Metabolic and Endocrine Abnormalities


Ovid: Lovell & Winter’s Pediatric Orthopaedics

Editors: Morrissy, Raymond T.; Weinstein, Stuart L.
Title: Lovell & Winter’s Pediatric Orthopaedics, 6th Edition
> Table of Contents > VOLUME 1 > 7 – Metabolic and Endocrine Abnormalities

7
Metabolic and Endocrine Abnormalities
Benjamin A. Alman
Andrew W. Howard

P.168


INTRODUCTION
Biologic Functions of Bone
Although orthopaedists tend to focus on the role of bone
in providing structural support for the body, bone also plays a crucial
role in maintaining mineral homeostasis in serum. The serum levels of
calcium and phosphorus need to be maintained under tight control to
allow for the normal function of a variety of cells. The cancellous
bone has a tremendously large surface area that allows for the rapid
transfer of minerals stored in it, such as calcium, to the serum. This
process occurs at over a million sites in the human skeleton, mediated
by osteoblast and osteoclast cells. A variety of endocrine, metabolic,
and cellular factors are crucial for maintaining this tight homeostatic
balance. Not only do these various factors maintain serum minerals at
their proper levels, but they also act to regulate the amount of bone
present. Because of the interrelationship between these metabolic and
endocrine factors on the one hand, and the distribution of minerals
between the bone and serum on the other, metabolic and endocrine
disorders alter the quantity and quality of bone. This same
interrelationship occasionally works in reverse; that is, disorders
that alter bone structure dysregulate the serum mineral balance (1).
Growing Bone
The effect of metabolic and endocrine disorders on the
skeleton is very different in children and adults; this is because many
endocrine and metabolic factors affect the growth plate. Chondrocytes
in the growth plate go through a coordinated process of
differentiation, beginning with a proliferative phase at the epiphyseal
side of the growth plate and progressing to terminal differentiation
and apoptotic cell death at the metaphyseal side of the physis.
Terminal differentiation is associated with the expression of type X
collagen and the formation of scaffolding for bone formation. Blood
vessels located adjacent to the physis in the metaphyseal bone bring
pleuripotential mesenchymal cells to the region; these cells
differentiate into osteoblasts and produce new bone on the scaffolding
left behind by the growth-plate chondrocytes. This coordinated
differentiation process results in the longitudinal growth of long
bones. The process of growth-plate chondrocyte differentiation needs to
be tightly regulated because, if growth-plate chondrocytes on one side
of the body go through this process at a different rate from those on
the other side, a limb length inequality would result. The process of
growth-plate chondrocyte maturation is regulated by both local and
systemic factors (2). In those conditions in
which these systemic factors are dysregulated, as is the case in
several endocrinopathies, there is an associated growth-plate
abnormality (3). In addition, some endocrine
factors that regulate bone mineral homeostasis, such as thyroid
hormone, also regulate the growth-plate chondrocytes. Therefore,
whereas thyroid hormone dysregulation has implications for bone density
in adults, in growing children thyroid hormone dysregulation can also
cause an abnormality in the growth plate.
FACTORS THAT REGULATE BONE DENSITY
Cells
Bone density is regulated by osteoblast, osteocyte, and
osteoclast cells that add to or break down bone. These cells are
regulated by local and systemic factors, some of which can be modulated
by the mechanical environment. All of these factors are interrelated in
a complex way that has still not been completely elucidated.
Osteoblasts
Osteoblasts are the main cells responsible for laying
down new bone in the form of osteoid. These cells are derived from
pleuripotential stromal precursor cells (sometimes called mesenchymal stem cells)
and are the active cells that lay down new bone during skeletal growth
and remodeling. These cells produce alkaline phosphatase, an enzyme
that is often used to identify osteoblasts and osteoblastic activity.
As the bone matures, osteoblasts become encased in the new bone. Once
the osteoblasts become encased in osteoid, they become relatively
quiescent and are termed osteocytes. In
mature bone, osteocytes are located extremely far away from neighboring
cells and communicate with other cells through long cytoplasmic
processes. The osteocytes remain quiescent until stimulated by hormonal
or mechanical factors to begin to reabsorb or lay down bone. Although
osteoblasts and osteocytes are thought of as cells responsible for
building new bone, they are also able to reabsorb small quantities of
bone; they are able to do this rapidly in comparison to osteoclasts,
which require cellular differentiation and recruitment in order to
reabsorb bone. Therefore, osteoblasts and osteocytes are the first
cells that the body activates when bone reabsorption is required (4).
Osteoclasts
Osteoclasts are derived from circulating monocytes.
After differentiation and recruitment to the site on the bone where
they are required, osteoclasts are able to reabsorb bone in a very
robust manner. Osteoclasts form a ruffled border that attaches to the
osteoid and secretes proteins that degrade the bone matrix. As such,
osteoclasts form active reabsorption cavities called Howship lacunae.
There is an intimate relation between osteocyte and osteoclast
activities, and many of the signals that activate osteoclasts are
mediated by osteocytes. For instance, parathyroid hormone (PTH) does
not directly regulate osteoclast activity,

P.169


but
conveys information via osteocytes, which produce secondary factors
that regulate the differentiation of monocytes to osteoclasts. The
major signaling pathway that is used by osteocytes to regulate
osteoclasts involves a member of the tumor necrosis factor superfamily
called RANKL,
its receptor, RANK, and a circulating inhibitor, osteoprotegerin (OPG).
The receptor, RANK, is present on osteoclast precursor cells; when
stimulated, it causes these precursors to differentiate into active
osteoclasts. RANKL is produced by osteocytes that are stimulated to
reabsorb bone. OPG also binds to RANKL, inhibiting its ability to
activate RANK, and thereby inactivating osteoclasts. The balance
between OPG and RANKL regulates the number of osteoclasts available (Fig. 7.1) (4,5).
Because OPG inhibits osteoclast production, its use represents a
promising approach to inhibiting osteoclastic activity and,
consequently, it has the potential to be developed into a useful
therapy for osteoporosis, neoplastic bone loss, and even the loosening
surrounding total joint implants (6,7).

Genetic Mechanisms Controlling Bone Density
In recent years, tremendous advances have been made in
understanding genes that regulate how bone cells develop. Much of this
information is covered in several review articles (1,2)
and is beyond the scope of this book. For the purpose of this chapter,
we consider three modulators of bone density: (a) calcium homeostasis,
(b) hormone factors, and (c) physical forces.
Hierarchy in the Regulation of Bone Mass
There is a hierarchy among the various factors that
regulate bone mass. Calcium mobilization overrides the other functions
of the skeleton. Calcium deficiency resulting from renal disease,
malabsorption, or a diet poor in calcium invariably causes bone loss,
which cannot be overcome by modulating any of the other factors that
regulate bone mass. The effects of hormones such as estrogen seem to be
more potent than the effect of physical forces; this is suggested by
the facts that exercise has only limited ability to maintain or restore
bone mass in postmenopausal women, and amenorrhoeic marathon runners
lose bone. Of the three modulators of bone mass—calcium availability,
hormonal factors, and physical forces—the last has the least pronounced
effects, although this is the one that orthopaedic surgery concentrates
most of its efforts on (4).
Figure 7.1
Expression of RANKL by osteoblasts and osteocytes activates RANK
receptor on preosteoclasts to cause differentiation into active
osteoclasts. Osteoprotegerin (OPG) is a circulating factor that can
also bind to RANK, but inhibits its ability to cause differentiation to
osteoclasts.
Calcium Homeostasis
Biologic Functions of Calcium
Calcium plays a crucial role in the irritability,
conductivity, and contractility of smooth and skeletal muscle, and in
the irritability and conductivity of nerves. Small changes in
extracellular and intracellular calcium levels lead to dysfunction of
these cells. In the case of neurons, cellular activity is inversely
proportional to calcium ion concentration, whereas for cardiac myocytes
there is a direct proportionality. Therefore, decreases in ionic
calcium concentration can lead to tetany, convulsions, or diastolic
death. Conversely, increases in the concentration of calcium can lead
to muscle weakness, somnolence, and ventricular fibrillation. It is
obviously important for the body to guard the concentration of ionized
calcium, thereby providing a rationale for the overriding importance of
calcium homeostasis in modulating bone density (8,9).
Normal Calcium Balance
Calcium is absorbed from the gut, stored in the bone,
and excreted primarily by the kidney. Therefore, diseases that affect
gut absorption or renal function have the potential to dysregulate
normal calcium homeostasis and bone mass. In addition, some conditions
that cause massive loss of bone mass, such as widespread metastatic
disease or prolonged bed rest, can also alter serum calcium levels.
Almost all of the body’s calcium is stored in the bones and is held in
the form of hydroxyapatite, a salt that is composed of calcium,
phosphorus, hydrogen, and oxygen (CaHPO) in very tiny crystals embedded
in the collagen fibers of the cortical and cancellous bone (10,11,12,13).
The small size of the crystals provides an enormous surface area, and
this factor, combined with the reactivity of the crystal surface and
the hydration shell that surrounds it, allows a rapid exchange process
with the extracellular fluid (ECF). This process converts the
mechanically solid structure of bone into a highly interactive
reservoir for calcium, phosphorus, and a number of other ions (12,14).
Serum Calcium and Phosphate
Hydroxyapatite is not freely soluble in water. At the pH
of body fluids, calcium and phosphate concentrations in serum exceed
the critical solubility product, and are predicted to precipitate into
a solid form. It is thought that various plasma proteins act to inhibit
the precipitation and

P.170


keep
these ions in solution. This metastable state is important for bone
structure, as it allows the deposition of hydroxyapatite with a minimal
expenditure of energy during bone formation. Unfortunately, it also
makes the occurrence of ectopic calcification and ossification easy,
because of increments in the levels of either or both of these ions.

Active Transport of Calcium—Regulators
Calcium cannot passively diffuse across mammalian cell
membranes, and requires an active transport machinery to move it into
or out of cells (8,12,14,15,16).
Although the mechanism to control this transport is regulated, in large
part, by the action of the active form of vitamin D, PTH, and the
concentration of phosphate (14,17,18),
a variety of other cell-signaling pathways also play their roles in
transporting calcium across cell membranes. These other cell-signaling
pathways, however, seem to act in specialized cell types under specific
physiologic states and, as such, are likely to play only a small role
in regulating the total serum calcium level. Therefore, PTH, vitamin D,
and phosphate are the three factors that play the most crucial roles in
the calcium transport process and, consequently, in maintaining the
normal extracellular soluble calcium level.
Parathyroid Hormone
PTH is produced by cells of the parathyroid glands, and
the expression level of PTH is regulated by the serum level of ionized
calcium. When serum calcium levels are low, there is an increase in PTH
expression and protein production, leading to increased PTH levels in
the serum. There are four parathyroid glands, and any one gland has the
potential to produce enough PTH to maintain calcium homeostasis. This
fact is of importance in the surgical management of thyroid neoplasia,
in which it is preferable to maintain the viability of at least one
parathyroid gland. PTH binds to a family of cell membrane receptors
called parathyroid hormone receptors (PTHR) that activate a number of
cell-signaling pathways. The pathway that has been studied the most in
the control of calcium is the one that regulates adenyl cyclase
activity, resulting in an increased cellular level of cyclic adenosine
monophosphate (cAMP). cAMP renders the cell membrane more permeable to
ionic calcium, and it induces the mitochondria, which are intracellular
storehouses for calcium, to release their calcium. These actions
increase the intracellular concentration of calcium, but do not promote
transport to the extracellular space, a function that also requires
vitamin D. PTH acts with 1,25-dihydroxyvitamin D to facilitate cellular
calcium transport in the gut, the renal tubule, and in the lysis of
hydroxyapatite crystals (16,17,19).
PTH directly stimulates osteoblasts to begin to degrade the surrounding
calcium-rich osteoid. Osteoclasts do not contain receptors for PTH, but
are stimulated by PTH activation in osteoblasts through induction of
the expression of RANKL, which activates osteoclasts (19,20).
Another action of PTH is to diminish the tubular reabsorption of
phosphate, thereby causing the renal excretion of phosphate (19,21,22).
Vitamin D
Active vitamin D is produced from provitamins through conversion steps in the skin, liver, and kidney (Fig. 7.2). The provitamins are ingested from animal fats (ergosterol) or synthesized by the liver (7-dehydrocholesterol) (10,16,23),
and are converted to calciferol and cholecalciferol by ultraviolet
light, a process that occurs in the skin. In the absence of ultraviolet
light, this conversion cannot occur; this explains why vitamin D
deficiency is associated with prolonged periods indoors away from
ultraviolet light sources, such as in chronically ill individuals, or
in individuals living in extremely cold climates (10,24). The compounds are then transported to the liver, where they are converted to 25-hydroxyvitamin D by a specific hydrolase (25,26,27,28,29).
Severe liver disease or the intake of drugs that block hydrolase
activity will inhibit the production of 25-hydroxyvitamin D, also
potentially leading to vitamin D deficiency. The final conversion
occurs in the kidney. In the presence of specific hydrolases and a
number of biochemical cofactors, 25-hydroxyvitamin D is converted to
either 24,25-dihydroxyvitamin D or 1,25-dihydroxyvitamin D. The latter
serves as the potent calcium transport promoter (30–32). A low serum
calcium level and a high PTH level cause conversion to the 1,25 analog,
whereas a high serum calcium level, a higher serum phosphate level, and
a low PTH level favor formation of 24,25-dihydroxyvitamin D, which is
less potent in activating calcium transport (Fig. 7.3) (30,33,34,35,36).
Serum phosphate also plays an important role here, because a high
concentration of phosphate shunts the 25-hydroxyvitamin D into the
24,25-dihydroxy form. Although the 24,25-dihydroxy form is less active
in regulating calcium levels, it has an important role in growth-plate
chondrocytes. This crucial role of the kidney in converting vitamin D,
as well as its important role in excreting excess calcium and
phosphorus, explains the particularly deleterious effect of renal
failure on bone homeostasis, causing vitamin D deficiency as well as
directly dysregulating normal calcium excretion. Because of the crucial
role of vitamin D in calcium metabolism, the National Academy of
Sciences and the American Academy of Pediatrics recommends 200 IU per
day of vitamin D (37). This dose will prevent
physical signs of vitamin D deficiency and maintain serum
25-hydroxyvitamin D at or above 27.5 nmol per L (11 ng per mL). The
generic name of 1,25-dihydroxyvitamin D is calcitriol.
Recent studies found that vitamin D also plays a variety of
extra-skeletal roles, including in modulation of the immune response,
and as a chemoprotective agent against certain cancers (38).
Dietary Calcium Intake
Dietary calcium is crucial for the maintenance of bone mass. Daily requirements vary depending upon the need

P.171



for calcium during periods of rapid bone growth. Recommendations from the American Academy of Pediatrics are summarized in Table 7.1 (8,24,39,40).
Adequate calcium in the diet during adolescent years is important for
the maintenance of bone mass over the long term, and orthopaedists
should counsel their patients about the importance of including
appropriate amounts of calcium, as well as vitamin D, in their diet.
Several dietary factors alter calcium absorption. Calcium salts are
more soluble in acid media; therefore the loss of the normal
contribution of acid from the stomach reduces the solubility of the
calcium salts and decreases the absorption of the ionized cation. A
diet rich in phosphate may decrease the absorption of calcium by
binding the cation to HPO2-4 and precipitating most of the ingested calcium as insoluble material (12,41).
Ionic calcium can be chelated by some organic materials with a high
affinity for the element, such as phytate, oxalate, and citrate.
Although these materials may remain soluble, they cannot be absorbed (12,41,42,43). Calcium, in the presence of a free fatty acid, forms an insoluble soap that cannot be absorbed (41,44).
Disorders of the biliary or enteric tracts, associated with
steatorrhea, are likely to reduce the absorption of calcium because it
forms an insoluble compound and because the fat-soluble vitamin D that
is ingested is less likely to be absorbed under these circumstances (45).

Figure 7.2 The conversion of vitamin D from the skin or from dietary sources takes place in the liver and kidney. A: In the liver, the enzyme vitamin D 25-hydrolase acts to form 25-hydroxy vitamin D. B: The second conversion of vitamin D takes place in the kidney, where at least two pathways have been described. The maintenance pathway [when the need is minimal, as defined by a normal calcium and phosphorus and low parathyroid hormone (PTH)
level] occurs in the presence of a specific enzyme (25-hydroxyvitamin D
24-hydrolase) and results in the less active 24,25-dihydroxyvitamin D.
If calcium transport is required, as signaled by the presence of low
serum calcium and phosphorus levels and a high PTH level, the body
converts the 25-hydroxyvitamin D to the much more active form,
1,25-dihydroxyvitamin D.
TABLE 7.1 DIETARY CALCIUM REQUIREMENTS
Age Calcium Requirement (mg/d)
0 to 6 mo 210
6 mo to 1 y 270
1 through 3 y 500
4 through 8 y 800
9 through 18 y 1300
Dietary Phosphate Intake
Phosphate (PO4) is absorbed lower in the
gastrointestinal (GI) tract than is calcium, and is freely transported
across the gut cell to enter the extracellular space, in which it
represents a major buffer system. Transport into and out of the bone is
passive and is related to the kinetics of the formation and breakdown
of hydroxyapatite crystals. Tubular reabsorption of phosphate, however,
is highly variable, with reabsorption ranging from almost 100% to less
than 50%. The principal factor that decreases tubular reabsorption of
phosphate is PTH.
Figure 7.3
The roles of the bone, kidneys, gastrointestinal tract, parathyroid
gland, and thyroid gland in calcium kinetics. These organs act to
maintain calcium in the extracellular fluid (ECF) at the appropriate levels for normal cellular function. Vitamin D and parathyroid hormone (PTH)
act to transport calcium ions across the gut wall and regulate renal
excretion and, thereby, bone calcium content. Depending on the need for
increased transport, 25-hydroxy vitamin D is converted to 24,25- or
1,25-dihydroxyvitamin D. A: In the normocalcemic state, an equilibrium between calcium intake and excretion is maintained by the various organs. B:
In the hypocalcemic state, a reduced concentration of calcium signals
the parathyroid glands to release more PTH, which acts at the levels of
the gut cell, renal tubule, and bone to increase transport of calcium
and rapidly replenish body fluids with it. An increase in PTH also
favors the synthesis of 1,25-dihydroxyvitamin D in the kidney and acts
to promote renal phosphate excretion by markedly diminishing the
tubular reabsorption of phosphate. C: In
the hypercalcemic state, low concentrations of calcium and PTH act
independently to diminish the synthesis of 1,25-dihydroxyvitamin D and
decrease transport of calcium in the gut cell, tubule, and bone.
Increased concentrations of calcium also cause the release of
calcitonin (CT) from the C-cells of the
thyroid gland, thereby diminishing calcium concentration. This
mechanism principally involves stabilizing the osteoclast and
decreasing its action on the bone, but it is not very effective in
humans.

P.172


Endocrine Factors
Sex Steroids
The most potent endocrine regulator of bone density is
estrogen. Most of the clinical and experimental data on the role of
estrogen in bone have been generated from studies of postmenopausal
women. However, clinical data from children with deficiencies in sex
hormones, such as in Turner syndrome, also show that a lack of estrogen
in growing girls is responsible for profound loss of bone density. The
exact mechanism by which estrogen regulates bone formation and loss is
unknown. Estrogen receptors are present on both osteoblasts and
osteoclasts, yet the cellular mechanism by which estrogen alters the
behavior of

P.173


these
cells is not clear. Studies in animals suggest that estrogen exhibits
at least some of its effects through the regulation of pleuripotential
stromal cells in the bone marrow, a process that may be mediated by
interleukin-6. Estrogen also suppresses the activation of osteoclasts
by inhibiting the activation of RANK in the precursor cells (46,47).

Androgens also seem to regulate bone mass, although the
mechanism is less well understood than it is in the case of estrogen.
Idiopathic hypogonadotropic hypogonadism is associated with decreased
bone mass, and there is an association between delayed puberty and low
bone mass in boys, suggesting a positive role for androgens in
regulating bone mass (48).
Thyroid Hormones
Thyroid hormones act in the cell nucleus, interacting
with nuclear proteins and DNA to increase the expression of a variety
of genes, ultimately regulating cell activity positively. Thyroid
hormone activates both osteoblasts and osteoclasts. The actual effect
on bone mass depends on the body’s balance between these two cell
types, and how well the normal control of the calcium level is able to
counteract their heightened activity. In general, the balance is in
favor of the osteoclast and, very often, increased thyroid hormone
levels lead to bone loss (49,50).
Corticosteroids
Corticosteroids have a variety of effects on cells; they
inhibit cellular activity in general, potentially decreasing the
ability of osteoblasts to lay down new bone. Corticosteroids also have
profound effects on the skeleton because of their effect on calcium
regulation in the kidney, where they increase calcium excretion; this
leads, secondarily, to elevated PTH levels, with consequent negative
effects on bone density (51,52).
Calcitonin
CT is produced by parafollicular thyroid cells. Although
CT causes inhibition of bone resorption by osteoclasts and osteoblasts,
and decreases reabsorption of calcium and phosphate in the kidneys in
animal models and cell cultures, it seems to play a minor role in
humans (53,54).
Mechanical Factors
Excessive reductions in bone strain produced by
weightlessness (microgravity in outer space) or immobilization
(paralysis, prolonged bed rest, or application of casts) can cause
significant bone loss, whereas strenuous athletic activity can augment
certain bones (54,55).
This effect is important in pediatric orthopaedic patients, in whom
many of the neuromuscular disorders are associated with decreased
weight-bearing and associated osteoporosis. Bone remodels itself
according to the mechanical stresses applied, a phenomenon termed Wolff’s law.
It is well known that the mechanical environment alters cell behavior
and gene expression, and it is thought that such a mechanism, most
likely acting through osteocytes, is responsible for the effect of
weight-bearing on bone density, as well as for the changes attributable
to Wolff’s law (56,57).
FACTORS THAT REGULATE GROWTH-PLATE CHONDROCYTES
In recent years, a number of signaling pathways that regulate the function of growth-plate chondrocytes have been elucidated (2).
General information about growth-plate development and its local
regulation is covered in the chapter on developmental biology. However,
it is apparent that growth-plate chondrocytes are affected, either
primarily or secondarily, by a variety of endocrine regulatory factors,
and so a brief review is warranted here.
Growth-plate chondrocytes at the epiphyseal side of the
growth plate reside in the resting zone; they begin to proliferate and
advance toward the metaphyseal side of the growth plate in the
proliferative zone; following this, they enter a prehypertrophic zone,
where they shift from proliferation to differentiation. It is also in
this prehypertrophic zone that important signals that regulate the
differentiation process, such as PTH-related protein and Indian
hedgehog, are present. Following this exposure, the cells undergo
hypertrophy, form columns in the hypertrophic zone, and then undergo
terminal differentiation and cell death. Blood vessels from the
metaphysis, adjacent to the terminally differentiated chondrocytes,
bring in new pleuripotential mesenchymal cells that differentiate into
osteoblasts, forming new bone on the scaffolding left behind by the
chondrocytes. This last region is sometimes called the zone of provisional calcification.
It is easy to imagine how hormones can tip the balance
in favor of or against the differentiation process in these cells. In
addition, agents that alter normal bone formation by osteoblasts can
also alter the growth plate by preventing the normal replacement of the
terminally differentiating chondrocytes with new bone. This inhibition
of normal ossification results in the characteristic growth plate
changes in rickets, in which there is an increased zone of terminal
differentiation. Endocrinopathies can also alter size and matrix
components in the various zones of the growth plate. Such disorders
that affect terminal differentiation may make the growth plate
mechanically weaker in this region, predisposing the patient to
conditions such as slipped capital femoral epiphysis (SCFE). In a
similar manner, it may make the growth-plate chondrocytes easier to
deform with compressive pressure, causing deformities such as genu
varum; this explains the high frequency of these growth-plate
deformities in children with endocrine disorders. As in bone,
mechanical factors can also play a role in growth plate function. The
Hueter-Volkmann Principle states that growth plates exhibit increased
growth in response to tension and decreased growth in response to
compression (58). Therefore, an endocrinopathy can cause growth-plate deformities, which can then be

P.174



exacerbated by the effect of the changing mechanical axis in the affected limb.

Similar to the situation in the bone, there is also a
hierarchal regulation of the growth plate, with endocrine factors
playing a dominant role over mechanical factors (3,59).
This fact is readily apparent in conditions such as rickets, where
surgery will not result in correction of the genu varum in the absence
of the correction of the underlying endocrinopathy in the growing
child. Therefore, it is important to avoid the temptation to undertake
surgical correction of a deformity in a growing child with an endocrine
disorder until the endocrinopathy is treated.
There are a large number of endocrine factors that play
a role in regulating growth-plate function. In many cases, not much is
known about the intracellular signaling mechanisms utilized by these
factors. Growth hormone plays an important role in regulating the
proliferation of growth-plate chondrocytes, and the process is mediated
by somatomedins. In the absence of growth hormone, there is a slowing
of growth-plate maturation, as well as a slowing of the rate of long
bone growth. Thyroid hormone also plays a role in regulating
chondrocyte activity, by increasing the metabolic and proliferative
rates of growth-plate chondrocytes. PTH may alter chondrocyte
maturation in growth plates because the PTH receptor PTHR1 is expressed
in prehypertrophic chondrocytes and its stimulation results in an
inhibition of terminal differentiation. Nutrition and insulin also
regulate growth-plate chondrocytes, in a manner similar to that of
growth hormone, by regulating growth-plate chondrocyte proliferation. A
lack of dietary protein exerts a negative control over the
somatomedins. An excess of glucocorticoids also inhibits growth, partly
by an inhibitory effect on protein synthesis in cartilage, but also by
interference with somatomedin production and action (3).
Although all these factors play roles in regulating growth-plate
chondrocytes, we will likely learn more in the coming years about the
roles they play in a variety of growth-plate pathologies, including
disorders such as SCFE, where it is well known that a variety of
endocrinopathies are predisposing conditions.
DISEASES OF BONE
Rickets
Context/Common Features
Rickets is described as a
clinical condition in which there is inadequate mineralization of
growing bone. Severe nutritional rickets was endemic in early
industrialized societies, particularly where sunlight was scarce.
Accordingly, severe rachitic deformities were commonly seen in the
early days of orthopaedics (60,61). In developed countries nutritional rickets is now a rarity, although it may present de novo to pediatric orthopaedists for diagnosis. Inherited forms of rickets are still commonly seen in the United States (62).
The surgeon should also be familiar with renal tubular abnormalities
that can result in rickets, as well as with the clinical entity of
renal osteodystrophy, which describes the bone disease associated with
end-stage renal disease and includes features of rickets as well as
secondary hyperparathyroidism.
The clinical manifestations of all forms of rickets are
similar, and therefore the clinical presentation will be covered first,
prior to a discussion of the various etiologies.
Clinical Presentation
Rickets is failure of or delay in calcification of newly
formed bone at long bone physes. The manifestations include changes in
growth-plate morphology, with decreased longitudinal growth and angular
deformities of the long bones. Osteomalacia, which is the failure of
mineralization of osteoid formed at cortical and trabecular surfaces,
often accompanies rickets in childhood. In the adult, osteomalacia is
the only result of the mechanisms that cause rickets in childhood.
The skeletal abnormalities of severe rickets present in
early childhood and often before the age of 2 years. The child may have
a history consistent with hypocalcemia in infancy, including apneic
spells, convulsions, tetany, and stridor prior to the age of 6 months (63).
The child is often hypotonic with delayed motor activity milestones for
sitting, crawling, and walking. There is proximal muscle weakness and
sometimes profuse sweating. Cardiomyopathy and respiratory and GI
infections may accompany the clinical presentation (64,65,66,67,68,69).
Skeletal deformities may be evident at every physis. The
wrists, elbows, and knees are thickened, and the long bones are short.
Genu varum or valgum may be present, as may coxa vara. Costochondral
enlargement leads to the characteristic rachitic rosary appearance of
the chest. Harrison’s sulcus is an indentation of the lower ribs caused
by indrawing against the soft bone. Kyphoscoliosis may be present.
Closure of the anterior fontanelle is delayed. Frontal and parietal
bossing of the skull is evident. Plagiocephaly may be related to the
effect of positioning on a soft skull. Delayed primary dentition is
common (70,71).
Radiographic Changes
The radiographic hallmark of rickets is widened and indistinct growth plates (Fig. 7.4). In a normal child, the distance between the metaphysis and epiphysis of the distal radius should never be greater than 1 mm (72).
Lateral expansion of the growth plates also occurs,
particularly with weight bearing. One should remember that crawling
children bear weight on their wrists, and this explains the thickened
wrists as well as knees. The metaphysis typically takes on a cupped and
splayed appearance. The long bones are short for age, and show evidence
of coxa vara and genu varum or valgum as described in the preceding
text. Radiography may reveal further evidence of osteomalacia. The
hallmark is Looser zones. These are

P.175


transverse
bands of unmineralized osteoid, which typically appear in the medial
aspect of the proximal femur and at the posterior aspect of the ribs.
These are described as pseudofractures and often have an osteosclerotic
reaction area around them. In an adult, they can progress to true
fractures. Acetabular protrusion and pathologic fractures complete the
radiographic signs of rickets (64,70,72,73).

Figure 7.4 Rickets. Changes caused by rickets can be seen (A) at the wrist and (B)
at the knees of this 1-year-old child with X-linked hypophosphatemia.
The growth plates are widened and the metaphyses are cupped,
particularly at the ulna and femur. At 4 years of age (C, D) the changes have resolved with medical treatment.
Overview of Classification of Rickets
Bone is mineralized by the crystallization of calcium
and phosphate in the presence of alkaline phosphatase enzyme. Calcium
and phosphate are maintained very close to their solubility coefficient
by a complex series of inhibitors. The control mechanisms in the
physiologic state have been discussed in earlier sections of this
chapter.
A useful way of approaching the management of rickets is
to consider those conditions that reduce the availability of calcium,
those conditions that reduce the availability of phosphate, and the
rare condition that reduces the availability of alkaline phosphatase at
the osteoblast-bone junction (Table 7.2).
Nutritional rickets and end organ insensitivity to calcitriol are
problems as far as calcium is concerned. X-linked hypophosphatemia is
the most common form of rickets seen today in the United States. It is
caused by renal tubular phosphate wasting in isolation (74).
Renal tubular abnormalities, including Fanconi syndrome, cause renal
wasting of phosphate, calcium, magnesium, and bicarbonate. Alkaline
phosphatase is deficient only in one rare recessive condition,
appropriately called hypophosphatasia.
TABLE 7.2 CLASSIFICATION OF RICKETS ACCORDING TO WHAT IS LACKING AT THE OSTEOBLAST-BONE INTERFACE
Calcium
   Nutritional rickets
      Vitamin D deficiency (common)
      Isolated calcium deficiency (rare)
      Combined calcium deficiency and marginal vitamin D intake (common)
   Gastrointestinal rickets
   1 α-hydroxylase deficiency
   End organ insensitivity to vitamin D
   Rickets of end-stage renal disease (renal osteodystrophy)

Phosphorus
   X-linked hypophosphatemia (common)
   Renal tubular abnormalities

Alkaline Phosphatase
   Hypophosphatasia

P.176


Finally, renal osteodystrophy is often discussed with
rickets, and appropriately so because many children with renal
osteodystrophy manifest signs of rickets. However, renal osteodystrophy
classically includes the changes of secondary hyperparathyroidism as
well as those of rickets.
Nutritional Rickets
Nutritional rickets had near universal prevalence in
northern industrialized societies in the 19th century. It has now
largely disappeared in developed countries. It remains a significant
clinical problem in the developing world; for example, a 66% prevalence
of clinical rickets was found in preschool children in Tibet in 2001 (75).
The main cause of nutritional rickets is vitamin D deficiency. Vitamin D3
(cholecalciferol) can be produced in the skin by a process that
requires ultraviolet B radiation, or it can be ingested in the diet.
The peak age of presentation of nutritional rickets is between 3 and 18
months in children who have inadequate exposure to sunlight and no
vitamin D supplementation in the diet (76,77,78). Breast milk is poor in vitamin D, and prolonged breast-feeding is a risk factor (79,80). Vitamin D is supplemented in dairy foods in North America, and diets deficient in dairy foods are therefore a risk factor (80,81,82). Two hundred international units of vitamin D is the recommended daily amount for preventing rickets (37,62).
Exposure to sunlight also prevents rickets. Two hours per week of
summer sunshine at the latitude of Cincinnati (39 degrees north) is
sufficient to produce adequate vitamin D in the skin. However, during
the winter months in Edmonton (52 degrees north), there is insufficient
UVB exposure to allow for adequate intrinsic production of vitamin D (62).
Although vitamin D deficiency is the principal cause of
nutritional rickets, it is possible to get rickets because of a
profoundly calcium-deficient diet, even in the presence of adequate
vitamin D intake (83). Probably much more
common is a subtle combination of calcium deficiency and vitamin D
deficiency interacting to produce dietary rickets (84). This interaction has been described among the modern Asian population in the United Kingdom (85,86) and black populations in the United States (71).
A diet that is low in calcium and high in phytate, oxalate, or citrate
(substances that bind calcium and are found in almost all fresh and
cooked vegetables) means that calcium availability in the diet is poor.
Vegetarians, especially those who avoid dairy products, are
particularly at risk. This lack of calcium produces an increase in PTH,
which in turn increases vitamin D catabolism. Vitamin D status may
already have been marginal because of low exposure to sunlight and poor
dietary intake. The increased catabolism of vitamin D, along with these
factors, results in vitamin D deficiency and the clinical presentation
of rickets. This combination of deficiencies of both calcium and
vitamin D is very prevalent among adolescents who present with rickets
in the United Kingdom and the United States.
Treatment of nutritional rickets involves adequate
provision of vitamin D. A dose of 5000 to 10,000 International Units
(IU) per day for 4 to 8 weeks should be prescribed, along with calcium
to the extent of 500 to 1000 mg per day in the diet (87).
Where daily dosing and compliance pose problems, much larger doses of
vitamin D (200,000 IU to 600,000 IU orally or intramuscularly) have
been given as single doses with good results (88).
Laboratory abnormalities in established nutritional
rickets may include a low-normal or decreased calcium ion
concentration, a low serum phosphate level, a low serum
25-hydroxyvitamin D3 level, and a high alkaline phosphate level. Alkaline phosphate level drops to normal in response to successful therapy.
Gastrointestinal Rickets
Even if adequate calcium and vitamin D are present in the diet, some GI diseases prevent their appropriate absorption (89).
Gluten-sensitive enteropathy, Crohn disease, ulcerative colitis,
sarcoidosis, and short-gut syndromes have been implicated. If liver
disease interferes with the production of bile salts, then fat
accumulates in the GI tract and prevents the absorption of fat-soluble
vitamins including vitamin D. The resulting deficiencies of vitamin D
and calcium cause bone disease in the same way as nutritional
deficiencies do, but the treatments need to be aimed at rectifying the
underlying GI problem as well as at supplementing the deficient vitamin
and mineral.
X-linked Hypophosphatemia
X-linked hypophosphatemia is the most common inherited etiology for rickets with a prevalence of 1 in 20,000 persons (90).
It is an X-linked dominant disorder; this means the female-to-male
patient ratio is approximately 2:1, and there is no male-to-male
transmission. Approximately one-third

P.177



of the cases are sporadic (90).
Individuals in whom the condition has sporadically occurred do transmit
the defect to their offspring. The defect is in a gene called PHEX (91).
This gene product indirectly regulates the transport of renal
phosphates. The defect at the kidney is isolated renal phosphate
wasting, leading to hypophosphatemia. In addition, a low or normal
kidney production of 1,25-dihydroxyvitamin D3 is observed, which is inappropriate in the hypophosphatemic state.

The clinical presentation includes rickets and mild shortening in stature (92,93). Dental abscesses occur in childhood, even prior to the development of dental caries (94).
Adults with the condition have osteomalacia accompanied by degenerative
joint disease, enthesopathies, dental abscesses, and short stature (95,96,97). The specific treatment for the condition is oral administration of phosphate and the active form of vitamin D3,
calcitriol, which is 1 α-hydroxylated. Treatment requires careful
metabolic monitoring. Hyperparathyroidism, soft tissue calcification,
and death caused by vitamin D intoxication have been problems with
medical therapy in the past. Calcitriol can be used in much lower doses
than the less-active vitamin D metabolites previously used, and is
thought to be a safer therapy (91,98,99). Angular deformities, particularly genu valgum, may persist after medical treatment, requiring osteotomy (100,101,102).
A small number of patients with McCune-Albright syndrome
also develop hypophosphatemic rickets. This syndrome includes patients
with café au lait spots, precocious puberty, and fibrous dysplasia of
multiple long bones. The syndrome is caused by constitutional
activation of the cyclic adenosine monophosphate-protein kinase A
(AMP-PKA) signaling pathway, and is related to genetic defects in G
signaling proteins (103).
1 α-Hydroxylase Deficiency
In 1961, Prader et al. described what was initially called vitamin D-dependent rickets (104);
this was because the early patients were treated with very large doses
of vitamin D. It turns out that such patients have 1 α-hydroxylase
deficiency and they can be treated with much smaller quantities of the
biologically active 1 α-hydroxylated calcitriol (105).
Typically, the patients are less than 24 weeks of age with weakness,
pneumonia, seizures, bone pain, and the skeletal bone changes of
rickets. Serum findings include low calcium and phosphorus, high
alkaline phosphatase and PTH, with a normal level of 25-hydroxyvitamin D3 but a markedly decreased level of 1,25-dihydroxyvitamin D3. The patients are not able to convert the accumulated 25-hydroxyvitamin D3 to its biologically active form of 1,25-dihydroxyvitamin D3 and therefore develop clinical rickets. The autosomal recessive genetic pattern has been described (106) and the specific mutations were initially described in 1997 (107), since which time at least 31 distinct mutations in the 1 α-hydroxylase gene have been identified (105).
Currently, the treatment consists of oral administration of activated vitamin D3 which is curative.
End Organ Insensitivity
In 1978, Marx et al. (108)
described two sisters with clinical rickets. The unusual clinical
feature was an exceedingly high circulating level of
1,25-dihydroxyvitamin D3. In patients having this condition, these levels can be 3- to 30-fold higher than normal (109).
A striking clinical finding is alopecia or near total loss of hair from
the head and body. These patients have an end organ insensitivity to
vitamin 3 (109). Treatment with
very high doses of vitamin D produces a variable but incomplete
clinical response. Intravenous high doses of calcium followed up with
oral calcium supplementation in large quantities has also been tried,
but as yet this rare form of rickets cannot be completely cured (105).
Renal Tubular Abnormalities
There are many causes of the Fanconi syndrome. This
syndrome implies failure of tubular reabsorption of the many molecules
smaller than 50 Da. The kidneys lose phosphate, calcium, magnesium,
bicarbonate, sodium, potassium, glucose, uric acid, and small amino
acids. With this renal tubular abnormality, there are multiple
mechanisms by which bone mineral homeostasis is disrupted (89,110).
As a result, such patients are short, with rickets and delayed bone
age. The predominant cause of bone disease is hypophosphatemia from
renal phosphate wasting, very similar to that seen in X-linked
hypophosphatemic rickets. Other mechanisms include calcium and
magnesium loss, metabolic acidosis caused by bicarbonate loss, renal
osteodystrophy (if renal disease is sufficiently advanced that less
1,25-dihydroxyvitamin D3 is produced), and finally, decreased calcium and phosphate reabsorption.
The treatment is similar to that of X-linked
hypophosphatemia: administration of oral phosphate and vitamin D.
Electrolyte imbalances from other causes need monitoring and treatment,
and the underlying renal disease should also be treated if possible.
Hypophosphatasia
Hypophosphatasia is another disease having clinical
overlap with rickets. Hypophosphatasia is caused by alkaline
phosphatase deficiency. Like most enzyme deficiencies, this is a
recessive condition with over 112 mutations described in the alkaline
phosphatase gene in chromosome 1 (111,112,113).
Clinically, alkaline phosphatase deficiency produces abnormal
mineralization of bone with a presentation of rickets in the child or
osteomalacia in the adult (114). Pathologic fractures can occur in children and in adults (115,116).
The fractures are accompanied by abnormal formation of dental cementum,
which causes loss of teeth. The primary teeth are lost early and with
minimal root resorption (117). Additional clinical manifestations may include failure to thrive, increased intracranial pressure, and craniosynostosis.
Figure 7.5
Radiograph of the pelvis of a patient with renal osteodystrophy shows
the marked changes of secondary hyperparathyroidism. Several brown
tumors are seen in the femoral shafts and ischial rami. These appear as
expanded destructive lesions, resembling primary or metastatic bone
tumors.

P.178


Hypophosphatasia has an estimated prevalence of 1 per 100,000 people (118).
There is a perinatal lethal form of the condition. A childhood form
presents with rickets at 2 or 3 years of age and remission of the
disease in adolescence. An adult form presents with mild osteomalacia
with pathologic fractures (111).
There is no satisfactory medical treatment for the
underlying defect. Bone marrow transplantation has been used
experimentally in severe cases, with the aim of repopulating the bone
marrow with osteoblasts capable of producing alkaline phosphatase (111).
Surgical treatment of femoral fractures and pseudofractures in the
adult has been reported, with rodding techniques found to be superior
to plating techniques in the abnormal bone (115).
Renal Osteodystrophy
Renal osteodystrophy describes the bony changes
accompanying end-stage renal disease; it is commonly seen in patients
on dialysis. The clinical presentation includes hyperparathyroidism as
well as rickets with osteomalacia in varying combinations (119,120,121,122).
Renal failure leads to inadequate clearance of phosphate
from the blood once the renal function drops below 25% to 30% of
normal. The hyperphosphatemia drives the solubility equilibrium to
produce hypocalcemia, which signals the parathyroid glands to produce
PTH, causing secondary hyperparathyroidism. The bony changes of
hyperparathyroidism then become evident. These include subperiosteal
erosions and brown tumors (Fig. 7.5). The
subperiosteal erosions are described as classically appearing on the
radial margins of the middle phalanges of digits 2 and 3 of the hand in
adults. In children, they can also be seen at the lateral aspects of
the distal radius and ulna and at the medial aspect of the proximal
tibia (Fig. 7.6) (123).
Prolonged stimulation of the parathyroid glands can produce sufficient
hyperplasia to make the glands remain autonomous and maintain a
hyperparathyroid state, even if the end-stage renal disease is treated
by transplantation. In this case, the ongoing hyperparathyroidism is
described as tertiary rather than secondary.
The other aspect of renal osteodystrophy is rickets. If
there is inadequate renal mass to produce sufficient
1,25-dihydroxyvitamin D3 rickets (clinical and radiographic)
will accompany renal osteodystrophy. The clinical manifestations can
include varus or valgus deformities at the knees or ankles, and
radiographic evidence of widened and deformed growth plates and
rickets/osteomalacia such as Looser zones (Fig. 7.7).
Treatment of renal osteodystrophy includes:
  • Dietary phosphate restriction
  • Phosphate binding agents, especially those which are calcium-containing
  • Vitamin D, particularly calcitriol, to
    decrease the secondary hyperparathyroidism as well as to treat clinical
    rickets or osteomalacia
  • Restoration of renal function by transplantation in order to improve the musculoskeletal manifestations
SCFE occurs frequently in patients with renal osteodystrophy but is not common in other presentations of rickets (Fig. 7.8) (124,125,126). The slip occurs through the metaphy-sis (120,127,128)
and occurs at a younger age, in children who are typically small
because of their chronic disease. Therefore, stabilization of the slip
should permit ongoing growth of the proximal femur if possible.
Unstable slips and avascular necrosis are rare in patients with renal

P.179



osteodystrophy, but avascular necrosis possibly associated with steroid use posttransplant has been reported (128).
If the child is young, the slip severe, and the bone disease has not
yet been treated medically, then traction plus medical treatment have
shown very good results. When considering surgical treatment of the
slipped epiphyses, the high incidence of bilaterality suggests that
both epiphyses should be stabilized. In young patients, a pinning
technique that allows for growth (smooth pins across the physis) can be
considered (127). Hardware cutout, including
pin protrusion into the joint, is more likely with the soft bone of
renal osteodystrophy, but has generally been associated with inadequate
medical control of hyperparathyroidism (127,128).

Figure 7.6 Renal osteodystrophy in an 8-year-old boy. A:
Radiographs of the hand show sclerosis, acroosteolysis, and soft tissue
calcification around the metacarpal phalangeal (MCP) and proximal
interphalangeal (PIP) joints. B: Radiographs of the knees show subperiosteal resorption at the medial border of the proximal tibia.
Figure 7.7 Renal osteodystrophy. A Looser zone is evident (arrow) in the medial femoral diaphysis.
Osteoporosis in Children
Implications for General and Lifelong Health
The National Institutes of Health (NIH) Consensus Panel
(2000) has defined osteoporosis as “a skeletal disorder characterized
by compromised bone strength predisposing to an increasing risk of
fracture.” The Consensus Panel notes that bone strength includes both
bone density and bone quality. Childhood osteoporosis can come from
numerous primary and secondary etiologies, as summarized in Table 7.3.
There are 10 million individuals in the United States
with osteoporosis, and 18 million more with low bone mass are at risk
for osteoporosis (129). We associate
osteoporosis with senescence, and certainly most of the individuals
currently affected are old and not likely to be consulting pediatric
orthopaedists. However, the NIH emphasizes that “sub-optimal bone
growth in childhood and adolescence is as important as bone loss to the

P.180


development
of osteoporosis.” The recommended intake of calcium for children aged 9
to 18 is 1300 mg per day, and it is estimated that only 10% of girls
and 25% of boys meet this minimum requirement (129).
Whereas consumption of dairy-based beverages supplying calcium has
declined, consumption of carbonated beverages has increased (130).
Phosphoric acid is used in cola soft drinks to maintain carbonation,
and teenage girls who drink soft drinks are three to four times more
likely to report fractures than those who do not, the association being
strongest among active girls drinking cola (131).
The pediatric orthopaedic surgeon has at least three reasons to be
interested in osteoporosis. First, osteogenesis imperfecta (OI) is an
example of primary osteoporosis with profound implications for medical
and surgical management during childhood. The understanding of this
condition is advancing rapidly. Second, many children treated by
pediatric orthopaedists already have, or are at risk for, secondary
osteoporosis—especially nonambulatory children with neuromuscular
disorders. Finally, prevention of the most common osteoporosis
associated with aging, and its attendant morbidity, has to include a
focus on acquisition of adequate bone mineral strength in childhood and
adolescence—something pediatric orthopaedic surgeons are ideally
positioned to promote among their patients, in their community, and
nationally (132,133).

Figure 7.8 Renal osteodystrophy in a 12-year-old boy. A:
An anteroposterior pelvis x-ray reveals an early slipped capital
femoral epiphysis on the right. Slipped capital femoral epiphysis is
common in renal osteodystrophy and rare in rickets. B:
Three years after fixation the right proximal femoral epiphysis remains
open and stable; the left hip now shows signs of epiphyseal avascular
necrosis and fragmentation.
TABLE 7.3 CLASSIFICATION OF CHILDHOOD OSTEOPOROSIS
Primary
   Structural gene abnormalities
      Osteogenesis imperfecta
      Marfan syndrome
      Ehlers-Danlos syndrome
      Bruck syndrome
   Genes important in bone development
   Homocystinuria
   Osteoporosis pseudoglioma syndrome
   Idiopathic juvenile osteoporosis

Secondary
   Neuromuscular
      Duchenne muscular dystrophy
      Cerebral palsy
      Myelomeningocele
   Endocrine
      Growth hormone deficiency
      Hyperthyroidism
      Disorders of puberty
   Drug-related
      Glucocorticoids
      Anticonvulsants
      Miscellaneous (methotrexate, heparin, cyclosporine)

Collagen Mutations—Osteogenesis Imperfecta
OI, or brittle bone disease, describes a spectrum of
clinical disorders that share the symptom of abnormal bone fragility.
Most patients with OI have disorders of collagen production that affect
either the quantity or quality of collagen produced. Many of these
disorders can be traced to specific mutations in collagen genes, but
there are myriad such mutations. The phenotype of OI is quite variable.
It includes mildly affected individuals of normal stature without any
skeletal deformities and also those with extensive bony fragility, who
suffer dozens of fractures during childhood, are short in stature, and
have deformed, bowed

P.181



extremities and abnormal facial appearance. The most severe form of OI is fatal in the perinatal period.

The orthopaedic surgeon may be involved in the operative
management of the fractures and deformities that result from OI.
Diagnosis of milder forms of OI among children with frequent fractures
is easy if the sclerae are abnormal (blue or gray) but can be
challenging if they are normal (white). It can be particularly
difficult to distinguish mild OI from inflicted injury.
Clinical Presentation
OI is a rare condition with an estimated prevalence of 1 in 15,000 to 1 in 20,000 children (134).
The hallmark of OI is brittle bones and the tendency to fracture, with
recurrent fractures occurring in childhood, in particular during the
preschool years. Bone pain is a feature in many patients. It is
described as chronic and unremitting and usually relates to old
fractures. Muscle weakness may be variously present. Ligamentous laxity
and joint hypermobility may be present. Wormian bones are present in
the skull in approximately 60% of patients. Abnormal collagen in the
eye leads to the blue or gray-blue sclerae classically associated with
OI. Abnormal collagen in the teeth leads to dentinogenesis imperfecta
(clinically small, deformed teeth which are “opalescent” due to a
higher ratio of transparent enamel to opaque dentin); this is present
in some, but not all, patients with OI (135–140).
The clinical presentation of OI is heterogeneous across
a very wide spectrum. The clinical classification scheme that is most
commonly used is based on the one first proposed by Sillence et al. in
1979 (140). The Sillence classification has
stood the test of time and is a useful way of dividing the phenotype.
This classification has recently been modified (138,141) to incorporate genetic and biochemical abnormalities (Table 7.4). Greater genetic understanding has led to the addition of extra types to the four OI types initially described by Sillence.
Type I Osteogenesis Imperfecta—Mild
Type I OI is nondeforming. Patients are of normal or
low-normal height and do not have limb deformities; they share the
hallmark of bony fragility and often have multiple fractures during
childhood. Fractures become less common after puberty. Blue sclerae are
present in type I OI. Fifty percent of patients also have presenile
deafness (142,143); this presents typically in the third decade of life and therefore is not helpful for diagnosing children (144).
The deafness itself has a conductive component and a sensorineural
component, and is sometimes of sufficient severity to require surgery
for the ossicles of the ears, or cochlear implantation in very severe
cases (145,146).
Fractures which occur in type I OI include spiral and transverse
fractures of long bones, particularly lower extremity bones. In
addition, the avulsion type of fractures, such as olecranon fractures (147)
and patellar fractures, are common and are related to the decreased
tensile strength of the bone because of its underlying low collagen
content.
TABLE 7.4 CLASSIFICATION OF OSTEOGENESIS IMPERFECTA
Type Skeletal Manifestation Sclerae Teeth Collagen Defect
I Mild Blue Normal (IA) or dentinogenesis imperfecta (IB) Quantitative deficiency, but normal collagen
II Lethal Abnormal collagen or severe quantitative deficiency
III Severe White Dentinogenesis imperfecta Abnormal collagen
IV Moderate White Normal (IVA) or dentinogenesis imperfecta (IVB) Abnormal collagen
Type II Osteogenesis Imperfecta—Lethal Perinatal
Type II OI is the most severe form of OI, and babies
with the condition die at or shortly after birth; they are born with
crumpled femora and crumpled ribs accompanied by pulmonary hypoplasia,
which usually leads to death. Central nervous system malformations and
hemorrhages are common due to the markedly abnormal collagen being
produced. Lethal OI can be diagnosed by prenatal ultrasonography.
Short, broad limbs are identified with low echogenicity and low
shadowing, and it is easier to see soft tissue features such as orbits
or arterial pulsations within the fetus. At present, it is not possible
to reliably distinguish on prenatal ultrasound between lethal type II
OI and severe but survivable OI type III described in the subsequent
text. Most patients with the lethal form of OI have blue sclerae,
although some are born with white sclerae (139,140,148,149).
Type III Osteogenesis Imperfecta—Severe
Type III OI is the most severe of the survivable types.
These patients have relatively large skulls but undeveloped facial
bones, leading to a characteristic triangular appearance of the face.
The sclerae of patients with type III OI are described as pale blue at
birth, but become normal in color by puberty. Patients are short, with
severe limb deformities including bowing and coxa vara (Fig. 7.9). Multiple vertebral compression fractures lead to severe scoliosis, kyphosis, and rib

P.182



P.183



cage deformity. Many patients use wheelchairs for mobility, or require
walking aids. Radiographic characteristics include very osteopenic
bones with deformity related to previous fracturing. A characteristic
popcorn appearance of the epiphysis and metaphysis occurs in early
childhood. The pedicles of the vertebrae are elongated. The vertebrae
are wedged and may assume a codfishlike biconcave morphology. Posterior
rib fractures are seen. Additional clinical features may include
basilar invagination of the skull; this can present with headache,
lower cranial nerve palsy, dysphagia, limb hyperreflexia, nystagmus,
hearing loss, or quadriparesis. These patients often have multiple
fractures when they are born, but they do not have the severe thoracic
deformities seen in type II OI. Fractures heal at the normal rate but
recur frequently during childhood, particularly in the preschool years;
some patients with this form of OI have over 100 fractures (148,149).

Figure 7.9 This female infant with severe osteogenesis imperfecta (OI) presented at 19 months of age with a left femoral fracture (A) which was treated in a spica cast and healed (B). A refracture was treated in a spica (C) with progressive varus. At age two, a second refracture occurred through the varus malunion (D) and was treated with open intramedullary (IM) Williams rodding (E).
Three years later the femur is intact and has grown distally, as
evidenced by the position of the rod and by the transverse metaphyseal
lines that occur with each pamidronate treatment cycle (F).
Type IV Osteogenesis Imperfecta—Moderate
Type IV OI describes patients with a moderate clinical
presentation. Most have short stature and many have bowing and
vertebral fractures, although these are not as severe as in patients
with type III OI. Most of the patients are ambulatory, although some
use walking aids. There is a wide range of ages at the first fracture
and number of fractures in patients with type IV OI. Dentinogenesis
imperfecta may or may not be present in these patients. The sclerae are
typically white (139,140).
Additional Types
The four main presentations described by Sillence cover
most of those who present with OI. There is some overlap in the
phenotypes that can make them difficult to distinguish—for instance,
type III and type IV OI. As the genetic defects are understood, at
least three more types of OI have been added, and these do not fit into
the scheme described in preceding text. Type V OI is described as a
hypercallus variety of OI (150). These patients develop profuse amounts of extraosseous callus following their fractures (Fig. 7.10),
and the presentation can be confused on clinical and radiographical
evidence with an osteosarcoma, although type V OI usually occurs in
much younger children. An additional clinical feature is the
ossification of interosseous membranes between the tibia and fibula,
and between the radius and ulna; this leads to the clinical sign of
diminished or absent pronation and supination of the forearm, which can
suggest the correct diagnosis. Type VI OI includes individuals who
phenotypically appear to have moderate or severe OI similar to a type
IV presentation (151). However, these patients
are known to have normal collagen, and they have a defect in new bone
mineralization without having any of the biochemical abnormalities or
growth-plate deformities associated with rickets. The exact etiology of
this condition remains uncertain. Type VII OI was initially described
in a cohort of First Nations individuals from Quebec, Canada (152).
These patients have rhizomelia, and the characteristic deformities are
coxa vara of the long bones. The bone is histologically similar to what
is seen in type I OI. Linkage analysis shows that the defect is on
chromosome 3, and therefore not in a collagen gene.
Figure 7.10
Osteogenesis imperfecta (OI). The excess callus formation around the
distal humerus following injury (no displaced fracture was seen) is
typical of type V OI.
Etiology of Osteogenesis Imperfecta
Most presentations of OI are caused by mutations within
the collagen 1A1 gene found on chromosome 7q, or mutations in the
collagen 1A2 gene found on chromosome 17q. The complete list of
mutations found is kept up-to-date at the OI mutation database at http://www.le.ac.uk/genetics/collagen.
Two different types of mutations produce OI. OI type I, which is the mild form, is a quantitative defect in collagen production resulting from a silenced allele of the collagen 1A1 gene (138,153,154).
This is usually the result of a premature stop codon within the gene
and results in the production of nonsense messenger RNA instead of
proper messenger RNA coding for the preprocollagen of the collagen
molecule. The nonsense messenger RNA is detected and destroyed, and the
result is production of a diminished number of α 1 chains, and
consequently a decreased

P.184


quantity
of normal collagen being produced. With the decreased quantity of
collagen, the bone is weakened and becomes more susceptible to
microfractures. Bone can sense its mechanical environment, and
microfractures cause a new round of bone removal and reformation,
leading to constant activation of bone remodeling in OI. This process
demands an increased transcriptional activity of type I collagen and
induces more osteoclasts and an increase in the excretion of collagen
degradation of products. Bone turns over rapidly, but remains of poor
quality. When growth ceases at puberty, the transcriptional activity
demand is reduced and bone turnover can slow down; then the bone
strength and architecture become closer to normal, and the fracture
rate drops.

OI types II, III, and IV are usually caused by the production of abnormal types of collagen (138,155).
The collagen molecule is a triple helix formed by spontaneous
self-assembly of three long linear procollagen molecules. These
procollagen molecules have a typical repeating amino acid pattern of
glycine XY-glycine XY-glycine XY. Glycine appears at every third
position because it is the smallest amino acid and can be folded into
the interior of the triple helix. Glycine substitutions place a larger
amino acid where the glycine residue belongs, so that the collagen
triple helix cannot assemble appropriately. The collagen molecule
begins assembling at the C-terminal end and assembles toward the
N-terminal end. If the mutation substitutes for a glycine close to the
C-terminus at the beginning of the molecule, then a very short strand
of abnormal gene product is produced, and the corresponding clinical
diseases are severe. If the glycine substitution mutation is toward the
distal N-terminus end, then a longer partial collagen molecule can be
formed, and the clinical phenotype is less severe.
The primary defect in OI is the osteoporosis produced by
an abnormal quantity or quality of collagen. Mechanically the bone is
more ductile, rather than being more brittle (156).
Secondary osteoporosis caused by immobilization following fractures or
surgery, or by decreased physical activity and weight bearing with
severe deformity, adds to the complications of the primary defect.
Prevention of secondary osteoporosis is an important concept when
treating fractures, planning surgery, or recommending general care.
Diagnosis
The clinical approach is the mainstay for the diagnosis
of OI. There is no single laboratory test that distinguishes
individuals with OI from those with normal bone. The clinical features
of severe OI type II and type III are distinct enough that physical
findings and plain radiography are usually sufficient for a diagnosis.
Patients with type I OI have blue or blue-gray sclerae and are readily
identified clinically. Normal babies may have blue sclerae until 1 year
of age, so this finding has diagnostic value only in the older child.
Patients with mild presentations of type IV OI are easy to diagnose if
they have dentinogenesis imperfecta, but those with normal teeth may
benefit from additional investigation, depending on the purpose of
making the diagnosis. Dual energy x-ray absorptiometry (DEXA) scanning
shows low levels of lumbar and femoral bone mineral density (BMD) in
patients with mild OI (157,158). Published values for BMD in healthy normal children are available for comparison (159).
Caution must be exercised in interpreting DEXA scans in children, and
overdiagnosis of osteoporosis is reported to be frequent (160);
this is because DEXA scanning reports BMD per square centimeter surface
area of bone, ignoring the third dimension (thickness of the bone in
the path of the photon), which is larger in adults, thereby leading to
a higher area density even if the true volumetric density were to
remain a constant.
Dozens of individual mutations have been found within
collagen 1A1 and collagen 1A2 genes producing the main phenotypes of
type I, type II, type III, and type IV OI (138,148,149,154,161).
As such, many new patients often have new mutations specific to
themselves. Accordingly, a DNA-based genetic test for OI is not
currently clinically useful. An intermediate level of testing involves
culturing dermal fibroblasts and studying the collagen that they
produce. Quantitatively abnormal collagen production can be detected in
87% of individuals with known OI. Conversely, 13% of individuals with
known OI would be missed by a cultured dermal fibroblast test. One
common question is whether a patient has OI or inflicted trauma. OI is
very rare, and inflicted injury remains significantly more prevalent.
Clinical diagnosis remains a gold standard to distinguish these two
entities, and cultured dermal fibroblast testing is not considered
useful as a routine part of such investigation (162).
In cases with legal implications, positive findings in the past
history, family history, clinical examination, or properly interpreted
DEXA scan may assist in the diagnosis of osteoporosis or OI. OI cannot
be entirely ruled out in patients with negative findings, but is a
highly unlikely diagnosis in the presence of findings suggesting child
abuse (discussed elsewhere). Finally, a diagnosis of OI does not
exclude the possibility of child abuse.
Osteogenesis Imperfecta—Medical Treatment
Cyclical administration of intravenous bisphosphonates has recently become popular in the pharmacologic management of severe
OI, but cannot currently be recommended for mild OI. Bisphosphonates
are widely used drugs based on the pyrophosphate molecule, which is the
only natural inhibitor of bone resorption. The exact mechanism of the
drug is unclear, but its primary action is at the level of the
osteoclast. Glorieux et al. reported the effects of bisphosphonate
treatment in an uncontrolled observational study of 30 patients with
severe OI (163). The intravenous dose given was
3 mg per kg of pamidronate per cycle by slow intravenous infusion at
4-month intervals. All patients were given 800 to 1000 mg calcium and
400 international units of vitamin D per day. There was marked

P.185


improvement
in the clinical status of the patients. BMD increased at an average
rate of 42% per year. There was an increase in the cortical width of
the metacarpals and in the size of the vertebral bodies. The average
number of fractures dropped from 2.3 per year to 0.6 per year. No
nonunions or delayed unions of any fractures were noted. Patients
reported a marked reduction in bone pain 1 to 6 weeks following
initiation of treatment. The only adverse effect noted was the acute
phase reaction comprising fever, back pain, and limb pain on day two of
the first cycle. This adverse effect was treated with acetaminophen and
did not recur with subsequent infusion cycles. Mobility improved in 16
of the 30 patients treated, with no change in the other 14. Growth
rates increased. Patients younger than 3 years showed a faster and more
pronounced response to the bisphosphonate. The direct effect of the
bisphosphonate is decreasing resorption and turnover of bone. The
resulting decrease in bone pain and fractures led to increased weight
bearing and mobility. It is likely that the increased weight bearing
and mobility would have resulted in further strengthening of bone and
muscle. Bisphosphonate treatment has become a standard for severe OI.
It should be noted that this drug does not address the basic
abnormality underlying OI, but it does alter the natural course of the
disease. Radiographs of patients with OI treated with bisphosphonates
show characteristic dense sclerotic lines that form at the growth
plate, one per treatment cycle (164).

Caution must be exercised in extending the indications
for bisphosphonate treatment to patients with milder forms of OI.
Reported clinical results apply only to patients with severe OI.
Osteopetrosis is a reported complication of bisphosphonates in humans (165). Animal studies suggest reduced longitudinal bone growth with these drugs (166,167,168).
Randomized controlled clinical trials are needed before routine
clinical use of bisphosphonates in milder forms of OI is considered.
Other medical treatments for OI include anabolic agents,
specifically human growth hormone. Human growth hormone is an anabolic
agent, and therefore stimulates increased bone turnover, causing a
higher demand for collagen transcription and perhaps exacerbating the
underlying abnormality while attempting to ameliorate the decreased
stature. Because growth hormone may have both beneficial and negative
effects in OI, clinical research results are required before
indications can be stated. We suggest at present that growth hormone be
used in patients with OI only in the context of clinical research
studies.
Osteogenesis Imperfecta—Surgical Treatment
Patients with mild (nondeforming) OI require little
modification of standard surgical treatments, whereas those with severe
(deforming) OI require multiple specialized surgical techniques and
implants.
The treatment of fractures in mild OI must be done in
such a way as to minimize disuse osteoporosis. Simple undisplaced
fractures can be treated with plaster cast or splints which should be
removed early to allow prompt return to function. Avulsion fractures of
the olecranon can be treated by tension band wiring and early motion
with good results (Fig. 7.11).
Severe OI requires specialized techniques. Children with
very fragile lower extremities can be supported with vacuum orthoses to
permit protected weight bearing (169).
Eventually, however, children with severe OI will be considered for IM
rodding of the long bones. This IM rodding is best carried out in
conjunction with medical treatment using bisphosphonates in an
interdisciplinary setting. Some centers propose that lower extremity
rodding be considered as soon as the patient stands and before walking
begins, at about the age of 18 months (148).
Upper extremity rodding is indicated if repeated long bone fractures
cause deformity which affects function, in particular, if the use of
crutches or a walker is compromised (148,170).
Rodding techniques have been developed to correct the
bowed long bones typical of severe forms of OI. Rodding allows for
correction of the deformity, puts the weight-bearing line along the
axis of the bone, is effective in soft bone, and allows for continued
growth.
Sofield et al. first described the technique of open
fragmentation and rodding of the long bones in 1952 and reported the
results in 1959 (171). He described a long open
exposure with removal of the entire diaphysis of the bone, division
with multiple transverse osteotomies, and reassembly over a rod.
Williams (172) developed a threaded end rod
that can be attached to an insertor (which simplifies placement of the
rod in the bone) and then detached. Bailey and Dubow developed a rod
with overlapping male and female shafts allowing for expansion of the
rod during normal bone growth (Fig. 7.12). The
rod is anchored into the epiphysis using a t-piece at each end.
Complications of extensible rods are frequent; Nicholas and James (173)
reported 56 roddings in 16 patients, with 6 failed expansions, 12 t-end
loosenings, 4 rod migrations, 6 bent rods, and 5 fractures. Separation
of the rods after growth led to fracture and bowing in every case, and
close radiographic surveillance was proposed. Comparison of
nonelongating to elongating rods has shown slightly higher complication
rates with elongating rods, although crimping the t-piece may reduce
this complication rate. A newer elongating rod design, the Fassier
Duval rod, has cancellous screw threads at either end to provide stable
anchorage in the epiphysis or metaphysis (174).
In one series, the need for revision due to bone growth and rod
migration was reported as being higher for nonelongating rods (175), but in another series it was reported as being approximately equal to that seen with elongating rods (176). Methods of rod exchange through percutaneous techniques have been described (177,178), and a stereotactic device to assist this has been developed (178) but is not in wide clinical use.
Tiley et al. reported on 129 roddings among 13 children, of whom 11 maintained or gained the ability to ambulate (179). Most reports suggest that ambulation is improved by

P.186



rodding (180,181,182,183), but one emphasizes the possibility of ambulatory status worsening in a large number of patients (184).
Ultimately, the prognosis with regard to walking is much more strongly
influenced by the subtype of OI than by the treatment (185,186,187),
but modern combinations of medical and surgical treatment combined with
rapid advances in the understanding of the biology of the disease may
one day change this situation.

Figure 7.11
Osteogenesis imperfecta (OI). This olecranon avulsion in a 14-year-old
boy with type 1 OI was treated with stable internal fixation and early
return to motion. Olecranon avulsions are frequently associated with a
diagnosis of OI.
Treatment of Scoliosis in Osteogenesis Imperfecta
Scoliosis (Fig. 7.13) can be very challenging to treat in patients with OI (188,189,190,191,192,193,194,195,196,197). Bracing appears to be ineffective at preventing curve progression even when curves are small (194,197).
Patients with large progressive curves may suffer from pulmonary
compromise, and bracing can cause rib deformities that worsen it. It is
unknown whether operative management of the scoliosis leads to improved
pulmonary function, quality of life, and survival. Accordingly,
decision making regarding surgery must be individualized for each
patient. Spinal fusions have a higher complication rate in OI, with 20
patients out of 60 experiencing a total of 33 major complications (194).
The most common complications were blood loss greater than 2.5 L (nine
cases), intraoperative hook pullout (five cases), postoperative hook
pullout (five cases), and pseudarthrosis (five cases). Strategies to
prevent hook pullout include load sharing via segmental
instrumentation, supplementation of hook site bone with methyl
methacrylate, and consideration of fusion without instrumentation.
Preoperative medical treatment (bisphosphonate) to strengthen the bone
is logical but results are not yet reported. Progression of the curve
postoperatively, with or without pseudarthrosis, may occur (194).
The natural history, expectations, likelihood of complications, and
likelihood of success must be carefully assessed for each patient, and
decisions not to embark on surgical reconstruction are sometimes
correct.
Figure 7.12 Osteogenesis imperfecta (OI). A: An extensible rod is used in this 7-year-old girl with severe OI. B: The rod has grown with the femur over a 3-year period.

P.187


Juvenile Idiopathic Osteoporosis
Idiopathic juvenile osteoporosis is a rare condition (133). Onset is typically 2 to 3 years prior to puberty, and patients present with vertebral (Fig. 7.14)
or long bone fractures and bone pain. Fractures are typically
metaphyseal in location. Kyphosis, scoliosis, and pectus carinatum
deformities may be present. BMD is decreased 2.5 standard deviations
below age-appropriate norms. Idiopathic juvenile osteoporosis is a
diagnosis of exclusion; that is, other primary and secondary causes of
osteoporosis must be excluded (Table 7.3).
There are no blood test abnormalities specific to the diagnosis, and
the genetic cause is as yet unknown. Treatment includes optimizing the
intake of calcium and vitamin D and promoting physical activity,
including weight bearing and strength training but avoiding trauma.
Bracing can be used to control vertebral pain and to prevent
progression of kyphotic deformities (198).
Judicious use of bisphosphonate therapy may be considered in severe
cases. A remarkable remission of the condition at puberty is common.
Other Forms of Primary Osteoporosis in Children
Apart from OI and idiopathic juvenile osteoporosis,
there are several other primary causes of reduced BMD in children.
Children with Ehlers Danlos syndrome, Marfan syndrome, and
homocystinuria have reduced bone mass. Those who suffer fractures or
bone pain can also be considered to have a form of primary
osteoporosis, but fractures are not as typical a feature of these
conditions as they are of OI. Like OI, these conditions are based on
known deficiencies in the production of structural proteins. These
conditions are discussed in other chapters.
Bruck syndrome is a rare primary form of osteoporosis.
The phenotype is similar to OI, with thin bones, fractures, bone pain,
and blue or white sclerae. Joint contractures are a distinctive
clinical characteristic (199). The condition is a result of failure to cross-link collagen fibrils, and is specific to bone tissue (200).
Osteoporosis pseudoglioma syndrome (OPPG) is an
autosomal recessive disorder phenotypically similar to OI but
accompanied by congenital blindness due to hyperplasia of the vitreous
humor (201,202). The mutation is in the low-density-lipoprotein-receptor–related protein 5 gene (203). Treatment with bisphosphonates has been successful (204).
Surgical management of fractures with intramedullary rodding techniques
has encountered complications related to severe fragility of the bone (205).
Secondary Osteoporosis in Children
Bone mass responds to biochemical, hormonal, and mechanical signals as described in the first part of this

P.188



chapter. Interference with normal homeostatic mechanisms leads to
reduced bone mass—for example, many children have disuse osteopenia
following fracture treatment. Resumption of load bearing after healing
allows this to normalize. In conditions where weight bearing is
reduced, persistent low bone mass can lead to low-energy fractures and
a cascade of repeated fractures in the same extremity.

Figure 7.13 Osteogenesis imperfecta (OI). Marked spinal deformity in a 5-year-old girl with severe OI.
Neuromuscular Disorders
The central nervous system exerts direct control over
BMD. The fat-derived hormone leptin acts on hypothalamic neurons that
mediate bone mass via the sympathetic nerves (206,207,208,209).
The clinical significance of this recent finding has not been
elucidated, but it may become important in managing osteoporosis and
related regional conditions such as reflex sympathetic dystrophy. At
present, the management of neuromuscular osteoporosis focuses on the
downstream effects of the neuromuscular abnormalities on load bearing
and nutrition.
Low bone mass is observed in patients with cerebral
palsy (CP) (210–214). Typically, it is nonambulatory patients with the
lowest bone masses who are at risk for pathologic and low energy
fractures, but reduced bone mass has been observed in the affected
limbs of ambulatory hemiplegics also (215,216),
thereby suggesting that both weight bearing and muscle forces play
roles in the establishment and maintenance of bone mass. After absence
of ambulation, undernutrition is the second most important predictor of
low bone mass in patients with CP (214,217).
Fractures and a “fracture cascade” of increasing disuse osteoporosis
from treatment-related immobilization can cause significant problems
for many patients (218,219). Fracture prevalence among children and adolescents with moderate to severe CP was 26% in one study (214).
The orthopaedic surgeon should focus as much on the prevention as on
the treatment of osteoporosis in the CP population. Increasing weight
bearing, optimizing nutrition, and minimizing the extent and duration
of immobilization following surgery are important. A controlled trial
of weight bearing in patients with CP showed significant increases in
femoral neck bone mineral content (220). Medical treatments including vitamin D and calcium (221) or bisphosphonates (222)
have also been reported to increase BMD in patients with CP, although
the precise indications for their prescription are unclear.
Approximately one-third of children without CP suffer a fracture during
childhood, so the 26% fracture prevalence among children with CP does
not seem out of line. Accordingly, if weight bearing is easy to achieve
and enjoyable for the child it can be recommended; but if

P.189


extensive
surgery, elaborate standers, and many hours of professional care are
required, then perhaps it is not of benefit to the child.

Figure 7.14
Juvenile idiopathic osteoporosis. This 14-year-old boy has marked
osteoporosis evident on plain radiographs [dual energy x-ray
absorptiometry (DEXA) scans indicated bone mineral densities 2.8
standard deviations below mean] and a healed compression fracture at T7.
Patients with Duchenne muscular dystrophy (DMD) lose
bone mass as the lower extremities weaken, with bone densities 1.6
standard deviations below the mean while they are still walking and 4
standard deviations below the mean once they become nonambulatory. Of
41 boys followed, 18 sustained fractures, 12 of which were lower
extremity fractures, and 4 of which caused loss of ambulation (223).
In boys with DMD, BMD has been reported as lower among those who were
treated with steroids (prednisone) than among untreated boys (224).
Deflazacort is a steroid medication which preserves muscle strength,
ambulation, and respiratory function as well as preventing the onset of
scoliosis. Despite being a glucocorticoid analog, it is reported to
have bone-sparing effects (225) and randomized
trial evidence (in juvenile arthritis) shows less bone loss among
patients treated with deflazacort compared with patients treated with
prednisone (226). Vertebral fractures have been reported in patients treated with deflazacort (227), whereas they are rare among untreated patients (223). No trials of bisphosphonates in DMD have been reported yet.
Endocrine/Metabolic Disorders
Growth Hormone Deficiency
Growth hormone deficiency results in extremely short
stature and low muscle mass. BMD is low in both adults and children,
and adults have a 2.7 times increased fracture rate compared with
normal controls (228). Treating the deficiency
with growth hormone improves longitudinal growth and muscle mass and
also improves calcium absorption, thereby exerting beneficial effects
on the osteoporosis both mechanically and biologically. Improved BMD
with growth hormone treatment has been documented in adults but
requires further study in children (229,230).
It is worth noting that growth hormone deficiency is the
second most common endocrinologic cause of SCFE (after hypothyroidism).
Ninety-two percent of children with growth hormone deficiency and SCFE
present with the slip after growth hormone
treatment has been initiated, perhaps because the growth plate becomes
more active and potentially weaker, mechanically, relative to the
larger size of the body. The prevalence of bilaterality in slips with
endocrinopathies has been reported to be as high as 61%, and therefore
prophylactic pinning of the contralateral side has been suggested by
some specialists (231). Among 2922 children
followed prospectively while receiving growth hormone in Australia and
New Zealand, only 10 slipped epiphyses were reported by clinicians
performing active surveillance for complications (232).
Hyperthyroidism
Hyperthyroidism leads to bone loss through increased
bone turnover. T3 directly stimulates osteoblasts, resulting in linked
osteoclast activation and bone resorption. The increase in serum
calcium and phosphate suppresses PTH and 1,25-hydroxy vitamin D
production, which then decreases calcium and phosphate absorption and
increases calcium excretion. The net result is bone resorption in a
high turnover state (133). Treatment of the underlying disease to achieve a euthyroid state is the appropriate management for the osteoporosis.
Disorders of Puberty
One third to one half of adult bone mass is accrued
during the pubertal years, and sex steroids are necessary for this
accelerated bone mineral accrual. Mineral accrual lags behind the
longitudinal growth spurt by a year or two, perhaps partially
explaining the increased fracture rate documented during and just after
the growth spurt (233). Over the last 30 years
there has been a statistically significant increase in the incidence of
distal forearm fractures in adolescents, but it is unclear whether this
relates to changes in diet or in physical activity (234).
Precocious puberty results in accelerated growth and
bone mineral accrual. It can be treated with gonadotropin-releasing
hormone analogs to preserve epiphyseal function and allow attainment of
greater final height. Such treatment does not adversely affect peak
bone mass. Constitutionally delayed puberty in boys is associated with

P.190



lower BMD, but it is unclear whether testosterone treatment improves overall mineral acquisition (133).

Athletic amenorrhea is common among young women training
for sport. Thirty-one percent of college level athletes who were not
using oral contraceptives reported athletic amenorrhea or
oligomenorrhea (235). Athletic amenorrhea plus disordered eating plus osteoporosis has been dubbed the “female athlete triad” (236,237). The full triad is rare among athletes (238) and was found to be nonexistent among female army recruits (239).
Not all sports are equal in propensity for bone loss. Gymnasts exhibit
higher bone mass than do runners despite similar prevalance of
amenorrhea (240); gymnasts have both higher muscle strength (241) and higher serum insulinlike growth factor 1 (IGF-1) (242)
than do runners, and these are both protective of bone mass. It must be
remembered that for the general population, more exercise means more
bone, as has been demonstrated in randomized trials of physical
activity in prepubescent and pubescent children (243,244).
Untreated anorexia nervosa is associated with 4% to 10% trabecular and cortical bone loss per year (245).
The long-term prevalence of fractures amongst people with anorexia
nervosa is 57% at 40 years follow-up, 2.9 times higher than that in the
general population (246). Estrogen alone is insufficient to restore BMD, particularly if nutrition is compromised (247,248).
Restoration of body mass, provision of adequate calcium and vitamin D,
and correction of the hormonal environment should all occur.
Primary amenorrhea from pure dysfunction of the
hypothalamic–pituitary axis was associated with osteoporosis in 3 and
osteopenia in 10 of 19 girls (ages 16 to 18), compared with 0 among 20
controls with regular cycles (249).
Drug-related Effects
Glucocorticoids
Steroids are commonly used to treat a variety of acute
and chronic medical conditions. Frequent short courses of oral
glucocorticoids (for asthma exacerbation) have been shown not to
adversely affect BMD in children (250), but
chronic glucocorticoid use is a well-established cause of osteoporosis
and is common among patients with juvenile arthritis, leukemia, and
organ transplantation. The underlying mechanism includes osteoblast
apoptosis and decreased intestinal calcium absorption and renal
reabsorption (133). Deflazacort is a bone-sparing glucocorticoid (225),
and randomized trial evidence (in juvenile arthritis) shows less bone
loss among patients treated with deflazacort than among patients
treated with prednisone (226). In adults using
glucocorticoids, there are now multiple trials demonstrating that
bisphosphonates are useful for prevention and treatment of bone loss
and for prevention of fractures (251,252,253,254). Case series in children have shown similar promising results (255).
Research is active in this area and it is likely that prevention and
treatment trials with children on long-term steroids will further
clarify the appropriate role for bisphosphonates, as well as for novel
agents such as subcutaneous PTH (256,257).
Anticonvulsants
An association between altered bone mineral metabolism and anticonvulsant drugs has long been discussed (258,259,260,261,262,263,264),
although the exact mechanism by which anticonvulsants interfere with
bone metabolism remains unknown. Induction of liver enzymes with
consequent increased vitamin D catabolism has been proposed (265,266), but other studies have reported vitamin D metabolism to be normal (267,268). There is laboratory evidence that phenytoin and carbamazepine exert a direct effect on bone cells (269).
Quantitative computerized tomographic studies of the distal radius in
patients using carbamazepine or valproic acid (for isolated epilepsy
without CP) showed decreased trabecular BMD but a compensatory increase
in cortical BMD in a high bone turnover state (270).
A controlled trial has shown that treatment with vitamin D and calcium
increases BMD in children with severe CP who are receiving
anticonvulsants (221).
Miscellaneous Factors
Other drugs reported to cause osteopenia and
osteoporosis in children include methotrexate, cyclosporine, and
heparin. Little information is available on prevention and treatment of
iatrogenic osteoporosis in children (133).
SCLEROSING BONE CONDITIONS IN CHILDREN
Osteopetrosis
Osteopetrosis is the most common and the most well known
of the sclerosing bony dysplasias. All of these conditions are
characterized by an imbalance between the formation and the resorption
of bone, favoring formation. In osteopetrosis, it is decreased or
completely failed resorption of bone due to an osteoclast defect that
tips the balance in favor of formation.
Three clinical presentations of osteopetrosis are traditionally described (271).
The infantile or autosomal recessive form is the most severe and can be
fatal in the first decade of life. Patients present with pathologic
fractures or with an exceedingly dense radiographic appearance of the
bones (Fig. 7.15); they may have cranial nerve
problems including optic nerve compression and blindness, facial nerve
dysfunction, and sensorineural deafness due to narrowing of the skull
foramina; they may have bone pain related to fractures or stress
fractures; and they frequently have infections of the bones or
mandibles. Searching nystagmus is described. Patients are pale because
of anemia and hepatosplenomegaly caused by extramedullary
hematopoeisis. A prominent forehead and broad upper skull with
hypertelorism are common features.
The milder autosomal dominant adult form of osteopetrosis can present with pathologic fractures and dull pain,

P.191



P.192



and patients have mild degrees of short stature and prominent
foreheads. Other individuals with the autosomal dominant form are so
mildly affected that the diagnosis is made incidentally on the basis of
the very curious appearance of the bones.

Figure 7.15 Six-month-old male infant with severe osteopetrosis and pancytopenia. A to E: Dense sclerotic bones at the pelvis (A), humerus (B), and forearm (C), without evident medullary cavities. D and E: After successful bone marrow transplant the bony architecture in the humerus (D) and forearm (E) were normalized.
An intermediate clinical type of osteopetrosis is
described; it includes patients with more severe phenotypes who survive
into adulthood. Classifications based on the identification of genetic
defects have expanded the list to at least six types of osteopetrosis,
but the genetic classification is not currently comprehensive enough to
replace the clinical classification of the disease (272).
The striking radiographic characteristic of
osteopetrosis is that the bones are dense white, without medullary
cavities, and appear marble-like. A bone within bone or endobone
appearance may be seen within the pelvis. Marked sclerosis of the end
plates of the vertebral bodies may give vertebrae a rugger jersey
appearance (Fig. 7.16). There may be
significant sclerosis of the base of the skull. Failure of metaphyseal
cutback by osteoclasts leads to an Erlenmeyer flask shape of the
epiphysis. Secondary deformities including progressive coxa vara or
apex lateral bowing of the femur may be present (271,273,274).
Figure 7.16 The classic rugger jersey appearance of the spine is seen in this 15-year-old girl with osteopetrosis.
Etiology
Osteopetrosis is caused by specific osteoclast defects
that prevent the osteoclast from carrying out its normal function in
reabsorption of bone. Many specific defects have been identified. These
include carbonic anhydrase deficiency, which prevents the osteoclasts
from acidifying the extracellular space at the ruffled border. This is
associated with the milder clinical forms and may also be associated
with distal renal tubular acidosis and with intercranial calcifications
that are unique to this form (275,276,277). Fatal infantile forms of osteopetrosis have been associated with defects in the genes coding for proton pump or chloride

P.193



channel protein production (278). More recent case reports describe iatrogenic osteopetrosis as being the result of bisphosphonate treatment (165).

All causes of osteopetrosis decrease the function of
osteoclasts in the reabsorption of bone. The histologic hallmark of
osteopetrosis is remnants of primary spongiosa within the bone. These
are areas of calcified cartilage, which are precursors to enchondral
bone formation and are normally removed during the first pass
remodeling in fetal life. The ongoing failure to resolve and remodel
bone leads to the skeletal manifestations of the disease, limits the
narrow space available, contributes to pancytopenia and
hepatosplenomegaly, and is also responsible for the cranial nerve
compression at the base of the skull. The predisposition toward
musculoskeletal infections is thought to be caused by a combination of
the abnormal bony architecture and the insufficiency of white blood
cells.
Medical Treatment
Bone marrow transplantation for the severe forms of infantile osteopetrosis was first described in 1977 (279) and has since been widely reported (280,281,282).
Successful bone marrow transplantation corrects both the skeletal and
hematologic abnormalities associated with osteopetrosis. However, not
all bone marrow transplantations are successful and not all patients
have survived. Additional medical treatment can include very high doses
of calcitriol to stimulate osteoclastic activity and the administration
of interferon γ to stimulate superoxide production by osteoclasts.
Long-term therapy with interferon γ in patients with osteopetrosis
increases bone resorption and hematopoeisis and improves leukocyte
function (283).
Orthopaedic Treatment
Most patients presenting to pediatric orthopaedists do so because of fractures (284). Most fractures in children with osteopetrosis will heal well if treated with closed means, although healing may be delayed (Fig. 7.17).
The principal exceptions are fractures of the femoral neck and
intratrochanteric region, which can be extremely difficult to manage (Fig. 7.18). Armstrong’s survey of the experience of the Pediatric Orthopaedic Society of North America (284)
described a high incidence of nonunion and varus deformity of the femur
if femoral neck and intertrochanteric fractures were treated by closed
means. The results of early operative treatment of femoral neck and
intertrochanteric fractures were good, but technical difficulties
involved in obtaining fixation in the dense bone were noted. Worn-out
or broken drills and even drivers were reported. Subtrochanteric and
femoral shaft fractures did well with both closed and open management.
Most tibial fractures were managed closed. Multiple fractures of the
tibia were reported in some patients, but could be well managed with
casts.
Cervical spine fractures are rare and involve the
posterior elements. These fractures are treated with immobilization.
Lumbar spondylolysis and listhesis have also been described (284,285)
and have responded well to nonoperative treatment in children and
adolescents, though occasionally requiring spinal fusion in adults.
Acquired coxa vara can be treated with proximal femoral
valgus osteotomies fixed by conventional plates or screws, but with
technical difficulties similar to those encountered in treating femoral
fractures.
Some patients with osteopetrosis get osteoarthritis of
the hip and knee during midlife. These can be treated with total knee
arthroplasty and total hip arthroplasty, although difficulties with
reaming the canal and cutting the bone surfaces have been noted.
Caffey Disease
Caffey disease, or infantile cortical hyperostosis,
characteristically presents between the ages of 6 weeks and 6 months.
The clinical features are those of an irritable child, sometimes with
fever, and with tender soft tissue swelling over the affected bone.
Radiographically, there is abundant subperiosteal new bone and,
eventually, thickening of the cortex (Fig. 7.19).
The underlying cause remains unknown. The disease is usually
self-limiting and may recur episodically, but typically resolves by the
age of 2 years, and in most cases does not require any active
orthopaedic treatment. Laboratory investigation may show an increased
erythrocyte sedimentation rate, an increased white blood cell count, an
increased level of alkaline phosphatase, and iron deficiency anemia.
Differential diagnoses include child abuse, infection, and metastatic
neoplasms. John Patrick Caffey was a radiologist who initially
described the condition while working on the radiographic presentation
of child abuse. Table 7.5 presents a listing of other causes of periosteal reaction and cortical thickening in infancy and early childhood.
Sporadic reports propose that there is a lethal form of Caffey disease that is fatal in either the perinatal (286) or infantile (287)
age group. Whether this remains part of the spectrum of Caffey disease
depends on the eventual elucidation of the underlying cause.
The recurrent forms of Caffey disease have been successfully treated with anti-inflammatory drugs including naproxen (288) or indomethacin (289).
The rationale for treating Caffey disease with nonsteroidal
antiinflammatory drugs is supported by the clinical observation that a
bone lesion and hyperostosis similar to that in Caffey disease is also
seen in children who are treated with a prolonged course of
prostaglandin E for patent ductus arteriosus (169,290).
Pyknodysostosis
Pyknodysostosis is similar to osteopetrosis in that it is a manifestation of a failure of bone resorption (Fig. 7.20). Patients with this condition do not produce cathepsin K,

P.194



which normally degrades bone proteins during bone resorption (291,292,293,294).
This is an inherited autosomal recessive condition. Patients are
somewhat short, with deformities including pectus excavatum,
kyphoscoliosis, an oblique angle of the mandible, failure of fusion of
the sutures of the skull in adulthood, proptosis, oblique nose, and
frontal bossing (295,296,297). Growth hormone therapy has been used for improving stature (297). Orthopaedic management is similar to that for osteopetrosis.

Figure 7.17 A to C: This 15-year-old boy with osteopetrosis sustained a fractured femoral shaft from a fall while running (A). He was successfully treated by traction (B) followed by spica casting (C). D: One-year follow-up shown.
Overproduction of Bone by Osteoblasts
There are several rare conditions in which osteoblasts
overproduce bone because of abnormalities in regulation. These
conditions are caused by defects in the transforming growth factor β
superfamily of proteins, which are known to regulate bone growth. The
three clinical conditions are melorheostosis, Camurati-Engelmann
disease, and sclerosteosis.
Figure 7.18
This 12-year-old girl with osteopetrosis has had bilateral femoral neck
fractures treated with open reduction and internal fixation. The left
has united but the right has had nonunion and cutout despite revision
of hardware. The bone has been likened to chalk—dense but
brittle—leading to a high rate of complications associated with
internal fixation.

P.195


Melorheostosis is best remembered by its Greek name,
which describes flowing wax on a burning candle. The condition is
characterized by new periosteal and endosteal bone that resembles
dripping wax radiographically. It typically affects one limb or one
side and produces irregularities of cortical bone, with a lesion or
rash overlying the skin. The etiology is believed to be down-regulation
of betaIG-H3 (298). Somatic mosaicism is thought to explain the usual single limb manifestation (299).
There may be stiffness in the soft tissues and shortening or
contracture of the affected extremity, and the lesion may present with
pain (300). Sarcomas arising in melorheostotic lesions have been reported (301,302,303). Ilizarov treatment has been useful in treating deformity and limb shortening (304,305), but complications relating to abnormal bone and soft tissue have been recorded (306).
Soft tissue contractures are often more problematic for the patient
than the bony deformity is, but they are difficult to treat
operatively, having an over 50% recurrence rate and associated
complications including distal ischemia (300).
Figure 7.19 The ulna is the most frequently affected bone in the extremities of patients with Caffey disease.
Camurati-Engelmann disease is an autosomal dominant
condition characterized by hyperostosis, typically with accumulation of
bone within the medullary canal and the diaphyseal region, and also in
the skull. Clinical features include an enlarged head, proptosis, thin
limbs, weak proximal muscle sometimes resulting in a waddling gait, and
musculoskeletal pain (296,307).
Spontaneous improvement at puberty has been described. Cranial nerve
abnormalities and increased intracranial pressure are possible
sequelae. The disorder is caused by an excess of active TGF β1 which
leads to continuous stimulation of osteoblastic bone deposition (272,308,309).
TABLE 7.5 CAUSES OF CONDITIONS ASSOCIATED WITH PERIOSTEAL REACTION AND CORTICAL THICKENING IN INFANCY AND EARLY CHILDHOOD
Cause Time of Presentation Characteristics
Physiologic periosteal reaction of newborn Age 1-6 mo Thin, even periosteal reaction symmetric along femora, tibiae, humeri
Congenital or genetic condition
Menkes kinky-hair syndrome Newborn Failure to thrive; X-linked
defective copper absorption; boys; sparse, kinky hair; central nervous
system degeneration, metaphyseal fractures, and periosteal reaction;
can be mistaken for abuse or rickets
Camurati-Engelmann
   (diaphyseal dysplasia)
Age 4-6 yr Autosomal dominant;
progressive midshaft thickening of long bones; waddling gait; normal
laboratory findings, except slight elevation of alkaline phosphatase
Infection
Osteomyelitis Any age Classic bacterial
osteomyelitis with lytic or blastic changes at metaphysis and
periosteal reaction as disease progresses; elevated ESR; viral and
fungal types exist; Salmonella osteomyelitis in sickle cell disease may begin at diaphysis; ESR not elevated
Congenital syphilis Age > 3 mo (severe spirochetal infection can cause fetal loss) Many manifestations possible;
osteochondritis with metaphyseal lytic lesions; diaphyseal osteitis;
periostitis; positive serology for syphilis
Inflammatory disease
Juvenile chronic arthritis Age 5-10 yr Periarticular reaction at phalanges, metacarpals, and metatarsals
Trauma
Accidental or nonaccidental injury
Any age Accidental injury should
result in local reaction consistent with age-appropriate activities
(i.e., single tibial reaction 7-10 d after injury in a child who is
walking); nonaccidental injury can result in multiple areas of
periosteal reaction inconsistent with age-appropriate activities
Burns Weeks to months after burn Usually a local response, but can have elements of hypertrophic osteoarthropathy
Metabolic disease
Hypertrophic osteoarthropathy Any age Usually associated with
physiologic abnormality in pulmonary, cardiac, or gastrointestinal
system; tibial, fibular, radial, ulnar diaphyseal and metaphyseal
involvement; clubbing
Hypervitaminosis A Age >9 mo Periosteal reaction in long
bones, typically ulnae and metatarsals; epiphyseal and metaphyseal
ossification abnormalities and physeal lesions possible; elevation of
serum vitamin A level
Scurvy (hypovitaminosis C) Age >9 mo Epiphyseal and metaphyseal
changes most prominent around knees, with associated subperiosteal
hemorrhages; abnormal vitamin C levels in serum or blood
Hyperphosphatemia Usually in second and third decades of life, but reported also in childhood and infancy Associated with tumoral calcinosis; periostitis in tubular bones; calcified soft tissue masses; elevated serum phosphorus level
Healing phase of rickets After treatment of rickets Periosteal reaction adjacent to healing growth plates
Prostaglandin-induced hyperostosis After weeks of prostaglandin E1 administration to maintain ductus patency in congenital heart disease Symmetric, diffuse periosteal reaction in long bones and ribs; elevated alkaline phosphatase level
Neoplasm
Acute leukemia 2-5 yr Diffuse osteopenia, metaphyseal rarefactions; commonly, symmetric periosteal reaction in long bones
Metastatic neuroblastoma, retinoblastoma Similar to leukemia, with presentations <2 yr Similar to leukemia
Idiopathic
Caffey disease 6 wk-6 mo Mandibular involvement,
asymmetric involvement of clavicle, scapula, ribs, or tubular bones
without associated lytic lesions; elevated alkaline phosphatase level
and ESR
ESR, erythrocyte sedimentation rate.

P.196



P.197


Sclerosteosis is a rare condition inherited in an autosomal recessive pattern (310).
It is characterized by skeletal overgrowth, particularly of the skull
and of the mandible. In addition, patients have sclerotic long bones
and gigantism. Bony syndactyly is a characteristic clinical marker (311,312,313,314,315,316).
The increased intracranial pressure associated with this condition may
lead to sudden death, so recognition of this condition is important (317). The underlying defect is loss of function and mutation in the SOST gene (318,319). No specific medical treatment is currently available.
Figure 7.20 Pyknodysostosis. This 12-year-old boy has the dense bones and acroosteolysis characteristic of pyknodysostosis.
REFERENCES
Introduction
1. Karsenty G. The complexities of skeletal biology. Nature 2003; 423:316–318.
2. Kronenberg HM. Developmental regulation of the growth plate. Nature 2003;423:332–336.
3. Siebler
T, Robson H, Shalet SM, et al. Glucocorticoids, thyroid hormone and
growth hormone interactions: implications for the growth plate. Horm Res 2001;56:7–12.
Factors that Regulate Bone Density
4. Harada S, Rodan GA. Control of osteoblast function and regulation of bone mass. Nature 2003;423:349–355.
5. Troen BR. Molecular mechanisms underlying osteoclast formation and activation. Exp Gerontol 2003;38:605–614.
6. Khosla S. Minireview: the OPG/RANKL/RANK system. Endocrinology 2001;142:5050–5055.
7. Doggrell SA. Present and future pharmacotherapy for osteoporosis. Drugs Today 2003;39:633–657.
8. Fourman P, Royer P. Calcium metabolism and bone, 2nd ed. Philadelphia, PA: FA Davis Co, 1968.
9. Rasmussen H. The calcium messenger system. N Engl J Med 1986;314:1094.
10. Avioli LV, Haddad JG. Progress in endocrinology and metabolism. Vitamin D: current concepts. Metab Clin Exp 1973;22:507.
11. Hoffman WS. The biochemistry of clinical medicine, 4th ed. Chicago, IL: Mosby Year Book, 1970.
12. Morgan B. Osteomalacia, renal osteodystrophy, and osteoporosis. Springfield, IL: Charles C. Thomas, 1973.
13. Widdowson EM, McCance RA. The metabolism of calcium, phosphorus, magnesium and strontium. Pediatr Clin North Am 1965;12:595.
14. Raisz LG. Bone metabolism and calcium regulation. In: Avioli LV, Krane SM, eds. Metabolic bone disease. New York: Academic Press, 1978.
15. Borle AB. Membrane transfer of calcium. Clin Orthop 1967; 52:267.
16. DeLuca HF, Vitamin D. New horizons. Clin Orthop 1971;78:423.
17. Arnaud C, Fischer J, Rasmussen H. The role of the parathyroids in the phosphaturia of vitamin D deficiency. J Clin Invest 1964:43.
18. Arnaud C, Tsao HS, Littledike T. Calcium homeostasis, parathyroid hormone and calcitonin: preliminary report. Mayo Clin Proc 1970;45:125.
19. Arnaud CD, Tenenhouse AM, Rasmussen H. Parathyroid hormone. Annu Rev Physiol 1967;29:349.
20. Harrison
HE, Harrison HC. The interaction of vitamin D and parathyroid hormone
on calcium, phosphorus and magnesium homeostasis in the rat. Metab Clin Exp 1964;13:952.

P.198


21. Rasmussen H. Ionic and hormonal control of calcium homeostasis. Am J Med 1971;50:567.
22. Bijovet OLM. Kidney function in calcium and phosphate metabolism. In: Avioli LV, Krane SM, eds. Metabolic bone disease. New York: Academic Press, 1977:49.
23. Harmeyer J, DeLuca HF. Calcium-binding protein and calcium absorption after vitamin D administration. Arch Biochem Biophys 1969;133:247.
24. Beal VA. Calcium and phosphorus in infancy. J Am Diet Assoc 1968;53:450.
25. Blunt JW, DeLuca HF. The synthesis of 25-hydroxycholecalciferol: a biologically active metabolite of vitamin D3. Biochemistry 1969;8:671.
26. Blunt JW, DeLuca HF, Schnoes HK. 25-Hydroxycholecalciferol: a biologically active metabolite of vitamin D3. Biochemistry 1968;7:3317.
27. Blunt JW, Tanaka Y, DeLuca HF. The biological activity of 25-hydroxycholecalciferol: a metabolite of vitamin D3. Proc Natl Acad Sci U S A 1968;61:1503.
28. Jones G, Schnoes HK, DeLuca HF. Isolation and identification of 1,25 dihydroxy vitamin D2. Biochemistry 1975;14:1250.
29. Lund J, DeLuca HF. Biologically active metabolite of vitamin D from bone, liver and blood serum. J Lipid Res 1966;7:739.
30. Cheeney RW. Current clinical applications of vitamin D metabolite research. Clin Orthop 1981;161:285.
31. Holick
MF, Schnoes HK, DeLuca HF, et al. Isolation and identification of
24,25-hydroxycholecalciferol, a metabolite of D3 made in the kidney. Biochemistry 1972;11:4251.
32. Holick
MF, Schnoes HK, DeLuca HF, et al. Isolation and identification of
1,24-dihydroxycholecalciferol: a metabolite of vitamin D active in
intestine. Biochemistry 1971;10:2799.
33. DeLuca HF. Parathyroid hormone as a trophic hormone for 1,25-dihydroxy vitamin D3, the metabolically active form of vitamin D. N Engl J Med 1972;287:250.
34. Boyle
IT, Gray RW, DeLuca HF. Regulation by calcium of in vivo synthesis of
1,25-dihydroxycholecalciferol and 21,25-dihydroxycholecalciferol. Proc Natl Acad Sci U S A 1971;68:2131.
35. Fraser DR, Kodicek E. Regulation of 25-hydroxycholecalciferol-1-hydrolase activity in kidney by parathyroid hormone. Nature 1973;241:163.
36. Rasmussen H, Wong M, Bikle D, et al. Hormonal control of 25-hydroxycholecalciferol to 1,25-dihydroxycholecalciferol. J Clin Invest 1972;51:2502.
37. Gartner
LM, Greer FR. Section on Breastfeeding and Committee on Nutrition.
American Academy of Pediatrics. Prevention of rickets and vitamin D
deficiency: new guidelines for vitamin D intake. Pediatrics 2003;111(4 Pt 1):908–910.
38. Lin R, White JH. The pleiotropic actions of vitamin D. Bioessays 2004;26:21–28.
39. Food and Nutrition Board. Recommended dietary allowances, 7th ed. Washington, DC, 1968.
40. Committee on Nutrition (AAP). Calcium requirements of infants, children, and adolescents. Pediatrics 1999;104:1152–1157.
41. Avioli LV. Intestinal absorption of calcium. Arch Intern Med 1972;129:345.
42. Bruce HM, Callow RK. Cereals and rickets: the role of inositalhexaphosphoric acid. Biochem J 1934;28:517.
43. Wills MR, Phillips JB, Day RC, et al. Phytic acid and nutritional rickets in immigrants. Lancet 1972;1:771.
44. DeToni G. Renal rickets with phospho-gluco-amino renal diabetes. Ann Paediatr 1956;187:42.
45. Tryfus H. Hepatic rickets. Ann Paediatr 1959;192:81.
46. Raisz LG, Rodan GA. Pathogenesis of osteoporosis. Endocrinol Metab Clin North Am 2003;32:15–24.
47. Manolagas SC, Kousteni S, Jilka RL. Sex steroids and bone. Recent Prog Horm Res 2002;57:385–409.
48. Vanderschueren D, Bouillon R. Androgens and bone. Calcif Tissue Int 1995;56:341–346.
49. Burman KD. Thyroid disease and osteoporosis. Hosp Pract 1997;32:71–73, 78–85.
50. Ramos-Remus C, Sahagun RM, Perla-Navarro AV. Endocrine disorders and musculoskeletal diseases. Curr Opin Rheumatol 1996;8:77–84.
51. Sambrook PN. Corticosteroid induced osteoporosis. J Rheumatol Suppl 1996;45:19–22.
52. Adachi JD. Corticosteroid-induced osteoporosis. Int J Fertil Womens Med 2001;46:190–205.
53. Wallach S, Farley JR, Baylink DJ, et al. Effects of calcitonin on bone quality and osteoblastic function. Calcif Tissue Int 1993; 52:335–339.
54. Ferretti JL, Cointry GR, Capozza RF, et al. Bone mass, bone strength, muscle-bone interactions, osteopenias and osteoporoses. Mech Ageing Dev 2003;124:269–279.
55. Berard
A, Bravo G, Gauthier P. Meta-analysis of the effectiveness of physical
activity for the prevention of bone loss in postmenopausal women. Osteoporos Int 1997;7:331–337.
56. Brighton
CT, Strafford B, Gross SB, et al. The proliferative and synthetic
response of isolated calvarial bone cells of rats to cyclic biaxial
mechanical strain. J Bone Joint Surg Am 1991;73:320.
57. Greco F, DePalma L, Speddia N, et al. Growth plate cartilage metabolic response to mechanical stress. J Pediatr Orthop 1989;9:520.
Factors that Regulate Growth Plate Chondrocytes
58. Hueter C. Anatomische studien an den extremitatengelenken neugebornener und erwachsener. Virchows Arch 1862;25:572.
59. Ohlsson C, Isgaard J, Tornell J, et al. Endocrine regulation of longitudinal bone growth. Acta Paediatrica (Oslo, Norway, 1992) 1993;82(Suppl 391):33–40; discussion 41.
Diseases of Bone
60. Whistler D. Disputatio medica inauguralis de morbo puerili anglorum quem patrio idiomate indigenae vocant the rickets. London: Wilhemi, Christiani, Boxii, 1645.
61. Glisson R. De
rachitide sive marbo puerili qui vulgo The Rickets Dicitur Tracttatus.
Adscitis in operis societatem Georgio Bate et Ahasuero Regemortero
. London: G Du-Gardi, 1650.
62. Pettifor JM. Nutritional Rickets. In: Glorieux F, Pettifor JM, Juppner H, eds. Pediatric bone: biology and diseases. San Diego, CA: Academic Press, 2003:541.
63. Buchanan
N, Pettifor JM, Cane RD, et al. Infantile apnoea due to profound
hypocalcaemia associated with vitamin D deficiency. A case report. S Afr Med J 1978;53:766–767.
64. Fraser DR, Salter RB. The diagnosis and management of the various types of rickets. Pediatr Clin North Am 1958;26:417.
65. Arnstein AR, Frame B, Frost HM. Recent progress in osteomalacia and rickets. Ann Intern Med 1967;67:1296.
66. Smith R. The pathophysiology and management of rickets. Orthop Clin North Am 1972;3:601.
67. Laditan AA, Adeniyi A. Rickets in Nigerian children: response to vitamin D. J Trop Med Hyg 1975;78:206.
68. Harrison HE, Harrison HC. Rickets then and now. J Pediatr 1975;87:1144.
69. Goel KM, Sweet EM, Logan RW, et al. Florid and subclinical rickets among immigrant children in Glasgow. Lancet 1976; 1:1141–1145.
70. Mughal Z. Rickets in childhood. Semin Musculoskelet Radiol 2002; 6:183–190.
71. Pettifor JM. Nutritional rickets. In: Glorieux F, Pettifor JM, Juppner H, eds. Pediatric bone: biology and diseases. San Diego, CA: Academic Press, 2003:541.
72. Steinbach HG, Noetzli M. Roentgen appearance of the skeleton in osteomalacia and rickets. AJR Am J Roentgenol 1964;91:955.
73. Opie WH, Muller CJ, Kamfer H. The diagnosis of vitamin D deficiency rickets. Pediatr Radiol 1975;3:105–110.
74. Holm IA, Econs MJ, Carpenter TO. Familial hypophosphatemia and related disorders. In: Glorieux F, Pettifor JM, Juppner H, eds. Pediatric bone: biology and diseases. San Diego, CA: Academic Press, 2003:603.
75. Harris NS, Crawford PB, Yangzom Y, et al. Nutritional and health status of Tibetan children living at high altitudes. N Engl J Med 2001;344:341–347.
76. Salimpour R. Rickets in Tehran: study of 200 cases. Arch Dis Child 1975;50:63.
77. Majid Molla A, Badawi MH, al-Yaish S, et al. Risk factors for nutritional rickets among children in Kuwait. Pediatr Int 2000; 42:280–284.
78. el Hag AI, Karrar ZA. Nutritional vitamin D deficiency rickets in Sudanese children. Ann Trop Paediatr 1995;15:69–76.
79. Biser-Rohrbaugh A, Hadley-Miller N. Vitamin D deficiency in breast-fed toddlers. J Pediatr Orthop 2001;21:508–511.

P.199


80. Feldman KW, Marcuse EK, Springer DA. Nutritional rickets. Am Fam Physician 1990;42:1311–1318.
81. Carvalho
NF, Kenney RD, Carrington PH, et al. Severe nutritional deficiencies in
toddlers resulting from health food milk alternatives. Pediatrics 2001;107:E46.
82. Rudolf M, Arulanantham K, Greenstein RM. Unsuspected nutritional rickets. Pediatrics 1980;66:72–76.
83. Marie PJ, Pettifor JM, Ross FP, et al. Histological osteomalacia due to dietary calcium deficiency in children. N Engl J Med 1982;307:584–588.
84. Clements MR, Johnson L, Fraser DR. A new mechanism for induced vitamin D deficiency in calcium deprivation. Nature 1987;325:62–65.
85. Henderson
JB, Dunnigan MG, McIntosh WB, et al. Asian osteomalacia is determined
by dietary factors when exposure to ultraviolet radiation is
restricted: a risk factor model. Q J Med 1990;76:923–933.
86. Henderson
JB, Dunnigan MG, McIntosh WB, et al. The importance of limited exposure
to ultraviolet radiation and dietary factors in the aetiology of Asian
rickets: a risk-factor model. Q J Med 1987;63:413–425.
87. Kruse K. Pathophysiology of calcium metabolism in children with vitamin D-deficiency rickets. J Pediatr 1995;126(5 Pt 1): 736–741.
88. Shah BR, Finberg L. Single-day therapy for nutritional vitamin D-deficiency rickets: a preferred method. J Pediatr 1994;125:487–490.
89. Mankin HJ. Metabolic bone disease. J Bone Joint Surg Am 1994;76:760.
90. Burnett
CH, Dent CE, Harper C, et al. Vitamin D-resistant rickets: analysis of
twenty-four pedigrees with hereditary and sporadic cases. Am J Med 1964;36:222.
91. Holm IA, Econs MJ, Carpenter TO. Familial hypophosphatemia and related disorders. In: Glorieux F, Pettifor JM, Juppner H, eds. Pediatric bone: biology and diseases. San Diego, CA: Academic Press, 2003:603.
92. Albright F, Butler AM, Bloomberg E. Rickets resistant to vitamin D therapy. Am J Dis Child 1937;54:529.
93. Harrison HE. The varieties of rickets and osteomalacia associated with hypophosphatemia. Clin Orthop 1957;9:61.
94. McWhorter AG, Seale NS. Prevalence of dental abscess in a population of children with vitamin D-resistant rickets. Pediatr Dent 1991;13:91–96.
95. Berndt M, Ehrich JH, Lazovic D, et al. Clinical course of hypophosphatemic rickets in 23 adults. Clin Nephrol 1996;45:33–41.
96. Chung WT, Niu DM, Lin CY. Clinical aspects of X-linked hypophosphatemic rickets. Acta Paediatr Taiwan 2002;43:26–34.
97. Stickler GB, Morgenstern BZ. Hypophosphataemic rickets: final height and clinical symptoms in adults. Lancet 1989;2:902–905.
98. Carpenter TO. New perspectives on the biology and treatment of X-linked hypophosphatemic rickets. Pediatr Clin North Am 1997;44:443–466.
99. Carpenter
TO, Keller M, Schwartz D, et al. 24,25 dihydroxyvitamin D
supplementation corrects hyperparathyroidism and improves skeletal
abnormalities in X-linked hypophosphatemic rickets—a clinical research
center study. J Clin Endocrinol Metab 1996;81:2381–2388.
100. Ferris B, Walker C, Jackson A, et al. The orthopaedic management of hypophosphataemic rickets. J Pediatr Orthop 1991;11:367.
101. Rohmiller
MT, Tylkowski C, Kriss VM, et al. The effect of osteotomy on bowing and
height in children with X-linked hypophosphatemia. J Pediatr Orthop 1999;19:114–118.
102. Loeffler RD, Sherman FC Jr. The effect of treatment on growth and deformity in hypophosphatemic vitamin D-resistant rickets. Clin Orthop 1982;162:4–10.
103. Ringel
MD, Schwindinger WF, Levine MA. Clinical implications of genetic
defects in G proteins: the molecular basis of McCune-Albright syndrome
and Albright hereditary osteodystrophy. Medicine 1996;75:171.
104. Prader
A, Illig R, Heierli E. Eine besondere form des primare vitamin D
resistenten rachitis mit hypocalcemie und autosomal-dominanten erbgang:
die hereditare pseudomangelrachitis. Helv Paediatr Acta 1961;16:452–468.
105. Portale
AA, Miller WL. Rickets due to hereditary abnormalities of Vitamin D
synthesis or action. In: Glorieux F, Pettifor JM, Juppner H, eds. Pediatric bone: biology and diseases. San Diego, CA: Academic Press, 2003:583.
106. De Braekeleer M, Larochelle J. Population genetics of vitamin D-dependent rickets in northeastern Quebec. Ann Hum Genet 1991;55(Pt 4):283–290.
107. Fu
GK, Lin D, Zhang MY, et al. Cloning of human 25-hydroxyvitamin D-1
alpha-hydroxylase and mutations causing vitamin D-dependent rickets
type 1. Mol Endocrinol 1997;11:1961–1970.
108. Marx SJ, Spiegel AM, Brown EM, et al. A familial syndrome of decrease in sensitivity to 1,25-dihydroxyvitamin D. J Clin Endocrinol Metab 1978;47:1303–1310.
109. Brooks
MH, Bell NH, Love L, et al. Vitamin D-dependent rickets, type II:
resistance of target organs to 1,25-dihydroxy-vitamin D. N Engl J Med 1978;298:996.
110. Chesney RW, Jones DP. Rickets due to renal tubular abnormalities. In: Glorieux F, Pettifor JM, Juppner H, eds. Pediatric bone: biology and diseases. San Diego, CA: Academic Press, 2003:633.
111. Cole DEC. Hypophosphatasia. In: Glorieux F, Pettifor JM, Juppner H, eds. Pediatric bone: biology and diseases, San Diego, CA: Academic Press, 2003:651.
112. Zurutuza L, Muller F, Gibrat JF, et al. Correlations of genotype and phenotype in hypophosphatasia. Hum Mol Genet 1999; 8:1039–1046.
113. Sato S, Matsuo N. Genetic analysis of hypophosphatasia. Acta Paediatr Jpn, Overseas edition, 1997;39:528–532.
114. McCane RA, Fairweathr DVI, Barrett AM, et al. Genetic, clinical, biochemical, and pathological features of hypophosphatasia. Q J Med 1956;25:523.
115. Coe JD, Murphy WA, Whyte MP. Management of femoral fractures and pseudofractures in adult hypophosphatasia. J Bone Joint Surg Am 1986;68:981.
116. Anderton JM. Orthopaedic problems in adult hypophosphatasia: a report of two cases. J Bone Joint Surg Br 1979;61:82.
117. Hu CC, King DL, Thomas HF, et al. A clinical and research protocol for characterizing patients with hypophosphatasia. Pediatr Dent 1996;18:17–23.
118. Fraser D. Hypophosphatasia. Am J Med 1957;22:730.
119. Kuizon
BD, Salusky IB. Renal osteodystrophy: pathogenesis, diagnosis, and
treatment. In: Glorieux F, Pettifor JM, Juppner H, eds. Pediatric bone: biology and diseases. San Diego, CA: Academic Press, 2003:679.
120. Oppenheim WL, Salusky IB, Kaplan D, et al. Renal osteodystrophy in children. In: Castells S, Finberg L, eds. Metabolic bone disease in children. New York: Marcel Dekker Inc, 1990:197.
121. Weller M, Edeiken J, Hodes PJ. Renal osteodystrophy. AJR Am J Roentgenol 1968;104:354.
122. Parfitt AM. Renal osteodystrophy. Orthop Clin North Am 1972; 3:681.
123. States LJ. Imaging of metabolic bone disease and marrow disorders in children. Radiol Clin North Am 2001;39:749–772.
124. Mehls O, Ritz E, Krempien B, et al. Slipped epiphyses in renal osteodystrophy. Arch Dis Child 1975;50:545.
125. Floman
Y, Yosipovitch Z, Licht A, et al. Bilateral slipped upper femoral
epiphysis: a rare manifestation of renal osteodystrophy. Case report
with discussion of its pathogenesis. Isr J Med Sci 1975;11:15.
126. Goldman AB, Lane JM, Salvati E. Slipped capital femoral epiphyses complicating renal osteodystrophy: a report of three cases. Radiology 1978;126:33.
127. Oppenheim WL, Bowen RE, McDonough PW, et al. Outcome of slipped capital femoral epiphysis in renal osteodystrophy. J Pediatr Orthop 2003;23:169–174.
128. Loder RT, Hensinger RN. Slipped capital femoral epiphysis associated with renal failure osteodystrophy. J Pediatr Orthop 1997; 17:205–211.
129. National Institutes of Health. Osteoporosis prevention, diagnosis, and therapy. NIH Consens Statement 2000;17:1–45.
130. Calvo MS. Dietary considerations to prevent loss of bone and renal function. Nutrition 2000;16:564–566.
131. Wyshak G. Teenaged girls, carbonated beverage consumption, and bone fractures. Arch Pediatr Adolesc Med 2000;154:610–613.
132. Leonard MB, Zemel BS. Current concepts in pediatric bone disease. Pediatr Clin North Am 2002;49:143–173.
133. Ward LM, Glorieux HF. The spectrum of pediatric osteoporosis. In: Glorieux F, Pettifor JM, Juppner H, eds. Pediatric bone: biology and diseases. San Diego, CA: Academic Press, 2003:401.

P.200


134. Orioli IM, Castilla EE, Barbosa-Neto JG. The birth prevalence rates for the skeletal dysplasias. J Med Genet 1986;23:328–332.
135. Seedorff KS. Osteogenesis imperfecta: a study of clinical features and heredity based on 55 Danish families comprising 180 affected persons. Copenhagen: Ejnar Munksgaard, 1949.
136. Smith
R, Francis MJO, Bauze RJ. Osteogenesis imperfecta: a clinical and
biochemical study of a generalized connective tissue disorder. Q J Med 1975;44:555.
137. Wynne-Davis R, Gormley J. Clinical and genetic patterns in osteogenesis imperfecta. Clin Orthop 1981;159:26.
138. Cole WG. The Nicholas Andry Award—1996. The molecular pathology of osteogenesis imperfecta. Clin Orthop Relat Res 1997;343:235–248.
139. Sillence DO. Osteogenesis imperfecta: an expanding panorama of variance. Clin Orthop 1981;159:11.
140. Sillence DO, Senn A, Danks DM. Genetic heterogeneity in osteogenesis imperfecta. J Med Genet 1979;16:101.
141. Cole WG. Etiology and pathogenesis of heritable connective tissue diseases. J Pediatr Orthop 1993;13:392.
142. Kuurila K, Grenman R, Johansson R, et al. Hearing loss in children with osteogenesis imperfecta. Eur J Pediatr 2000;159:515–519.
143. Imani
P, Vijayasekaran S, Lannigan F. Is it necessary to screen for hearing
loss in the paediatric population with osteogenesis imperfecta? Clin Otolaryngol 2003;28:199–202.
144. Kuurila K, Kaitila I, Johansson R, et al. Hearing loss in Finnish adults with osteogenesis imperfecta: a nationwide survey. Ann Otol Rhinol Laryngol 2002;111:939–946.
145. van der Rijt AJ, Cremers CW. Stapes surgery in osteogenesis imperfecta: results of a new series. Otol Neurotol 2003;24:717–722.
146. Szilvassy J, Jori J, Czigner J, et al. Cochlear implantation in osteogenesis imperfecta. Acta Otorhinolaryngol 1998;52:253–256.
147. Stott NS, Zionts LE. Displaced fractures of the apophysis of the olecranon in children who have osteogenesis imperfecta. J Bone Joint Surg Am 1993;75:1026–1033.
148. Zeitlin L, Fassier F, Glorieux FH. Modern approach to children with osteogenesis imperfecta. J Pediatr Orthop B 2003;12: 77–87.
149. Plotkin H, Primorac D, Rowe D. Osteogenesis imperfecta. In: Glorieux F, Pettifor JM, Juppner H, eds. Pediatric bone: biology and diseases. San Diego, CA: Academic Press, 2003:443.
150. Glorieux FH, Rauch F, Plotkin H, et al. Type V osteogenesis imperfecta: a new form of brittle bone disease. J Bone Miner Res 2000;15:1650–1658.
151. Glorieux
FH, Ward LM, Rauch F, et al. Osteogenesis imperfecta type VI: a form of
brittle bone disease with a mineralization defect. J Bone Miner Res 2002;17:30–38.
152. Ward LM, Rauch F, Travers R, et al. Osteogenesis imperfecta type VII: an autosomal recessive form of brittle bone disease. Bone 2002;31:12–18.
153. Pettinen RP, Lichtenstein JR, Martin GR, et al. Abnormal collagen metabolism in cultured cells in osteogenesis imperfecta. Proc Natl Acad Sci U S A 1975;72:586.
154. Barsh
GS, David KE, Byers PH. Type I osteogenesis imperfecta: a nonfunctional
allele for pro alpha (I) chains for type I procollagen. Proc Natl Acad Sci U S A 1982;79:3838.
155. Peltonen
L, Palotie A, Prockop DJ. A defect in the structure of type I
procollagen in a patient who had osteogenesis imperfecta: excessive
mannose in the COOH-terminal peptide. Proc Natl Acad Sci U S A 1980;77:6179.
156. Alman B, Frasca P. Fracture failure mechanisms in patients with osteogenesis imperfecta. J Orthop Res 1987;5:139–143.
157. Moore
MS, Minch CM, Kruse RW, et al. The role of dual energy x-ray
absorptiometry in aiding the diagnosis of pediatric osteogenesis
imperfecta. Am J Orthop 1998;27:797–801.
158. Zionts LE, Nash JP, Rude R, et al. Bone mineral density in children with mild osteogenesis imperfecta. J Bone Joint Surg Br 1995;77:143–147.
159. Faulkner RA, Bailey DA, Drinkwater DT, et al. A. Bone densitometry in Canadian children 8–17 years of age. Calcif Tissue Int 1996;59:344–351.
160. Gafni
RI, Baron J. Overdiagnosis of osteoporosis in children due to
misinterpretation of dual-energy x-ray absorptiometry (DEXA). J Pediatr 2004;144:253–257.
161. Minch CM, Kruse RW. Osteogenesis imperfecta: a review of basic science and diagnosis. Orthopedics 1998;21:558.
162. Marlowe A, Pepin MG, Byers PH. Testing for osteogenesis imperfecta in cases of suspected non-accidental injury. J Med Genet 2002;39:382–386.
163. Glorieux FH, Bishop NJ, Plotkin H, et al. Cyclic administration of pamidronate in children with severe osteogenesis imperfecta. N Engl J Med 1998;339:947–952.
164. Grissom LE, Harcke HT. Radiographic features of bisphosphonate therapy in pediatric patients. Pediatr Radiol 2003;33:226–229.
165. Whyte MP, Wenkert D, Clements KL, et al. Bisphosphonate-induced osteopetrosis. N Engl J Med 2003;349:457–463.
166. Evans
KD, Lau ST, Oberbauer AM, et al. Alendronate affects long bone length
and growth plate morphology in the OIM mouse model for osteogenesis
imperfecta. Bone 2003;32:268–274.
167. Lepola VT, Hannuniemi R, Kippo K, et al. Long-term effects of clodronate on growing rat bone. Bone 1996;18:191–196.
168. Miller
SC, Jee WS, Woodbury DM, et al. Effects of N, N, N’,
N’-ethylenediaminetetramethylene phosphonic acid and
1-hydroxyethylidene-1,1-bisphosphonic acid on calcium absorption,
plasma calcium, longitudinal bone growth, and bone histology in the
growing rat. Toxicol Appl Pharmacol 1985;77:230–239.
169. Letts
M, Monson R, Weber K. The prevention of recurrent fractures of the
lower extremities in severe osteogenesis imperfecta using vacuum pants:
a preliminary report in four patients. J Pediatr Orthop 1988;8:454.
170. Gargan MF, Wisbeach A, Fixsen JA. Humeral rodding in osteogenesis imperfecta. J Pediatr Orthop 1996;16:719–722.
171. Sofield
HA, Millar EA. Fragmentation, realignment, and intramedullary rod
fixation of deformities of the long bones in children: a ten-year
appraisal. J Bone Joint Surg Am 1959;41:1371.
172. Williams PF. Fragmentation and rodding in osteogenesis imperfecta. J Bone Joint Surg Br 1965;47:23.
173. Nicholas RW, James P. Telescoping intramedullary stabilization of the lower extremities for severe osteogenesis imperfecta. J Pediatr Orthop 1990;10:219.
174. Fassier F, Glorieux FH. Surgical management of osteogenesis imperfecta. In: Duparc. J, ed. Surgical techniques in orthopaedics and traumatology. New York: Elsevier Science, 2003.
175. Gamble
JG, Strudwick WJ, Rinsky LA, et al. Complications of intramedullary
rods in osteogenesis imperfecta: Bailey-Dubow rods versus nonelongating
rods. J Pediatr Orthop 1988;8:645.
176. Porat
S, Heller E, Seidman DS, et al. Functional results of operation in
osteogenesis imperfecta: elongating and nonelongating rods. J Pediatr Orthop 1991;11:200.
177. Ryoppy S, Alberty A, Kaitila I. Early semiclosed intramedullary stabilization in osteogenesis imperfecta. J Pediatr Orthop 1987; 7:139.
178. Middleton RWD, Frost RB. Percutaneous intramedullary rod interchange in osteogenesis imperfecta. J Bone Joint Surg Br 1987;69:429.
179. Tiley F, Albright JA. Osteogenesis imperfecta: treatment by multiple osteotomy and intramedullary rod insertion. J Bone Joint Surg Am 1973;55:701.
180. Luhmann
SJ, Sheridan JJ, Capelli AM, et al. Management of lower-extremity
deformities in osteogenesis imperfecta with extensible intramedullary
rod technique: a 20-year experience. J Pediatr Orthop 1998;18:88–94.
181. Engelbert RH, Pruijs HE, Beemer FA, et al. Osteogenesis imperfecta in childhood: treatment strategies. Arch Phys Med Rehabil 1998;79:1590–1594.
182. Antoniazzi F, Mottes M, Fraschini P, et al. Osteogenesis imperfecta: practical treatment guidelines. Paediatric Drugs 2000;2: 465–488.
183. Mulpuri K, Joseph B. Intramedullary rodding in osteogenesis imperfecta. J Pediatr Orthop 2000;20:267–273.
184. Khoshhal KI, Ellis RD. Effect of lower limb Sofield procedure on ambulation in osteogenesis imperfecta. J Pediatr Orthop 2001; 21:233–235.
185. Engelbert RH, Uiterwaal CS, Gulmans VA, et al. Osteogenesis imperfecta in childhood: prognosis for walking. J Pediatr 2000; 137:397–402.
186. Engelbert
RH, Uiterwaal CS, Gulmans VA, et al. Osteogenesis imperfecta: profiles
of motor development as assessed by a postal questionnaire. Eur J Pediatr 2000;159:615–620.
187. Engelbert
RH, Beemer FA, van der Graaf Y, et al. Osteogenesis imperfecta in
childhood: impairment and disability—a follow-up study. Arch Phys Med Rehabil 1999;80:896–903.

P.201


188. Engelbert
RH, Uiterwaal CS, van der Hulst A, et al. Scoliosis in children with
osteogenesis imperfecta: influence of severity of disease and age of
reaching motor milestones. Eur Spine J 2003;12:130–134.
189. Janus
GJ, Finidori G, Engelbert RH, et al. Operative treatment of severe
scoliosis in osteogenesis imperfecta: results of 20 patients after halo
traction and posterior spondylodesis with instrumentation. Eur Spine J 2000;9:486–491.
190. Widmann
RF, Bitan FD, Laplaza FJ, et al. Spinal deformity, pulmonary
compromise, and quality of life in osteogenesis imperfecta. Spine 1999;24:1673–1678.
191. Engelbert
RH, Gerver WJ, Breslau-Siderius LJ, et al. Spinal complications in
osteogenesis imperfecta: 47 patients 1–16 years of age. Acta Orthop Scand 1998;69:283–286.
192. Hanscom
DA, Winter RB, Lutter L, et al. Osteogenesis imperfecta. Radiographic
classification, natural history, and treatment of spinal deformities. J Bone Joint Surg Am 1992;74:598–616.
193. Hanscom DA, Bloom BA. The spine in osteogenesis imperfecta. Orthop Clin North Am 1988;19:449–458.
194. Yong-Hing K, MacEwen GD. Scoliosis associated with osteogenesis imperfecta. J Bone Joint Surg Br 1982;64:36–43.
195. Benson DR, Newman DC. The spine and surgical treatment in osteogenesis imperfecta. Clin Orthop Relat Res 1981;159:147–153.
196. Cristofaro RL, Hoek KJ, Bonnett CA, et al. Operative treatment of spine deformity in osteogenesis imperfecta. Clin Orthop Relat Res 1979;139:40–48.
197. Benson DR, Donaldson DH, Millar EA. The spine in osteogenesis imperfecta. J Bone Joint Surg Am 1978;60:925–929.
198. Jones ET, Hensinger RN. Spinal deformity in idiopathic juvenile osteoporosis. Spine 1981;6:1.
199. Breslau-Siderius
EJ, Engelbert RH, Pals G, et al. Bruck syndrome: a rare combination of
bone fragility and multiple congenital joint contractures. J Pediatr Orthop B 1998;7:35–38.
200. Bank
RA, Robins SP, Wijmenga C, et al. Defective collagen crosslinking in
bone, but not in ligament or cartilage, in Bruck syndrome: indications
for a bone-specific telopeptide lysyl hydroxylase on chromosome 17. Proc Natl Acad Sci U S A 1999;96:1054–1058.
201. Frontali M, Stomeo C, Dallapiccola B. Osteoporosis-pseudoglioma syndrome: report of three affected sibs and an overview. Am J Med Genet 1985;22:35–47.
202. McDowell CL, Moore JD. Multiple fractures in a child: the osteoporosis pseudoglioma syndrome. A case report. J Bone Joint Surg Am 1992;74:1247–1249.
203. Gong Y, Slee RB, Fukai N, et al. LDL receptor-related protein 5 (LRP5) affects bone accrual and eye development. Cell 2001; 107:513–523.
204. Zacharin M, Cundy T. Osteoporosis pseudoglioma syndrome: treatment of spinal osteoporosis with intravenous bisphosphonates. J Pediatr 2000;137:410–415.
205. Kasten P, Bastian L, Schmid H, et al. Failure of operative treatment in a child with osteoporosis-pseudoglioma syndrome. Clin Orthop Relat Res 2003;410:262–266.
206. Ducy P, Schinke T, Karsenty G. The osteoblast: a sophisticated fibroblast under central surveillance. Science 2000;289: 1501–1504.
207. Ducy P, Amling M, Takeda S, et al. Leptin inhibits bone formation through a hypothalamic relay: a central control of bone mass. Cell 2000;100:197–207.
208. Takeda S, Elefteriou F, Levasseur R, et al. Leptin regulates bone formation via the sympathetic nervous system. Cell 2002;111: 305–317.
209. Flier JS. Physiology: is brain sympathetic to bone? Nature 2002;420:619, 621–622.
210. King W, Levin R, Schmidt R, et al. Prevalence of reduced bone mass in children and adults with spastic quadriplegia. Dev Med Child Neurol 2003;45:12–16.
211. Henderson
RC, Lark RK, Gurka MJ, et al. Bone density and metabolism in children
and adolescents with moderate to severe cerebral palsy. Pediatrics 2002;110(1 Pt 1):E5.
212. Ihkkan
DY, Yalcin E. Changes in skeletal maturation and mineralization in
children with cerebral palsy and evaluation of related factors. J Child Neurol 2001;16:425–430.
213. Tasdemir HA, Buyukavci M, Akcay F, et al. Bone mineral density in children with cerebral palsy. Pediatr Int 2001;43:157–160.
214. Henderson RC, Lin PP, Greene WB. Bone-mineral density in children and adolescents who have spastic cerebral palsy. J Bone Joint Surg Am 1995;77:1671–1681.
215. Lin PP, Henderson RC. Bone mineralization in the affected extremities of children with spastic hemiplegia. Dev Med Child Neurol 1996;38:782–786.
216. Naftchi NE, Viau AT, Marshall CH, et al. Bone mineralization in the distal forearm of hemiplegic patients. Arch Phys Med Rehabil 1975;56:487–492.
217. Duncan
B, Barton LL, Lloyd J, et al. Dietary considerations in osteopenia in
tube-fed nonambulatory children with cerebral palsy. Clin Pediatr 1999;38:133–137.
218. Brunner R, Doderlein L. Pathological fractures in patients with cerebral palsy. J Pediatr Orthop B 1996;5:232–238.
219. Bischof
F, Basu D, Pettifor JM. Pathological long-bone fractures in residents
with cerebral palsy in a long-term care facility in South Africa. Dev Med Child Neurol 2002;44:119–122.
220. Chad
KE, Bailey DA, McKay HA, et al. The effect of a weight-bearing physical
activity program on bone mineral content and estimated volumetric
density in children with spastic cerebral palsy. J Pediatr 1999;135:115–117.
221. Jekovec-Vrhovsek
M, Kocijancic A, Prezelj J. Effect of vitamin D and calcium on bone
mineral density in children with CP and epilepsy in full-time care. Dev Med Child Neurol 2000;42:403–405.
222. Henderson
RC, Lark RK, Kecskemethy HH, et al. Bisphosphonates to treat osteopenia
in children with quadriplegic cerebral palsy: a randomized,
placebo-controlled clinical trial. J Pediatr 2002; 141:644–651.
223. Larson CM, Henderson RC. Bone mineral density and fractures in boys with Duchenne muscular dystrophy. J Pediatr Orthop 2000;20:71–74.
224. Bianchi ML, Mazzanti A, Galbiati E, et al. Bone mineral density and bone metabolism in Duchenne muscular dystrophy. Osteoporos Int 2003;14:761–767.
225. Markham A, Bryson HM. Deflazacort: a review of its pharmacological properties and therapeutic efficacy. Drugs 1995;50:317–333.
226. Loftus
J, Allen R, Hesp R, et al. Randomized, double-blind trial of
deflazacort versus prednisone in juvenile chronic (or rheumatoid)
arthritis: a relatively bone-sparing effect of deflazacort. Pediatrics 1991;88:428–436.
227. Talim B, Malaguti C, Gnudi S, et al. Vertebral compression in Duchenne muscular dystrophy following deflazacort. Neuromuscul Disord 2002;12:294–295.
228. Wuster C. Fracture rates in patients with growth hormone deficiency. Horm Res 2000;54(Suppl 1):31–35.
229. Cowell CT, Wuster C. The effects of growth hormone deficiency and growth hormone replacement therapy on bone. A meeting report. Horm Res 2000;54(Suppl 1):68–74.
230. Clanget
C, Seck T, Hinke V, et al. Effects of 6 years of growth hormone (GH)
treatment on bone mineral density in GH-deficient adults. Clin Endocrinol 2001;55:93–99.
231. Loder RT, Wittenberg B, DeSilva G. Slipped capital femoral epiphysis associated with endocrine disorders. J Pediatr Orthop 1995;15:349–356.
232. Cowell CT, Dietsch S. Adverse events during growth hormone therapy. J Pediatr Endocrinol Metab 1995;8:243–252.
233. Bailey
DA, Wedge JH, McCulloch RG, et al. Epidemiology of fractures of the
distal end of the radius in children as associated with growth. J Bone Joint Surg Am 1989;71:1225–1231.
234. Khosla
S, Melton LJ III, Dekutoski MB, et al. Incidence of childhood distal
forearm fractures over 30 years: a population-based study. JAMA 2003;290:1479–1485.
235. Beals KA, Manore MM. Disorders of the female athlete triad among collegiate athletes. Int J Sport Nutr Exerc Metab 2002; 12:281–293.
236. Kazis K, Iglesias E. The female athlete triad. Adolesc Med 2003; 14:87–95.
237. Skolnick AA. ‘Female athlete triad’ risk for women. JAMA 1993;270:921–923.
238. Khan
KM, Liu-Ambrose T, Sran MM, et al. New criteria for female athlete
triad syndrome? As osteoporosis is rare, should osteopenia be among the
criteria for defining the female athlete triad syndrome? Br J Sports Med 2002;36:10–13.

P.202


239. Lauder TD, Williams MV, Campbell CS, et al. The female athlete triad: prevalence in military women. Mil Med 1999;164:630–635.
240. Robinson
TL, Snow-Harter C, Taaffe DR, et al. Gymnasts exhibit higher bone mass
than runners despite similar prevalence of amenorrhea and
oligomenorrhea. J Bone Miner Res 1995;10:26–35.
241. Bale P, Doust J, Dawson D. Gymnasts, distance runners, anorexics: body composition and menstrual status. J Sports Med Phys Fitness 1996;36:49–53.
242. Snow CM, Rosen CJ, Robinson TL. Serum IGF-I is higher in gymnasts than runners and predicts bone and lean mass. Med Sci Sports Exerc 2000;32:1902–1907.
243. Fuchs
RK, Bauer JJ, Snow CM. Jumping improves hip and lumbar spine bone mass
in prepubescent children: a randomized controlled trial. J Bone Miner Res 2001;16:148–156.
244. McKay
HA, Petit MA, Schutz RW, et al. Augmented trochanteric bone mineral
density after modified physical education classes: a randomized
school-based exercise intervention study in prepubescent and early
pubescent children. J Pediatr 2000;136: 156–162.
245. Maugars YM, Berthelot JM, Forestier R, et al. Follow-up of bone mineral density in 27 cases of anorexia nervosa. Eur J Endocrinol 1996;135:591–597.
246. Lucas
AR, Melton LJ III, Crowson CS, et al. Long-term fracture risk among
women with anorexia nervosa: a population-based cohort study. Mayo Clin Proc 1999;74:972–977.
247. Seeman
E, Szmukler GI, Formica C, et al. Osteoporosis in anorexia nervosa: the
influence of peak bone density, bone loss, oral contraceptive use, and
exercise. J Bone Miner Res 1992;7: 1467–1474.
248. Golden
NH, Lanzkowsky L, Schebendach J, et al. The effect of
estrogen-progestin treatment on bone mineral density in anorexia
nervosa. J Pediatr Adolesc Gynecol 2002;15:135–143.
249. Csermely T, Halvax L, Schmidt E, et al. Occurrence of osteopenia among adolescent girls with oligo/amenorrhea. Gynecol Endocrinol 2002;16:99–105.
250. Ducharme
FM, Chabot G, Polychronakos C, et al. Safety profile of frequent short
courses of oral glucocorticoids in acute pediatric asthma: impact on
bone metabolism, bone density, and adrenal function. Pediatrics 2003;111:376–383.
251. Adachi
JD, Roux C, Pitt PI, et al. A pooled data analysis on the use of
intermittent cyclical etidronate therapy for the prevention and
treatment of corticosteroid induced bone loss. J Rheumatol 2000;27:2424–2431.
252. Adachi JD, Bensen WG, Brown J, et al. Intermittent etidronate therapy to prevent corticosteroid-induced osteoporosis. N Engl J Med 1997;337:382–387.
253. Reid
DM, Hughes RA, Laan RF, et al. Efficacy and safety of daily risedronate
in the treatment of corticosteroid-induced osteoporosis in men and
women: a randomized trial. European Corticosteroid-induced Osteoporosis
Treatment Study. J Bone Miner Res 2000;15:1006–1013.
254. Wallach
S, Cohen S, Reid DM, et al. Effects of risedronate treatment on bone
density and vertebral fracture in patients on corticosteroid therapy. Calcif Tissue Int 2000;67:277–285.
255. Noguera A, Ros JB, Pavia C, et al. Bisphosphonates, a new treatment for glucocorticoid-induced osteoporosis in children. J Pediatr Endocrinol Metab 2003;16:529–536.
256. Lane
NE, Sanchez S, Modin GW, et al. Bone mass continues to increase at the
hip after parathyroid hormone treatment is discontinued in
glucocorticoid-induced osteoporosis: results of a randomized controlled
clinical trial. J Bone Miner Res 2000; 15:944–951.
257. Lane
NE, Sanchez S, Modin GW, et al. Parathyroid hormone treatment can
reverse corticosteroid-induced osteoporosis. Results of a randomized
controlled clinical trial. J Clin Invest 1998;102:1627–1633.
258. Dent CE, Richens A, Rowe DJF, et al. Osteomalacia with long-term anticonvulsant therapy in epilepsy. Br Med J 1970;4:69.
259. Frame B. Hypocalcemia and osteomalacia associated with anticonvulsant therapy. Ann Intern Med 1971;74:294.
260. Aponte CJ, Petrelli MP. Anticonvulsants and vitamin D metabolism. JAMA 1973;225:1248.
261. Crosley CJ, Chee C, Berman PH. Rickets associated with long-term anticonvulsant therapy in a pediatric outpatient population. Pediatrics 1975;56:52.
262. Winnacker
JL, Yeager H, Saunders RB, et al. Rickets in children receiving
anticonvulsant drugs: biochemical and hormonal markers. Am J Dis Child 1977;131:286.
263. Morijiri Y, Sato T. Factors causing rickets in institutionalised handicapped children on anticonvulsant therapy. Arch Dis Child 1981;56:446–449.
264. Timperlake
RW, Cook SD, Thomas KA, et al. Effects of anticonvulsant drug therapy
on bone mineral density in a pediatric population. J Pediatr Orthop 1988;8:467.
265. Hahn
TJ, Halstead LR. Anticonvulsant drug-induced osteomalacia: alterations
in mineral metabolism and response to vitamin D3 administration. Calcif Tissue Int 1979;27:13–18.
266. Hahn TJ. Bone complications of anticonvulsants. Drugs 1976; 12:201–211.
267. Henderson RC. Vitamin D levels in noninstitutionalized children with cerebral palsy. J Child Neurol 1997;12:443–447.
268. Camfield CS, Delvin EE, Camfield PR, et al. Normal serum 25-hydroxyvitamin D levels in phenobarbital-treated toddlers. Dev Pharmacol Ther 1983;6:157–161.
269. Feldkamp
J, Becker A, Witte OW, et al. Long-term anticonvulsant therapy leads to
low bone mineral density—evidence for direct drug effects of phenytoin
and carbamazepine on human osteoblast-like cells. Exp Clin Endocrinol Diabetes 2000;108:37–43.
270. Rieger-Wettengl
G, Tutlewski B, Stabrey A, et al. Analysis of the musculoskeletal
system in children and adolescents receiving anticonvulsant monotherapy
with valproic acid or carbamazepine. Pediatrics 2001;108:E107.
Sclerosing Bone Conditions in Children
271. Shapiro F. Osteopetrosis. Current clinical considerations. Clin Orthop Relat Res 1993;294:34–44.
272. Key LLJ, Ries WL. Sclerosing bony dysplasia. In: Glorieux F, Pettifor JM, Juppner H, eds. Pediatric bone: biology and diseases. San Diego, CA: Academic Press, 2003:473.
273. Milgram JW, Jasty M. Osteopetrosis. J Bone Joint Surg Am 1982;64:912.
274. Hinkel CL, Beiler DD. Osteopetrosis in adults. AJR 1955;74:46.
275. Whyte MP. Carbonic anhydrase II deficiency. Clin Orthop Relat Res 1993;294:52–63.
276. Sly
WS, Whyte MP, Sundaram V, et al. Carbonic anhydrase II deficiency in 12
families with the autosomal recessive syndrome of osteopetrosis with
renal tubular acidosis and cerebral calcification. N Engl J Med 1985;313:139–145.
277. Sly
WS, Hewett-Emmett D, Whyte MP, et al. Carbonic anhydrase II deficiency
identified as the primary defect in the autosomal recessive syndrome of
osteopetrosis with renal tubular acidosis and cerebral calcification. Proc Natl Acad Sci U S A 1983;80: 2752–2756.
278. Cleiren
E, Benichou O, Van Hul E, et al. Albers-Schonberg disease (autosomal
dominant osteopetrosis, type II) results from mutations in the ClCN7
chloride channel gene. Hum Mol Genet 2001;10:2861–2867.
279. Ballet JJ, Griscelli C, Coutris C, et al. Bone marrow transplantation in osteopetrosis. Lancet 1977;2:1137.
280. Coccia BF, Krivit W, Cervenka J. Successful bone marrow transplantation for infantile osteopetrosis. N Engl J Med 1980;302:701.
281. Sieff CA, Levinsky RJ, Rogers DW, et al. Allogeneic bone-marrow transplantation in infantile malignant osteopetrosis. Lancet 1983;1:437.
282. Kaplan
FS, August CS, Fallon MD, et al. Successful treatment of infantile
malignant osteopetrosis by bone-marrow transplantation. J Bone Joint Surg Am 1988;70:617.
283. Key LLJ, Rodgriguiz RM, Willi SM, et al. Long-term treatment of osteopetrosis with recombinant human interferon gamma. N Engl J Med 1995;332:1594.
284. Armstrong
DG, Newfield JT, Gillespie R. Orthopedic management of osteopetrosis:
results of a survey and review of the literature. J Pediatr Orthop 1999;19:122–132.
285. Szappanos L, Szepesi K, Thomazy V. Spondylolysis in osteopetrosis. J Bone Joint Surg Br 1988;70:428–430.
286. de Jong G, Muller LM. Perinatal death in two sibs with infantile cortical hyperostosis (Caffey disease). Am J Med Genet 1995;59: 134–138.
287. Pazzaglia UE, Byers P, Beluffi G, et al. Pathology of infantile cortical hyperostosis (Caffey’s disease). J Bone Joint Surg Am 1985;67:1417.

P.203


288. Thometz JG, DiRaimondo CA. A case of recurrent Caffey’s disease treated with naproxen. Clin Orthop Relat Res 1996;324:304–309.
289. Couper RT, McPhee A, Morris L. Indomethacin treatment of infantile cortical periostosis in twins. J Paediatr Child Health 2001;37:305–308.
290. Ueda
K, Saito A, Nakano H, et al. Cortical hyperostosis following long-term
administration of prostaglandin E in infants with cyanotic congenital
heart disease. J Pediatr 1980;97:834.
291. Everts
V, Hou WS, Rialland X, et al. Cathepsin K deficiency in pycnodysostosis
results in accumulation of non-digested phagocytosed collagen in
fibroblasts. Calcif Tissue Int 2003;73:380–386.
292. Goto T, Yamaza T, Tanaka T. Cathepsins in the osteoclast. J Electron Microsc 2003;52:551–558.
293. Motyckova G, Fisher DE. Pycnodysostosis: role and regulation of cathepsin K in osteoclast function and human disease. Curr Mol Med 2002;2:407–421.
294. Fujita
Y, Nakata K, Yasui N, et al. Novel mutations of the cathepsin K gene in
patients with pycnodysostosis and their characterization. J Clin Endocrinol Metab 2000;85:425–431.
295. Edelson JG, Obad S, Geiger R, et al. Pycnodysostosis. Orthopedic aspects with a description of 14 new cases. Clin Orthop Relat Res 1992;280:263–276.
296. Vanhoenacker FM, De Beuckeleer LH, Van Hul W, et al. Sclerosing bone dysplasias: genetic and radioclinical features. Eur Radiol 2000;10:1423–1433.
297. Soliman
AT, Ramadan MA, Sherif A, et al. Pycnodysostosis: clinical, radiologic,
and endocrine evaluation and linear growth after growth hormone
therapy. Metab Clin Exp 2001;50:905–911.
298. Kim
JE, Kim EH, Han EH, et al. A TGF-beta-inducible cell adhesion molecule,
betaig-h3, is downregulated in melorheostosis and involved in
osteogenesis. J Cell Biochem 2000;77:169–178.
299. Fryns JP. Melorheostosis and somatic mosaicism. Am J Med Genet 1995;58:199.
300. Younge D, Drummond D, Herring J, et al. Melorheostosis in children. Clinical features and natural history. J Bone Joint Surg Br 1979;61-B:415–418.
301. Murphy M, Kearns S, Cavanagh M, et al. Occurrence of osteosarcoma in a melorheostotic femur. Ir Med J 2003;96:55–56.
302. Brennan DD, Bruzzi JF, Thakore H, et al. Osteosarcoma arising in a femur with melorheostosis and osteopathia striata. Skeletal Radiol 2002;31:471–474.
303. Bostman OM, Holmstrom T, Riska EB. Osteosarcoma arising in a melorheostotic femur. A case report. J Bone Joint Surg Am 1987;69:1232–1237.
304. Choi IH, Kim JI, Yoo WJ, et al. Ilizarov treatment for equinoplanovalgus foot deformity caused by melorheostosis. Clin Orthop Relat Res 2003;414:238–241.
305. Atar
D, Lehman WB, Grant AD, et al. The Ilizarov apparatus for treatment of
melorheostosis. Case report and review of the literature. Clin Orthop Relat Res 1992;281:163–167.
306. Griffet J, el Hayek T, Giboin P. Melorheostosis: complications of a tibial lengthening with the Ilizarov apparatus. Eur J Pediatr Surg 1998;8:186–189.
307. Vanhoenacker FM, Janssens K, Van Hul W, et al. Camurati-Engelmann disease. Review of radioclinical features. Acta Radiol 2003;44:430–434.
308. Janssens
K, ten Dijke P, Ralston SH, et al. Transforming growth factor-beta 1
mutations in Camurati-Engelmann disease lead to increased signaling by
altering either activation or secretion of the mutant protein. J Biol Chem 2003;278:7718–7724.
309. Campos-Xavier B, Saraiva JM, Savarirayan R, et al. Phenotypic variability at the TGF-beta1 locus in Camurati-Engelmann disease. Hum Genet 2001;109:653–658.
310. Beighton P, Davidson J, Durr L, et al. Sclerosteosis—an autosomal recessive disorder. Clin Genet 1977;11:1–7.
311. Hamersma H, Gardner J, Beighton P. The natural history of sclerosteosis. Clin Genet 2003;63:192–197.
312. Itin PH, Keseru B, Hauser V. Syndactyly/brachyphalangy and nail dysplasias as marker lesions for sclerosteosis. Dermatology 2001;202:259–260.
313. Stephen LX, Hamersma H, Gardner J, et al. Dental and oral manifestations of sclerosteosis. Int Dent J 2001;51:287–290.
314. Cremin BJ. Sclerosteosis in children. Pediatr Radiol 1979;8: 173–177.
315. Beighton P, Cremin BJ, Hamersma H. The radiology of sclerosteosis. Br J Radiol 1976;49:934–939.
316. Beighton
P, Durr L, Hamersma H. The clinical features of sclerosteosis. A review
of the manifestations in twenty-five affected individuals. Ann Intern Med 1976;84:393–397.
317. du Plessis JJ. Sclerosteosis: neurosurgical experience with 14 cases. Br J Neurosurg 1993;78:388–392.
318. Balemans
W, Ebeling M, Patel N, et al. Increased bone density in sclerosteosis
is due to the deficiency of a novel secreted protein (SOST). Hum Mol Genet 2001;10:537–543.
319. Brunkow
ME, Gardner JC, Van Ness J, et al. Bone dysplasia sclerosteosis results
from loss of the SOST gene product, a novel cystine knot-containing
protein. Am J Hum Genet 2001; 68:577–589.

This website uses cookies to improve your experience. We'll assume you're ok with this, but you can opt-out if you wish. Accept Read More