137. Gallo, S. et al. Effect of different dosages of oral 158. Dong, Q. & Miller, W. L. Vitamin D 25‑hydroxylase 324–331 (2003).
vitamin D supplementation on vitamin D status in deficiency. Mol. Genet. Metab. 83, 197–198 (2004). 179. Connor, J. et al. Conventional therapy in adults with
healthy, breastfed infants. JAMA 309, 1785–1792 159. Casella, S. J., Reiner, B. J., Chen, T. C., Holick, M. F. X‑linked hypophosphatemia: effects on enthesopathy
(2013). & Harrison, H. E. A possible genetic defect in and dental disease. J. Clin. Endocrinol. Metab. 100,
This paper is a randomized controlled clinical trial 25‑hydroxylation as a cause of rickets. J. Pediatr. 124, 3625–3632 (2015).
assessing four different doses of vitamin D given to 929–932 (1994). 180. Carpenter, T. O. et al. Randomized trial of
healthy breastfed infants in the first year of life 160. Glorieux, F. H. & Pettifor, J. M. Vitamin D/dietary the anti‑FGF23 antibody KRN23 in X‑linked
and demonstrates that 400 IU daily will achieve a calcium deficiency rickets and pseudo-vitamin D hypophosphatemia. J. Clin. Invest. 124, 1587–1597
25(OH)D concentration of >50 nmol per litre. deficiency rickets. Bonekey Rep. 3, 524 (2014). (2014).
138. Hatun, S¸., Ozkan, B. & Bereket, A. Vitamin D 161. Reade, T. M. et al. Response to crystalline 181. Imel, E. A. et al. Prolonged correction of serum
deficiency and prevention: Turkish experience. 1α‑hydroxyvitamin D3 in vitamin D dependency. phosphorus in adults with X‑linked hypophosphatemia
Acta Paediatr. 100, 1195–1199 (2011). Pediatr. Res. 9, 593–599 (1975). using monthly doses of KRN23. J. Clin. Endocrinol.
This paper reports the impact of a nationwide 162. Delvin, E. E., Glorieux, F. H., Marie, P. J. Metab. 100, 2565–2573 (2015).
vitamin D supplementation programme in reducing & Pettifor, J. M. Vitamin D dependency: This paper details the long-term biochemical
the incidence of nutritional rickets in a region of replacement therapy with calcitriol. J. Pediatr. 99, response to anti‑FGF23 antibody therapy in XLH.
Turkey. 26–34 (1981). 182. Carpenter, T. O. et al. Randomized, open-label, dose-
139. Moy, R. J., McGee, E., Debelle, G. D., Mather, I. 163. Wang, J. T. et al. Genetics of vitamin D 1α‑hydroxylase finding, phase 2 study of KRN23, a human
& Shaw, N. J. Successful public health action to reduce deficiency in 17 families. Am. J. Hum. Genet. 63, monoclonal anti‑FGF23 antibody, in children with
the incidence of symptomatic vitamin D deficiency. 1694–1702 (1998). X‑linked hypophosphatemia (XLH). Presented at
Arch. Dis. Child. 97, 952–954 (2012). 164. Edouard, T. et al. Short- and long-term outcome of Endocrine Society’s 98th Annual Meeting and Expo
140. Umaretiya, P. J. et al. Maternal preferences for patients with pseudo-vitamin D deficiency rickets (2016).
vitamin D supplementation in breastfed infants. treated with calcitriol. J. Clin. Endocrinol. Metab. 96, 183. Insogna, K. A phase 3 randomized, 24 week, double-
Ann. Fam. Med. 15, 68–70 (2017). 82–89 (2011). blind, placebo-controlled study evaluating the efficacy
141. Fatani, T. et al. Differential low uptake of free This paper details the long-term clinical and of burosumab, an anti‑FGF23 antibody, in adults with
vitamin D supplements in preterm infants: the Quebec biochemical response to treatment of X‑linked hypophosphatemia (XLH) [abstract LB‑1159].
experience. BMC Pediatr. 14, 291 (2014). 1α‑hydroxylase deficiency with 1,25(OH)2D. Presented at the Annual Meeting of the American
142. Huynh, J. et al. Vitamin D in newborns. A randomised 165. Chaturvedi, D., Garabedian, M., Carel, J.‑C. Society of Bone and Mineral Research (2017).
controlled trial comparing daily and single oral bolus & Léger, J. Different mechanisms of intestinal calcium 184. Econs, M. J. et al. Autosomal dominant
vitamin D in infants. J. Paediatr. Child Health 53, absorption at different life stages: therapeutic hypophosphataemic rickets is associated with
163–169 (2017). implications and long-term responses to treatment in mutations in FGF23. Nat. Genet. 26, 345–348
143. Hollis, B. W. et al. Maternal versus infant vitamin D patients with hereditary vitamin D‑resistant rickets. (2000).
supplementation during lactation: a randomized Horm. Res. Paediatr. 78, 326–331 (2012). 185. Imel, E. A. et al. Iron modifies plasma FGF23
controlled trial. Pediatrics 136, 625–634 (2015). 166. Balsan, S. et al. Long-term nocturnal calcium differently in autosomal dominant hypophosphatemic
144. Black, L. J., Seamans, K. M., Cashman, K. D. infusions can cure rickets and promote normal rickets and healthy humans. J. Clin. Endocrinol.
& Kiely, M. An updated systematic review and meta- mineralization in hereditary resistance to Metab. 96, 3541–3549 (2011).
analysis of the efficacy of vitamin D food fortification. 1,25‑dihydroxyvitamin D. J. Clin. Invest. 77, 186. Carpenter, T. O. Effects of KRN23, an anti‑FGF23
J. Nutr. 142, 1102–1108 (2012). 1661–1667 (1986). antibody, in patients with tumor induced osteomalacia
145. Khadgawat, R. et al. Impact of vitamin D fortified milk 167. Hochberg, Z., Tiosano, D. & Even, L. Calcium therapy and epidermal nevus syndrome: results from an
supplementation on vitamin D status of healthy school for calcitriol-resistant rickets. J. Pediatr. 121, ongoing phase 2 study. Presented at the Annual
children aged 10–14 years. Osteoporos. Int. 24, 803–808 (1992). Meeting of the American Society of Bone and Mineral
2335–2343 (2013). 168. Yagi, H. et al. A new point mutation in the Research (2016).
146. Tylavsky, F. A., Cheng, S., Lyytikäinen, A., deoxyribonucleic acid-binding domain of the 187. Reid, I. R. et al. X‑Linked hypophosphatemia.
Viljakainen, H. & Lamberg-Allardt, C. Strategies to vitamin D receptor in a kindred with hereditary Medicine (Baltimore) 68, 336–352 (1989).

NATURE REVIEWS | DISEASE PRIMERS VOLUME 4 | ARTICLE NUMBER 17101 | 19


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

188. Che, H. et al. Impaired quality of life in adults with knockout mice. Endocrinology 151, 4607–4612 fibroblast growth factor 23 and phosphate
X‑linked hypophosphatemia and skeletal symptoms. (2010). homeostasis in women. J. Bone Miner. Res. 28,
Eur. J. Endocrinol. 174, 325–333 (2016). 195. Liu, E. S. et al. c‑Raf promotes angiogenesis during 1793–1803 (2013).
This paper highlights the extent and predictors of normal growth plate maturation. Development 143, 203. Martin, A., David, V. & Quarles, L. D. Regulation and
impaired QOL in adults with XLH. The presence of 348–355 (2016). function of the FGF23/Klotho endocrine pathways.
enthesopathy was a key determinant of reduced 196. Gattineni, J. et al. Regulation of renal phosphate Physiol. Rev. 92, 131–155 (2012).
QOL, whereas treatment with phosphate transport by FGF23 is mediated by FGFR1
supplements and/or vitamin D was linked to a and FGFR4. Am. J. Physiol. Renal Physiol. 306, Author contributions
better mental health score. F351–F358 (2013). Introduction (T.O.C. and J.M.P.); Epidemiology (L.M.W. and
189. Ruppe, M. D. et al. Effect of four monthly doses of a 197. Shimada, T. et al. FGF‑23 is a potent regulator of S.A.A.); Mechanisms/pathophysiology (T.O.C., J.M.P., A.A.P.
human monoclonal anti‑FGF23 antibody (KRN23) on vitamin D metabolism and phosphate homeostasis. and S.A.A.); Diagnosis, screening and prevention (N.J.S.);
quality of life in X‑linked hypophosphatemia. J. Bone Miner. Res. 19, 429–435 (2003). Management (T.O.C. and J.M.P.); Quality of life (L.M.W.);
Bone Rep. 5, 158–162 (2016). 198. Andrukhova, O. et al. FGF23 acts directly on renal Outlook (A.A.P.); Overview of Primer (T.O.C. and J.M.P.).
190. Mäkitie, O. et al. Early treatment improves growth proximal tubules to induce phosphaturia through
and biochemical and radiographic outcome in X‑linked activation of the ERK1/2–SGK1 signaling pathway. Competing interests
hypophosphatemic rickets. J. Clin. Endocrinol. Metab. Bone 51, 621–628 (2012). T.O.C. is a consultant for and is participating in clinical trials
88, 3591–3597 (2003). 199. Imel, E. A., DiMeglio, L. A., Hui, S. L., Carpenter, T. O. of burosumab with Ultragenyx Pharmaceutical and has
191. Wöhrle, S. et al. Pharmacological inhibition of & Econs, M. J. Treatment of X‑linked served as a consultant for Alexion. A.A.P. has received hono-
fibroblast growth factor (FGF) receptor signaling hypophosphatemia with calcitriol and phosphate raria from and is participating in a clinical trial of burosumab
ameliorates FGF23‑mediated hypophosphatemic increases circulating fibroblast growth factor 23 with Ultragenyx Pharmaceutical. L.M.W. has received hono-
rickets. J. Bone Miner. Res. 28, 899–911 (2013). concentrations. J. Clin. Endocrinol. Metab. 95, raria from and is participating in a clinical trial of burosumab
192. Yuan, B. et al. Hexa‑D‑arginine treatment increases 1846–1850 (2010). with Ultragenyx Pharmaceutical and has been a consultant
7B2•PC2 activity in hyp-mouse osteoblasts and 200. Clinkenbeard, E. L. et al. Neonatal iron deficiency to Alexion. J.M.P. is a consultant for Biomedical Systems
rescues the HYP phenotype. J. Bone Miner. Res. 28, causes abnormal phosphate metabolism by elevating Corporation. N.J.S. and S.A.A. declare no competing
56–72 (2012). FGF23 in normal and ADHR mice. J. Bone Miner. Res. interests.
193. Goetz, R. et al. Isolated C‑terminal tail of FGF23 29, 361–369 (2014).
alleviates hypophosphatemia by inhibiting 201. Schouten, B. J., Hunt, P. J., Livesey, J. H., How to cite this article
FGF23‑FGFR-Klotho complex formation. Proc. Natl Frampton, C. M. & Soule, S. G. FGF23 elevation Carpenter, T. O. et al. Rickets. Nat. Rev. Dis. Primers. 3,
Acad. Sci. USA 107, 407–412 (2010). and hypophosphatemia after intravenous iron 17101 (2017).
194. Miedlich, S. U., Zhu, E. D., Sabbagh, Y. polymaltose: a prospective study. J. Clin. Endocrinol.
& Demay, M. B. The receptor-dependent Metab. 94, 2332–2337 (2009). Publisher’s note
actions of 1,25‑dihydroxyvitamin D are required 202. Wolf, M., Koch, T. A. & Bregman, D. B. Effects of Springer Nature remains neutral with regard to jurisdictional
for normal growth plate maturation in NPt2a iron deficiency anemia and its treatment on claims in published maps and institutional affiliations.

20 | ARTICLE NUMBER 17101 | VOLUME 4 www.nature.com/nrdp


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.

You might also like