Anda di halaman 1dari 38

C H A P T E R

58
Osteoporosis Associated with
Chronic Kidney Disease
Susan M. Ott1, Grahame Elder2
1University of Washington, Seattle, WA, USA, 2Department of Renal Medicine, Westmead Hospital; Osteoporosis and
Bone Biology Division, Garvan Institute of Medical Research, Sydney, Australia

INTRODUCTION significant reductions in the eGFR; for example early


diabetes or conditions interfering with tubular transport,
Since the kidney is essential for regulation of mineral such as Fanconi syndrome. These patients may have
metabolism, bone disease is an inevitable consequence of bone disease, due to acidosis or hypophosphatemia, but
renal damage. Treatment is challenging due to the com- that is beyond the scope of this chapter. Here we will
plex physiology and the dearth of clinical data, and is consider the management of patients with CKD stages
complicated by concern about vascular calcifications. A 3 to 5, defined as eGFR 30 to 59 mL/minute/1.73 m2
newly defined term CKD-MBD (chronic kidney disease (CKD stage 3), eGFR 15 to 29 mL/minute/1.73 m2
mineral and bone disorder) highlights the connections (CKD stage 4), and eGFR less than 15 mL/minute/1.73
between the biochemical and hormonal abnormalities of m2 (CKD stage 5) or on dialysis (5D) (Fig. 58.1). These
mineral metabolism, the vascular calcifications, and the patients often show changes of CKD-MBD.
skeletal disease [1]. A systematic review of studies worldwide found the
median prevalence of CKD in adults (using the MDRD
formula) is 7.2%, and, in those older than 64 years of age,
STAGES OF CHRONIC KIDNEY DISEASE this increases to about 30% [2]. Women had a higher rate
than men, African-Americans had a lower prevalence
The stages of chronic kidney disease (CKD) are than Caucasians, and Asian populations had a relatively
defined by the presence of a reduced glomerular filtra- high prevalence. These estimated rates of prevalence
tion rate (GFR), or when the GFR is normal, by other will vary depending on the method used to calculate
indicators of renal disease. In clinical practice the GFR is kidney function. For example, using the Cockcroft–Gault
estimated (eGFR) using a Modification of Diet in Renal equation, the prevalence of CKD is higher than when
Disease (MDRD) formula based on the serum creatinine, using the MDRD equation, especially in older popula-
the age, race, and gender of a person. CKD stages 1 and tions [2]. These methods both use serum creatinine,
2 refer to kidney diseases that are not associated with which depends on the muscle mass. Cystatin C, another

FIGURE 58.1  Stages of chronic kidney disease.

Osteoporosis. http://dx.doi.org/10.1016/B978-0-12-415853-5.00058-3 1387 Copyright © 2013 Elsevier Inc. All rights reserved.
1388 58.  OSTEOPOROSIS ASSOCIATED WITH CHRONIC KIDNEY DISEASE

marker used to determine GFR, is independent of mus- as a patient’s age, sex, and racial background. Changes
cle mass and may, therefore, be a more sensitive method of ROD will also be superimposed on, and coexist with,
for detecting CKD, particularly in elderly patients. underlying bone pathology such as postmenopausal
osteoporosis, and will be affected by drugs such as gluco-
corticoids. However, the predominant bone pathology of
patients with CKD stages 3 to 5 results from abnormali-
EPIDEMIOLOGY
ties of mineral metabolism caused by CKD. Treatments
also impact bone histopathology, particularly the use of
Renal Osteodystrophy
vitamin D receptor activators (VDRAs), calcimimetics,
The term renal osteodystrophy (ROD) encompasses the choice of phosphate binder, the mode of dialysis, and
histologic changes in bone that develop as a consequence the dialysate used. Descriptively, these histologic find-
of CKD [1]. Clearly these will vary in severity and pattern ings can be categorized by alterations in bone turnover,
with CKD duration and stage, together with factors such mineralization, and volume (TMV). These are discussed

TABLE 58.1  Prevalence and Incidence of Fractures in Dialysis Patients


Prevalence (%) Incidence (%/year)

Author Year n Patients Any/hip Spine Any Hip

Pendras 1966 19 First HD patients 47 – – –

Rubini 1969 29 HD 27 – – –

Parfitt 1972 16 HD 44 25 – –

Piraino 1988 16 HD: fibrosis – – 4.8 –

Yamaguchi 1996 124 HD 10 11 – –

Atsumi 1999 187 HD – 21 – –

Gerakis 2000 62 HD 11 – – –

Alem 2000 182,493 Males – – – 0.74

Alem 2000 143,971 Females – – – 1.36

Coco 2000 1272 HD all in unit – – – 1.39

Stehman-Breen 2000 4952 HD – – – 0.69

Ball 2002 101,039 USRDS on tx list – – – 0.29

Jamal 2002 104 HD > 55 years 52 33 – –

Kaji 2002 183 HD 7.6 (hip) – – –

Urena 2003 70 HD 30 7 – –

Block 2004 40,538 HD, Fresenius – – 0.52 –

Inaba 2005 114 PD > 65 years – 18 – –

Danese 2006 9007 USRDS – – – 0.65

Elder 2006 242 Pre-Tx – 28 – –

Ersoy 2006 292 PD 10 – – –

Jadoul 2006 12,782 HD, DOPPS 2.6 (hip) – 2.56 0.89

Jamal 2006 52 HD > 50 years old 52 – – –

Kaneko 2006 7159 USRDS – – – 1

Mitterbauer 2007 1777 HD – – 4.1 –

Mares 2008 73 HD – 21 – –

Iimori 2012 485 HD – 6 1.9 –

DOPPS: Dialysis Outcomes and Practice Patterns Study; HD: hemodialysis; PD: peritoneal dialysis; tx: transplant; USRDS: United States Renal Data System.
Source: data from KDIGO (2009) [4] and Iimori et al. (2011) [123].

VII.  SPECTRUM OF SECONDARY OSTEOPOROSIS


Pathophysiology 1389
in the section about skeletal manifestations of CKD later In the alendronate Fracture Intervention Trial, 45% of
in this chapter. subjects had CKD stage 3 or 4 CKD [15] as did 37% of
patients in the denosumab study [16] and 62% in the ral-
Fracture Rates oxifene study [17].
Abnormal bone quality or quantity resulting from
ROD can lead to increased fragility and an increased risk
of fracture. In 1966, Pendras and coworkers [3] report- PATHOPHYSIOLOGY
ing their experience with the first 22 patients to receive
long-term hemodialysis, found that bone and mineral Abnormal Mineral Metabolism
disorders emerged as one of the most troublesome com-
plications, with fractures occurring in 47%. Since then, Overview
several studies of fracture prevalence and incidence have The pathophysiology of mineral metabolism in
been reported (Table 58.1) [4]. Prevalence rates have var- renal failure is complex (Fig. 58.2). The kidney plays an
ied from 10% to 40% in general dialysis populations and essential role in metabolism of calcium, phosphate, and
the incidence of hip fracture for patients commencing magnesium, not only by renal reabsorption and excre-
dialysis from 1989 to 1996 in the US was 4.4-fold that of tion, but also by secretion of systemic hormones. These
residents of Olmstead County [5]. Not surprisingly, frac- changes result in a spectrum of bone disease that can
tures occur more commonly in elderly patients, in women, prove ­difficult to predict for an individual patient.
in diabetic patients, in those using g
­ lucocorticoids, and in Calcium and phosphate balance are central to many
those with longer exposure to dialysis. of these pathways, and bone and the kidney work
Fracture risk is also increased in earlier stages of CKD cooperatively to maintain mineral balance. To main-
[6]. In CKD stages 3 and 4, hip fractures occur two to tain homeostasis as GFR declines, the amount of a sub-
three times more commonly than among people without stance generated (or absorbed from the intestine) must
CKD [4,7], and an increased risk has been reported in equal the amount excreted. Substances such as creati-
both black and white women [8]. In elderly women [9] nine that are filtered with little tubular handling remain
and men [10] those with higher cystatin C levels have in balance, because rising plasma levels increase their
significantly more fractures, and this may be a more filtration and excretion by each remaining nephron.
sensitive indicator of fracture risk than the creatinine, However potassium, calcium, and phosphate would
because cystatin, as noted above, does not depend on be lethal if their levels rose to maintain balance like cre-
muscle mass [10]. For elderly patients, levels of eGFR atinine does. These solutes undergo both filtration and
less than 45 mL/minute/1.73 m2 are also associated with tubular handling, so that with progressive nephron
a near doubling of hip-fracture-related mortality both loss, each intact nephron excretes a greater fraction of
before and after adjustment for confounding factors [11]. its filtered load. For example, while each nephron may
need to excrete only 10% of its filtered phosphate load
Coexistence of Osteoporosis and to maintain homeostasis at normal GFR, 90% may need
Chronic Kidney Disease to be excreted by residual nephrons at low GFR levels.
Early stages of CKD are common in elderly people Despite these adaptations, as CKD progresses the abil-
who have low bone density because both osteoporosis ity of remaining nephrons to excrete the daily load of
(defined as a T-score lower than –2.5) and CKD increase phosphate is eventually exceeded, resulting in positive
with aging. In the Third National Health Assessment balance. The intestinal absorption also contributes to the
and Nutritional Examination Survey (NHANES III), balance of minerals, and with increasing kidney dam-
84% of women with osteoporosis had CKD stage 3 to 4. age the 1,25-dihydroxyvitamin D (1,25D) levels decrease
The survey also found that 26% of women and 17% of resulting in lower calcium and phosphate absorption.
men with CKD stages 3 to 5 had coexisting osteoporosis The dietary intake is another important factor that will
[12]. These percentages varied with the age distribution contribute the overall mineral balance in these patients.
of the population sampled.
Clinical trials of postmenopausal osteoporosis therapy Phosphate
have generally excluded patients with known kidney The phosphate bond is essential to life because it
disease based on serum creatinine levels. Consequently, allows cells to control energy. Phosphorylation is nec-
the proportion of patients with CKD in these trials is essary for enzyme activation, deoxyribonucleic acid
lower than in the overall population. Nevertheless, 45% (DNA) replication, and glycolysis. Intracellular and
of subjects in the pivotal teriparatide Fracture Preven- extracellular concentrations of phosphate and calcium
tion Trial had eGFR levels less than 80 mL/minute/1.73 m2 are tightly controlled by inter-related regulatory mecha-
[13] and 52% of subjects in the pooled risedronate trials nisms, in which the kidney plays a pivotal role. Phos-
had eGFR levels lower than 50 mL/minute/1.73 m2 [14]. phate is ubiquitous in foods, particularly those high

VII.  SPECTRUM OF SECONDARY OSTEOPOROSIS


1390 58.  OSTEOPOROSIS ASSOCIATED WITH CHRONIC KIDNEY DISEASE

in protein. This makes it challenging to reduce dietary calcium, but unlike phosphate, calcium does not shift
phosphate. Its active gastrointestinal absorption is medi- into cells, and any excess must be excreted by the ­kidneys
ated by 1,25D and its excretion is predominantly renal. or deposited in the skeleton or soft tissues.
Excretion is regulated by parathyroid hormone (PTH)
and fibroblastic growth factor 23 (FGF23), both of which Magnesium
cause internalization and degradation of the proximal To maintain normal serum magnesium levels, around
tubular sodium-dependent phosphate co-transporter 100 mg (4.1 mmol) must be excreted daily by the kid-
(NPT)2a. Klotho, another hormone secreted in the neph- neys. The principal storage sites for magnesium are
ron, is a coreceptor for FGF23, and is necessary for its muscle (20%) and bone (67%), where it is complexed
homing to target organs [18,19]. These hormones that with apatite crystals at a calcium:magnesium ratio of
regulate phosphate homeostasis are further discussed in 50:1. About 20% to 30% of total bone magnesium exists
Chapter 16 (Kumar and Tebben). as a surface-limited ion, which is freely exchangeable
Positive phosphate balance resulting from progres- with serum magnesium [21]. Like other intracellular
sive CKD is often reflected by increased serum phos- ions, serum magnesium levels do not necessarily reflect
phate concentrations, which are directly associated with total body balance.
cardiovascular risk and mortality [20]. Low phosphate Most disturbances of magnesium metabolism result
levels are much less common in CKD, although hypo- in renal or gastrointestinal magnesium losses, so CKD
phosphatemia may occur when patients undertake long is unusual for being a cause of positive magnesium bal-
dialysis hours, such as with nocturnal hemodialysis. ance. However, hypomagnesemia can occur in CKD,
related to drugs such as loop diuretics, calcineurin
Calcium inhibitors, aminoglycosides, cisplatin, amphotericin B,
Like phosphate, serum calcium concentrations are and proton pump inhibitors, and following parathyroid-
also maintained within strict limits. High extracellu- ectomy or transplantation. Epidemiological studies sug-
lar calcium can change membrane potentials necessary gest that low levels may be a risk factor for osteoporosis.
for cardiac, muscular, and neurological function. Low Elevation of magnesium levels is generally caused
extracellular calcium also predisposes to arrhythmias by the use of magnesium-containing antacids and
and muscle dysfunction, and chronically low calcium phosphate binders and by the dialysate magnesium
levels lead to osteoporosis or osteomalacia. Just as for content. High magnesium levels may inhibit miner-
phosphate, 1,25D increases the intestinal absorption of alization of bone and reduced dialysate magnesium

FIGURE 58.2  Complex effects of kidney disease on skeletal metabolism. The dashed lines represent “frustrated” feedback loops. For example,
higher phosphate levels increase parathyroid hormone (PTH) secretion, which normally decreases serum phosphate. However, chronic kidney
disease interrupts this feedback loop because kidney damage prevents a phosphaturic response to PTH. The resulting skeletal changes are shown
in the rectangles, and can vary from high to low bone formation and resorption rates. In addition, both high resorption and low formation could
contribute to vascular calcifications. BMP: bone morphogenetic protein; FGF: fibroblastic growth factor; IGF: insulin-like growth factor.

VII.  SPECTRUM OF SECONDARY OSTEOPOROSIS


Pathophysiology 1391
concentrations have been associated with an improve- particularly when mineral balance is positive as occurs
ment in osteomalacia [22]. However, long-term con- in many patients on dialysis. Other factors involved with
sequences of positive magnesium balance and the ­vascular calcification are described later in this chapter.
abnormal magnesium to calcium ratio found in bone
biopsy studies [23] remains unclear. Magnesium com- Parathyroid Hormone, Vitamin D and
pounds are effective phosphate binders [24], which
FGF23 and Klotho
may reduce the development of arrhythmias and in
observational studies of patients on dialysis, higher The three major circulating hormones that control
serum levels may have improved survival [25] and mineral metabolism are involved in complex feedback
slow the progression of vascular calcification. How- networks, as shown in Figure 58.3. The direct conse-
ever, the only prospective evaluation of vascular calci- quences of renal disease include an inability to excrete
fication in magnesium-supplemented ­dialysis patients phosphate and reduced production of 1,25D and Klotho.
was small and n­ onrandomized [26]. This indirectly leads to increased concentrations of PTH
and FGF23.
The Role of Bone Turnover
With progressive CKD and a reduced renal excretory Parathyroid Hormone
capacity for calcium and phosphate, bone is a key site Secondary hyperparathyroidism is seen in many
for mineral exchange and sequestration. Bone acts as patients with CKD 3 to 5, and PTH concentrations
a “sink” to buffer the exchangeable mineral pool, with increase as kidney function deteriorates. The “intact”
hydrated osteoid covering the bone surfaces and chan- PTH (iPTH) assays measure not only PTH 1-84 but also
nels acting as a potential site for immediate exchange. long C-terminal fragments, particularly PTH 7-84. These
The ability of bone to continue to perform the function fragments are normally metabolized by the kidney
of mineral sequestration is predicated on normal bone and can contribute up to 50% of PTH measured by the
physiology. Low bone turnover results in quiescent bone “intact” assay (as compared to the “biointact” assay) in
surfaces. The collagen matrix is more densely packed later stages of CKD [27]. Several factors contribute to sec-
with minerals and has lower water content, and thus is ondary hyperparathyroidism, including low circulating
more highly mineralized. This would make it more diffi- 1,25D, low serum calcium, and high serum phosphate
cult to deposit further mineral into the bone. In addition, [28]. These factors increase the PTH concentrations by
the osteocytes are older, so that the ability of this bone increasing the amount of PTH secreted per cell, and/or
to exchange calcium and phosphate rapidly diminishes. by increasing parathyroid cell proliferation. This can lead
In contrast, high turnover often results in a net efflux to diffuse glandular hypertrophy followed by the devel-
of mineral from bone. In both cases, loss of the bone opment of nodules with similar gene profiles in each
“sink” may increase the risk of vascular calcification, nodule and finally a stage of uninodular hyperplasia,

FIGURE 58.3  A shows the feedback network for mineral-regulating hormones in normal persons and B indicates disruptions with chronic
kidney disease. Red lines indicate inhibition and blue arrows indicate stimulation. With kidney disease, the phosphate, parathyroid hormone
(PTH) and fibroblastic growth factor (FGF)23 increase but the calcium, 1,25D and Klotho decrease. Phosphate increases because the kidney can no
longer excrete it, and the kidney is no longer able to respond to the phosphaturic actions of increased FGF23 and PTH levels.

VII.  SPECTRUM OF SECONDARY OSTEOPOROSIS


1392 58.  OSTEOPOROSIS ASSOCIATED WITH CHRONIC KIDNEY DISEASE

which is more resistant to medical treatments. Malignant


transformation is not common [29].
The progressive decline in levels of circulating 1,25D is
a major factor in these changes, because this hormone nor-
mally suppresses PTH levels by several mechanisms. 1,25D
directly inhibits transcription of the PTH gene [30] exerts
an antiproliferative effect on parathyroid cells through
induction of p21 gene expression and upregulates expres-
sion of the calcium-sensing receptor (CaR) [31], which sen-
sitizes parathyroid cells to the inhibiting effects of serum
calcium. 1,25D also indirectly reduces PTH ­secretion by
increasing gastrointestinal calcium absorption.
Low calcium levels also cause secondary hyperpara-
thyroidism by stimulating PTH secretion via the CaR.
With persistent hypocalcemia the expression of the PTH FIGURE 58.4  Vitamin D levels in chronic kidney disease (CKD)
gene is enhanced, due to greater stability of messenger stages with different parathyroid hormone (PTH) concentrations.
ribonucleic acid (mRNA) for PTH and when hypocal- Source: reproduced from Patel et al. (2011) [38], with permission.
cemia is even more prolonged parathyroid hyperplasia
ensues [32]. Increased levels of serum phosphate also this relationship is not evident for patients on dialysis.
result in hyperparathyroidism by stabilizing PTH mRNA Reduced 1,25D exacerbates hypocalcemia, and together
and when prolonged, hyperphosphatemia reduces p21 these contribute to the development of secondary hyper-
and contributes to parathyroid gland hypertrophy [33]. parathyroidism. When used in management, calcitriol
A further mechanism of parathyroid hypertrophy has and its analogs reduce bone turnover, improve miner-
been identified. Reduced levels of 1,25D or high levels of alization, and stabilize or improve bone volume [39].
dietary phosphate induce translocation to the cell surface However, oversuppression of bone turnover is a major
of a tumor necrosis factor-α-converting enzyme (TACE), concern. Less well studied are the effects of 1,25D on
which results in release of the growth promoter trans- blood vessel walls and vascular calcification, which vary
forming growth factor (TGF)-α [33,34]. In turn, TGF-α with 1,25D concentration and other factors [40,41].
binds to and activates the epidermal growth factor recep-
tor (EGFR). Activated EGFR then translocates to the FGF23 and Klotho
nucleus, where it functions as an autocrine signal for cell FGF23 is an endocrine hormone secreted by the osteo-
growth. With phosphate restriction, parathyroid TGF-α cytes that acts on the proximal tubule to increase phosphate
returns to normal levels rapidly, which suggests that low- excretion, and on the parathyroid gland to reduce secretion
ering dietary phosphate may counteract uremia-induced of PTH [42]. FGF23 also directly reduces 1-α-hydroxylation
parathyroid hyperplasia by preventing parathyroid TGF-α of vitamin D [43] and increases the activity of the catabolic
enhancement in addition to inducing expression of p21. 24-hydroxylase enzyme (CYP 24A1). This latter action
FGF23 reduces PTH by suppressing the PTH gene enhances the conversion of 1,25D to 1,24,25(OH)3D and
[30], but in CKD the parathyroids are resistant to this reduces the availability of substrate to the 1-alpha hydrox-
action, perhaps due to lower levels of Klotho [35]. ylase enzyme by converting 25(OH)D to 24,25(OH)2D
[44]. Klotho is a protein expressed in the kidneys and
Vitamin D ­parathyroids that is ­necessary for the action of FGF23.
A decline in 25-hydroxyvitamin D (25(OH)D) levels
often accompanies progressive CKD [36,37], although FACTORS THAT INCREASE CIRCULATING
values may be similar to those in the general community. FGF23 LEVELS
In contrast, 1,25D levels decline progressively, due to an Multiple different factors modify FGF23 levels [45]
early rise in FGF23 levels inhibiting 1 alpha-hydroxylase and some of these interact with each other (Fig. 58.5). One
(CYP27B1) activity as well as inducing 24-hydroxylase driving factor for FGF23 production is a chronic increase
activity (CYP24A1), resulting in 1,25D breakdown. With in phosphate load, although the mechanism is uncer-
worsening CKD, proximal tubular damage reduces the tain, and some studies do not show an increased FGF23
ability of tubular cells to produce 1,25D and concen- immediately after increasing the serum phosphate [46].
trations are generally low; this may be exacerbated by Increased serum phosphate levels also stimulate the post
poor sunlight exposure and reduced 25(OH)D levels in translational O-glycosylation of FGF23, which protects it
patients who are not supplemented with cholecalciferol. from cleavage and increases its stability [47]; the enzyme
Low levels of 25(OH)D are inversely related to PTH involved being coded by the polypeptide N-acetylgalac-
concentrations from CKD stage 1 to 4 [38] (Fig. 58.4) but tosaminyltransferase 3 (GALNT3) gene. FGF23 is also

VII.  SPECTRUM OF SECONDARY OSTEOPOROSIS


Pathophysiology 1393
had lower serum phosphate levels and increased renal
tubular fractional excretion of phosphate compared to
controls [59]. This suggests that hyperphosphatemia
is not the only stimulus for FGF23 production. Simi-
lar results were seen in another study of 3879 patients
with CKD stage 2 to 4. The mean serum phosphate and
median PTH levels were still in the normal range, but
median FGF23 was markedly greater than in healthy
populations, and increased significantly with decreas-
ing eGFR [60]. In children with CKD, levels of FGF23
are also reported to increase prior to the development of
hyperphosphatemia [61,62].
FIGURE 58.5  Interactions between mediators of fibroblastic
growth factor (FGF)23. Phosphate loading increases parathyroid hor- In a mouse model of renal injury, the serum FGF23
mone (PTH) and inhibits 1,25D. Phosphate loading also increases levels were increased prior to elevations in blood urea
FGF23 but the action is indirect. PTH directly stimulates FGF23 and nitrogen (BUN) and serum creatinine, but the gene tran-
1,25D, and also inhibits sclerostin and PHEX. 1,25D increases FGF23 scription in osteoblasts was not increased until later
directly. Sclerostin inhibits the PTH-stimulation of FGF-23, but also
stages. The source of the early increase in FGF23 was not
inhibits PHEX. PHEX and Dmp1 both inhibit FGF23 through uncertain
mechanisms. Leptin and low iron both stimulate FGF23. clear, but the authors thought it could represent changes
in excretion or post-transcriptional processing of the
hormone [63]. Most of the FGF23 in patients with CKD
cleaved by the protein furin, so any factors that inhibit is intact [64], suggesting that there may be something in
furin would increase intact FGF23 levels [48]. FGF23 pro- uremic serum that inhibits the cleavage of FGF23 [48].
duction is decreased by both PHEX and Dmp1, through
uncertain mechanisms; mutations in either of these FGF23 LEVELS AND CLINICAL OUTCOMES
­proteins cause hypophosphatemia and rickets. High FGF23 concentrations are associated with poor
FGF23 is upregulated by PTH and by 1,25D. 1,25D clinical outcomes, including cardiovascular disease, both
acts through osteocyte vitamin D receptors to increase in patients with renal disease [65–67] and in the general
FGF23 transcription [49,50], while PTH increases FGF23 population [68]. Compared with the lowest quartile of
both directly and indirectly [51,52]. The direct PTH effect FGF23, each subsequent quartile was associated with a
is via activation of the protein kinase A (PKA) pathway progressively higher risk for death. However, it is not
in osteoblasts, which increases FGF23 mRNA levels and clear if these problems are caused by FGF23 or if it is a
the indirect PTH effect is by the conversion of 25(OH)D surrogate marker for disease.
to 1,25D as well as by decreasing PHEX. On a clinical
level, lowering PTH levels in hemodialysis patients with FGF23 LEVELS AND BONE DISEASE
cinacalcet also lowers levels of FGF23 [53] and FGF23 Emerging evidence suggests that high levels of FGF23
levels fall following parathyroidectomy [26]. The wnt could have negative effects on bone tissue. For exam-
signaling pathway is also involved in FGF23 modula- ple, exogenous FGF23 applied to cultured osteoblastic
tion, through pathways with opposing effects; PTH- MC3T3.E1 cells in the presence of soluble Klotho inhib-
stimulated increases in FGF23 expression are blocked ited mineralization and proliferation [69]. In patients on
by sclerostin [54], but sclerostin also inhibits PHEX [55]. hemodialysis, FGF23 values did not correlate to bone
Studies also suggest that leptin increases FGF23 [45]. mass, or to serum levels of bone specific alkaline phos-
Iron deficiency increases FGF23 secretion in normal phatase or C-telopeptide [70]. However, in stage 1 to 4
people [56], but the protein is degraded so that the intact CKD, FGF23 levels were inversely correlated with bone
molecule levels return toward normal while the frag- mineral density at the hip [57]. A study of 105 Japanese
ments remain increased. In CKD stage 1 to 4 a study patients with CKD found that those with vertebral
found that the serum iron was inversely correlated to ­fractures had higher FGF23 levels [71].
serum FGF23 using an assay that measured both intact
and C-terminal fragments [57]. However, iron infusions KLOTHO
also can increase FGF23 in dialysis patients [58]. It is pos- Klotho is a protein named after the first of the Moi-
sible that iron infusions inhibit furin, which stabilizes rae, who were the three Greek goddesses controlling
the intact FGF23 levels [48]. man’s fate. Klotho spun the thread of life, so it was
appropriate to assign her name to a transgenic mouse
FGF23 IN EARLY CHRONIC KIDNEY DISEASE gene that conferred longevity above the wild type when
An increase in FGF23 levels is detected in very early overexpressed, but upon deletion resulted in prema-
CKD, such as in otherwise healthy kidney donors who ture aging and a shortened life span [72,73]. Klotho is

VII.  SPECTRUM OF SECONDARY OSTEOPOROSIS


1394 58.  OSTEOPOROSIS ASSOCIATED WITH CHRONIC KIDNEY DISEASE

a 130-kDa single-pass transmembrane protein expressed taking estrogen have lower phosphate and lower FGF23
in the nephron, the parathyroids, and the choroid plexus. levels then women not taking estrogen [83].
Although expression is higher in the distal nephron,
Klotho is also found in the proximal tubule [18]. Klotho Insulin and Osteocalcin
is necessary for the actions of FGF23, which does not In osteoblasts, insulin is anabolic, acting through the
have a specific receptor in the proximal tubule, but acts osteoblast insulin receptor and stimulating proliferation
through a more general FGF23 receptor. Once Klotho and differentiation [84]. Osteocalcin, long known to be
binds to this nonspecific FGF receptor, it converts to a secreted by osteoblasts and an important bone protein, is
receptor that will bind FGF23 [74]. Klotho is also impor- now known to also be a systemic hormone that increases
tant in the regulation of insulin-like growth factor (IGF) insulin sensitivity [85]. States of high bone turnover are
and in tubular calcium transport. associated with higher uncarboxylated osteocalcin.
Diabetes is a common cause of CKD and a known risk
factor for osteoporosis and fractures (see also Chapter
Acid-Base
55). Diabetic patients tend to have low bone formation,
With acute metabolic acidosis, calcium and bicar- which could exacerbate the diabetes because they will
bonate are released from bone in a chemical reaction have lower uncarboxylated osteocalcin. Osteoblasts also
that does not depend on cellular activity, although with mediate the adverse effects of glucocorticoids on fuel
chronic metabolic acidosis, osteoclastic bone resorp- metabolism [86].
tion is increased as well [75]. The released bicarbonate A study of 189 hemodialysis patients found that under-
buffers excess acid, but at the expense of loss of bone carboxylated osteocalcin (ucOC) correlated inversely
mineral. Respiratory acidosis does not cause these prob- with the plasma glucose, hemoglobin A1C, and gly-
lems, suggesting that the bicarbonate concentration is cated albumin. Diabetic patients had lower serum ucOC
the key factor. Since diseased kidneys are less capable levels [87].
of secreting an acid load, these mechanisms may cause
bone loss, even in early stages of CKD. In healthy elderly
Malnutrition, Insulin-Like Growth Factor
patients a randomized clinical trial for 2 years showed
an increase in bone density as well as markers of bone Patients with CKD usually have nutritional prob-
formation with potassium citrate supplementation [76]. lems. Their diets must be restricted because they can-
Another study showed decreases in the urine calcium not excrete excess phosphate or potassium or sodium.
excretion with potassium bicarbonate [77]. A Cochrane Nausea and anorexia limit the intake of food. This mal-
review found that the evidence for benefits of acid cor- nutrition would be expected to play a role in the devel-
rection is very limited in patients with CKD, but small opment of bone disease, but this has not been carefully
trials suggested there may be some beneficial effects on studied.
bone metabolism [78]. Serum levels of IGF-I must be interpreted with cau-
tion, because the bioavailability of IGF-I is altered in kid-
Abnormal Endocrine Hormones ney disease [88]. In the US population, serum IGF-I is
positively associated with estimated GFR [89]. However,
Gonadal Hormones Jehle et al. [90] studied 319 patients with various stages
Patients with late stages of CKD frequently have of kidney disease, including dialysis patients, and found
abnormalities in gonadal hormones, which can con- that total and free IGF-I levels remained constant in most
tribute to bone disease [79]. Many premenopausal patients. The IGF-I was positively correlated with the
women with stage 4 or 5 CKD have oligomenorrhea or bone formation rate.
amenorrhea, and serum estrogen levels are decreased. Studies report various levels in dialysis patients
However, albumin and sex hormone binding globulin [91,92]. The bioavailable levels of IGF may be low even
(SHBG) may also be decreased, so free estradiol levels when total levels are increased because the binding pro-
can be higher than suggested by total estradiol concen- teins are also increased. Han et al. [92] found lower IGF-I
trations [80]. Men may also have low total testosterone levels in patients on dialysis than in normal subjects.
levels. These patients also had abnormalities in their muscle
In men with stage 3 CKD, estradiol levels are associ- strength.
ated with a slight but significant decrease in the serum
phosphate, although the relationship is stronger in men
Wnt-Signaling Pathway
with normal kidney function [81]. This may be due to
an estrogen-induced increase in levels of FGF23 as In 2001, identification of the high bone mass gene
described in a rodent model [82]. However, in women as LRP-5, a coreceptor for the wnt-signaling path-
with heart disease and mild to moderate CKD, those way, led to numerous studies which documented the

VII.  SPECTRUM OF SECONDARY OSTEOPOROSIS


Pathophysiology 1395
importance of this pathway in skeletal metabolism [93] Accumulation of Substances
(see also Chapter 18). Sclerostin and Dickkopf (Dkk-1)
are two soluble proteins which inhibit wnt signaling Aluminum
by binding to LRP-5. Secreted frizzled-related proteins About a decade after long-term dialysis was first used,
(sFRPs) also inhibit the pathway. Sclerostin is secreted some centers reported a unique syndrome of dementia.
almost exclusively by osteocytes and tonically inhibits The first symptom was stuttering during dialysis, and
bone formation. When the bone is subjected to mechani- this rapidly progressed to seizures, coma, and death
cal forces, the affected osteocytes stop secreting scleros- within 6 months. The patients also had osteomalacia.
tin and bone formation commences at that location [94]. The etiology was a mystery and the most likely cause
Dkk-1 is expressed by many kinds of cells, including seemed to be a virus. Several centers had epidemics of
those in the marrow and vasculature. this disease, while others did not have any cases. Even-
Alterations to wnt signaling and levels of wnt inhibi- tually these outbreaks were linked to the aluminum lev-
tory proteins are seen in patients undergoing dialysis [95] els in the water used for the dialysate. Municipal water
and in animal models of CKD [96]. In a mouse model, an companies may add aluminum to deflocculate water,
early increase in expression of SOST (the sclerostin gene) and in several cities the dialysis centers started seeing
led to reduced wnt signaling and consequently reduced dementia soon after the water department began using
osteoprotegerin (OPG). this form of water treatment. Meanwhile, investigators
In bone biopsies from 60 patients with stage 5 CKD, had described two forms of osteomalacia in dialysis
Cejka and colleagues [95] investigated the associations patients: one responded well to calcitriol but the other
between serum levels of sclerostin or Dkk-1 and histo- was resistant and caused by aluminum [103]. Thereaf-
morphometric parameters of bone turnover, mineraliza- ter, all dialysis centers began using methods to purify
tion, and volume (TMV). Serum sclerostin levels were the water, and the incidence of aluminum-related osteo-
higher in CKD patients than in normal women. Scleros- malacia dropped dramatically. Currently only sporadic
tin correlated negatively with iPTH and with histomor- cases are seen, with high doses of oral aluminum or
phometric parameters of turnover, osteoblastic number, with accidental contamination of dialysate or parenteral
and function in both unadjusted and adjusted analyses. solutions.
Furthermore, sclerostin was superior to PTH for positive
prediction of high bone turnover. Dkk-1, however, did Iron
not show correlations with any of the parameters of bone Iron accumulation in the bone can also cause osteoma-
turnover measured on bone biopsies. lacia in dialysis patients. On bone biopsies, iron deposi-
Although wnt signaling is beneficial to the skeleton, tion is seen along the surface of the bone, particularly
it also acts in other systems, where it may play a nega- at the mineralizing front [23,104]. This can be seen in
tive role. For example, in the kidneys wnt signaling leads patients with frequent blood transfusions. Osteoporosis
to podocyte dysfunction, proteinuria [97], and increased is seen in patients with hemochromatosis who have nor-
epithelial damage [98]. Wnt inhibition is involved with mal renal function [105]. The iron appears to be directly
vascular calcifications, as discussed later in this chapter. toxic to the osteoblasts [106] (see also Chapter 48, Marcus
and Shane).
Bone Morphogenetic Protein-7 Strontium
Bone morphogenetic protein (BMP)-7 is an impor- Some animal studies suggest that accumulation of
tant promoter of osteoblast differentiation that is strontium may also cause osteomalacia [107]. In rats,
produced primarily by the kidneys, so that levels are the effects are complex and dose-dependent; some had
reduced in CKD [99]. This growth factor is necessary adynamic bone and others developed osteomalacia
for the differentiation of preosteoblasts into mature [108]. Strontium levels are increased in bones of dialy-
osteoblasts. The accumulation of preosteoblasts in sis patients who had osteomalacia [109,110]. This has
the marrow had traditionally been termed “fibro- not been seen in osteoporotic patients who have been
sis” because the long spindle-shaped cells resembled treated with strontium, but they did not have kidney
­fibrocytes [100]. disease [111]. There are no clinical trials using strontium
BMP-7 has been used successfully to treat the ady- in patients with CKD. Use of strontium as a treatment for
namic bone disorder that developed in adult mice after osteoporosis is discussed in Chapter 84.
the induction of renal impairment [101]. Because BMP-7
upregulates the expression of osteoblast PTH receptors, β2 Microglobulin (Amyloid)
higher levels of PTH may be required for regulation of β2 microglobulin is a byproduct of immunoglobulin
stem cell niches when levels of BMP-7 are reduced by metabolism, and is normally excreted by the kidney. This
renal damage [102]. protein is too large to be efficiently dialyzed, and it may

VII.  SPECTRUM OF SECONDARY OSTEOPOROSIS


1396 58.  OSTEOPOROSIS ASSOCIATED WITH CHRONIC KIDNEY DISEASE

accumulate in patients after many years of hemodialy- ubiquitin-proteosome system. Abnormalities in insulin
sis. The protein aggregates form β-pleated sheets, which /IGF-I signaling activate caspase-3, a protease that dis-
results in dialysis-related amyloid, but the distribution rupts the complex structure of muscle proteins [118].
is different than that of other forms of amyloid. Deposits In dialysis patients, grip strength is reduced compared
are seen in the bone near the joints, where it binds to col- to normal controls, and is lower in dialysis patients
lagen and stimulates osteoclastic bone resorption, form- with higher serum myostatin levels [92]. Relationships
ing cystic lesions [112]. Amyloid also deposits around the between nerve and bone are being discovered, and the
tendons of the wrist and causes carpal tunnel syndrome neuropathy associated with CKD may play a role in the
(Fig. 58.6) [113]. High-flux dialysis membranes are more bone disease. Neuropathy also can be a factor in devel-
efficient at removing the β2 microglobulin [114]. opment of stress fractures of the feet due to reduced
proprioception.
Low Activity and Poor Muscle Strength
Patients with CKD 4 or 5 have multiple medical SKELETAL MANIFESTATIONS OF
problems that may result in decreased physical activity. CHRONIC KIDNEY DISEASE
Measurements of muscle strength by the timed get up
and go test, and by the 6-minute walk, predict fractures Fractures
in patients with CKD. In one study the area under the
receiver operating curve (ROC) with these muscle tests Fractures Predict Mortality and Further Fractures
was 0.90, which was higher than the area using high- Mortality in patients with CKD stage 5D who have
resolution peripheral quantitative computed tomogra- had a hip fracture is about twice as high as mortality in
phy (HRpQCT) at the radius (area under the ROC 0.80) patients of similar age and gender who have not had a
[115,116]. Muscle strength is just one factor that can lead fracture. Coco et al. [119] followed 1272 hemodialysis
to falls, and 44% of dialysis patients older than 65 years patients over 10 years and observed that the mortality
had more than one fall in the previous year [117]. In for CKD stage 5D patients with a hip fracture was 2.7
elderly men and women treated for osteoporosis, CKD times higher than in fracture-free hemodialysis patients.
is a risk factor for falls [6]. Mittalhenkle et al. [120] recorded hip fracture cases over
Muscle wasting (or sarcopenia) is highly preva- 5.5 years, and the mortality incidence was 2.15 times
lent among patients with CKD, and may not be sim- greater in the cases as in matched controls. Danese et al.
ply responsive to protein feeding. In these patients [121] evaluated 9007 patients and found that a history of
there are complex mechanisms that stimulate loss of hip, vertebral, or pelvic fracture was associated with an
skeletal muscle, involving activation of mediators that age- and sex-adjusted mortality rate that was 2.7 times
stimulate the adenosine triphosphate (ATP)-dependent greater than for the other dialysis patients. Kaneko et al.
[122] found that the adjusted hazard ratio for mortal-
ity was 1.95 in patients with long-bone fractures, using
data from 7159 subjects in the Dialysis Morbidity and
­Mortality Study.
In the general population, previous fractures as an
adult are strongly associated with the risk of a subse-
quent fracture. In patients with CKD stage 5D, a ­vertebral
fracture identified on a radiograph increased the risk of a
new fracture by over seven-fold [121].

Relationship of Parathyroid Hormone and Alkaline


Phosphatase to Fractures
In patient with CKD stage 5D, several large prospec-
tive studies relating serum PTH to fractures have pro-
vided inconsistent results [4,123]. Cross-sectional studies
have also evaluated this relationship and in general were
negative (Table 58.2). However, a case controlled cohort
study did find a 31% (95% confidence interval (CI) 0.57
to 0.83, p < 0.001) reduction in global fracture risk follow-
FIGURE 58.6  Dialysis amyloid. The radiograph demonstrates ing parathyroidectomy [124].
cysts in the radius and carpal bones. Source: reprinted from Adams (2002) An association between high serum alkaline phospha-
[113], with permission. tase levels and the risk of fractures has been reported [125].

VII.  SPECTRUM OF SECONDARY OSTEOPOROSIS


Skeletal Manifestations Of Chronic Kidney Disease 1397

Bone Pain osteitis fibrosa (high bone turnover with “fibrosis”), and
mixed (high bone turnover with abnormal mineraliza-
Whereas ordinary osteoporosis does not cause pain tion). There is no consensus on the exact measurements
until the patient experiences a fracture, ROD is associ- to define these categories. A review of 2340 biopsies
ated with bone pain, especially when there is osteoma- performed on Brazilian patients for clinical indications
lacia. Severe secondary hyperparathyroidism may also found that the frequency of fractures was significantly
cause bone pain. higher in those with osteomalacia compared with other
forms [127]. Another study that followed 62 patients for
Bone Histology 5 years after bone biopsy found a higher rate of fractures
in those with adynamic bone disease [128].
Types of Renal Osteodystrophy Bone volume was not previously used to categorize
In ROD, bone formation rate varies from unmeasur- biopsies, but on average those with adynamic bone have
ably low to very high [126]. Superimposed upon this lower bone volume, and those with high turnover have
spectrum are variations in the degree of mineralization increased cancellous bone volume (Fig. 58.7).
of the bone matrix and variations in bone volume, which A comprehensive literature review of the prevalence
combine to determine the bone density. An international of types of bone disease in CKD in shown in Figure 58.8.
committee sponsored by Kidney Disease: Improving The review analyzed studies carried out between 1983
Global Outcomes (KDIGO) concurred that the most and 2006 [4]. Differing prevalences of bone disease types
important parameters to describe ROD are turnover, observed between studies are due to differing classifi-
mineralization, and volume (TMV) [1,4]. Thus, instead cation methods, in addition to differences related to
of a one-dimensional axis, the TMV system defines a geographical areas, genetic background, and treatment
more comprehensive three-dimensional space. modalities.
Most older studies of ROD have classified biopsies A report from 2011 shows that the patterns of ROD
based on the bone formation rate or presence of fibrosis, may be evolving. Malluche analyzed 630 bone biopsies
as well as the degree of mineralization. The six traditional performed in patients with CKD stage 5, from 2003 to
categories are: normal, adynamic (low bone turnover 2008, using the TMV system. About 58% of patients had
and normal mineralization), osteomalacia (low bone low bone turnover, which is a shift from older studies
turnover and abnormal mineralization), high turnover, showing a preponderance of high turnover related to
hyperparathyroidism. Only 3% of patients had abnormal
TABLE 58.2  Relationship between Fractures and Parathyroid mineralization. In addition, 73% of the white patients
Hormone with low bone volume had low bone turnover [129].
Author Year n Fractures
On bone biopsies, the bone volume in black dialysis
patients is higher than that in whites and the bone for-
Coco 2000 1272 High risk with low PTH mation rate is mostly normal or elevated in blacks but
Stehman- 2000 4952 No relation mostly low in whites. This is in contrast to black patients
Breen with normal renal function who have low bone forma-
Block 2004 40,538 Weak direct association, p = 0.035 tion. In the dialysis population the black patients had
higher PTH levels than white patients (661 vs. 274 pg/
Danese 2006 9007 Higher risk with low or high PTH mL), but at any level of PTH, bone turnover was lower in
Jadoul 2006 12,782 Relative risk = 1.7 if PTH > 900 blacks than in whites, indicating resistance to the resorp-
Mitterbauer 2007 1774 No relation
tive effects of PTH. The black patients were treated more
often with vitamin D or its metabolites. In addition,
Iimori 2012 485 Higher risk with low or high PTH black dialysis patients have a greater survival rate than
Urena 2003 70 No relation white patients [130].
Negri 2004 65 No relation
Parathyroid Hormone in Relation to Bone
Inaba 2005 114 No relation Histology
Elder 2006 242 More after PTX otherwise not related The classic findings of renal hyperparathyroidism
Ersoy 2006 292 No relation
(still termed “osteitis fibrosa”) are increased numbers
of multinucleated osteoclasts, woven bone, blurry tet-
Jamal 2006 52 No relation racycline labels, increased cancellous bone volume
Atsumi 1999 187 More with low PTH but decreased cortical thickness, and intratrabecular
­tunneling (Fig. 58.9).
PTH: parathyroid hormone. Studies above line are cohort studies and below the
line are cross-sectional. The classic term is a misnomer; there is no active
Source: data from KDIGO (2009) [4]. inflammation, the “fibrosis” is predominantly cellular

VII.  SPECTRUM OF SECONDARY OSTEOPOROSIS


1398 58.  OSTEOPOROSIS ASSOCIATED WITH CHRONIC KIDNEY DISEASE

FIGURE 58.7  A, Categories of renal osteodystrophy according to turnover and mineralization. The bone volumes shown are typical values
but these results vary according to the population studied. B, Histological sections of bone from patients showing adynamic (upper left), high
bone turnover (upper right), osteomalacia (lower left) and mixed disease (lower right). The bones were stained with Goldner’s stain which colors
mineralized bone blue/green and unmineralized osteoid orange. (See color plate.) Source: A, From the KDIGO guidelines [4], with permission; B, from
Ott (2009) [126], with permission.

with some extracellular matrix, which can disappear for skeletal resistance to PTH in patients CKD-MBD.
rapidly when PTH is lowered. The cells are not fibro- This is why guidelines recommend that serum levels of
blasts, but inchoate preosteoblasts that proliferated PTH should be higher than the range seen in normal
under the influence of PTH but have not achieved individuals [4].
mature differentiation. This can be corrected by addi- Correlations between PTH and bone formation vary
tion of BMP-7 [100,131–133]. However, the bone widely (Fig. 58.10). The older studies tended to find
response to PTH is not consistent, and there is evidence better correlations between PTH and bone formation

VII.  SPECTRUM OF SECONDARY OSTEOPOROSIS


Skeletal Manifestations Of Chronic Kidney Disease 1399

FIGURE 58.8  Chart showing percentage of patients with different categories of renal osteodystrophy, from studies reported between 1983 and
2006 [4]. AD: adynamic bone disease; OF: osteitis fibrosa; OM: osteomalacia. Source: from KDIGO (2009) [4], with permission.

(A)

(B)

FIGURE 58.9  Histological sections from patients with osteitis fibrosa. A shows intratrabecular tunneling, B shows multiple large osteoclasts.
(See color plate.)

rates, whereas more recent studies show poor corre- Bone Biochemical Markers
lations. This follows a trend for findings of adynamic
bone disease with high PTH levels. The reasons for Markers Studied in Chronic Kidney Disease
the recent poor correlations between PTH and bone In patients with normal renal function, serum bio-
formation are not clear, but could involve differences chemical markers correlate with bone formation or
in the assays for PTH, secular changes in the dialysis resorption, and they predict fractures in elderly popula-
population with more diabetic and elderly patients, tions. The interpretation of these markers is altered with
differences in therapies, and differences in racial com- decreasing renal function, particularly for those mark-
position of the studies [4]. ers that are excreted by the kidneys, so that the elevated

VII.  SPECTRUM OF SECONDARY OSTEOPOROSIS


1400 58.  OSTEOPOROSIS ASSOCIATED WITH CHRONIC KIDNEY DISEASE

concentrations merely represent accumulation instead of One study evaluated CKD patients without known
increased activity of the bone cells. cardiovascular disease and found that reduced TRAP-
Markers that have been studied in CKD include the 5b and elevated b-ALP were both associated with an
collagen-based measurements of bone formation and increase in the relative risk of cardiovascular mortality
resorption (procollagen type I N-terminal propeptice [139]. These somewhat paradoxical findings suggest that
(PINP) and N-terminal telopeptide (NTX) or C-terminal much more work needs to be done to understand fully
telopeptide (CTX)), alkaline phosphatase, and bone alka- the clinical utility of such biomarkers.
line phosphatase (b-ALP), receptor activator of nuclear
factor kappaB ligand (RANKL), osteocalcin, OPG [134], Bone Mineral Density
and tartrate-resistant acid phosphatase (TRAP). Serum
levels of alkaline and acid phosphatases are not altered Bone Mineral Density in Dialysis Patients
by renal dysfunction, but the collagen-based markers are Figure 58.11 shows the average values of BMD in
excreted by the kidney. patients with CKD stage 5D [140]. These values are
expressed as z-scores, which compare BMD in the
Relationship to Bone Histology patients to the BMD from the reference values of age-
Figure 58.10 shows correlations between tetracycline- and gender-matched people in the community. Most of
based bone formation and several bone turnover mark- the studies found that spine BMD was close to normal
ers. b-ALP has better correlation to bone formation rate and relatively better than BMD at the hip or radius. This
than PTH or osteocalcin. The b-ALP also has higher is consistent with observations from bone biopsies that
predictive value for the diagnosis of high or low bone show thin cortices but increased cancellous bone volume.
turnover.
Studies Relating Bone Mineral Density to Fracture
Markers of Bone Turnover and Clinical Outcomes in Chronic Kidney Disease Stage 5
In cohorts of elderly women, most of whom have BMD measurements using dual-energy X-ray absorp-
early stages of CKD, serum biochemical markers of bone tiometry (DXA) measure only one aspect of bone
turnover have been associated with fractures [135–137] strength. This is particularly true in CKD, because fac-
(see also Chapter 67). tors in addition to bone density play a large role in deter-
In patients with CKD stages 4 or 5 there are limited mining the strength of the bone and the risk of a fracture.
data that relate serum markers to fractures. Urena et al. The KDIGO report suggested that in patients with CKD
[138] found that cross-laps TM (CTx) and b-ALP were not stages 3 to 5D with evidence of CKD-MBD, BMD test-
different between fracture and nonfracture cases in a sur- ing not be performed routinely, because BMD does not
vey of 70 dialysis patients. Limori followed 485 hemodi- predict fracture risk as it does in the general population,
alysis patients and did not see a difference in fracture rate and BMD does not predict the type of ROD [4]. Since
according to the baseline b-ALP, but the b-ALP measure- then a prospective study of dialysis patients from Japan
ments prior to the fracture date were higher than those at found that hip or whole body bone density at baseline
the end of the study in patients without a fracture [123]. was associated with fracture incidence. At the total hip,
the adjusted hazard ratio for a fracture was 0.65 for each
standard deviation increase in the BMD [123].
Figure 58.12 shows results of studies assessing the
relationship of fracture to BMD in patients on dialysis.
The results of these studies are mixed and differences
are not as great as in similar studies of postmenopausal
women from the general population [141]. There are sev-
eral potential reasons for the poor performance of DXA
in patients with CKD stage 5. Vertebral measurements
may overestimate BMD due to arthritic conditions,
scoliosis, and aortic calcifications. CKD patients have
poor bone quality that cannot be measured by DXA.
Abnormal microarchitecture, mineralization density,
FIGURE 58.10  Graph showing correlation coefficients between crystal deposition in the bone matrix, or abnormalities
markers and the bone formation rate in various studies. The corre- in the matrix itself could all contribute to loss of bone
lation coefficients are on the y-axis, and the studies are arranged in strength. Patients with CKD, especially those with high
chronological order. Symbols connected by a line were from the same
study. Symbol size is proportional to number of subjects [4]. b-ALP:
serum PTH, have increased cancellous bone volume but
bone alkaline phosphatase, OC: osteocalcin, X-link: N-telopeptide or decreased cortical thickness [142]. Furthermore, patients
C-telopeptide. Source: from KDIGO (2009) [4], with permission. with CKD may experience more trauma to the skeleton if

VII.  SPECTRUM OF SECONDARY OSTEOPOROSIS


Skeletal Manifestations Of Chronic Kidney Disease 1401
they have more frequent falls. A study of fall prevalence to 0.80 for bone density tests, in comparison to 0.90
in osteoporotic patients showed a greater risk in those for neuromuscular tests in the same group of patients
with eGFR less than 65 mL/minute/1.73 m2, especially [115,116].
when they were taking glucocorticoids [6]. Nickolas et al. [143] measured BMD by DXA and HR-
Increases in BMD do not always mean the bone pQCT in 91 patients with CKD stages 2 to 5, one-third
strength is improved, because there may be concomitant of whom had a prevalent fracture. The BMD by DXA
worsening of the bone quality. In contrast, it is difficult was 7% to 10% lower in those with a fracture. By HR-
to imagine that a loss of bone density could be beneficial, pQCT, measurements of cortical thickness at the radius
except in rare cases of osteosclerosis. and tibia were also lower in patients who had a fracture,
whereas trabecular density was lower only at the radius.
Cross-Sectional Studies in Chronic Kidney Disease The best areas under the ROC curves were 0.66 for femo-
Stages 3 to 5 ral neck DXA, the HR-pQCT at the radius and at the cor-
A Canadian study included 211 patients, age over tical bone of the tibia. There was better discrimination
18 years, with stage 3 to 5 CKD. A history of fracture between those with and without a fracture in patients
since age 40 years, or a prevalent vertebral fracture, with longer duration of CKD.
was present in 74 patients. Bone density was measured The studies of patients with early stages of CKD are
using DXA and HR-pQCT at the radius. The odds consistent with those in the general population (ROC
ratio of fracture for each standard deviation decrease area 0.61 for the total hip for osteoporotic fractures [141]).
in measurement was 1.56 for DXA of the ultradistal This is not surprising because there is a large overlap
radius and 1.45 for trabecular thickness by HR-pQCT between persons with osteoporosis and persons with
at the radius. The area under the ROC curve was 0.73 age-related decreases in GFR.

FIGURE 58.11  Plot of bone density in dialysis patients. Circles are studies in women, diamonds in men, and squares both. Size of points are
proportional to numbers of subjects. Source: from Ott (2009) [140], with permission.

VII.  SPECTRUM OF SECONDARY OSTEOPOROSIS


1402 58.  OSTEOPOROSIS ASSOCIATED WITH CHRONIC KIDNEY DISEASE

FIGURE 58.12  Studies that measured bone density in patients with fractures. All but the Iimori study were cross-sectional [123,140]. * = results
adjusted for age. Fx: fracture; nr: not reported.

EXTRASKELETAL CALCIFICATIONS
with end-stage renal failure was the development of
tumoral calcinosis [3]. Phosphate did not dialyze well,
Factors that Modify Mineralization
and the hyperphosphatemia resulted in huge mineral
Extracellular fluids are normally supersaturated with deposits. These are still seen in patients who are non-
respect to calcium and phosphate, and inhibitors of calcifi- compliant with their phosphate control (Fig. 58.14).
cation are necessary to prevent deposition (see Fig. 58.13). When concentrations of phosphate and calcium exceed
In general, patients with CKD have increased levels of a solubility threshold, the product will precipitate in
minerals and calcification promoters, but decreased lev- soft tissues.
els of inhibitors. An exception is OPG, which inhibits
calcification but is increased in CKD patients. OPG is a
calcification inhibitor, and transgenic mice lacking OPG
Calciphylaxis
have severe vascular calcifications. It is likely that the An unusual calcification of the small cutaneous
elevated levels in CKD are a reaction to this pro-calcific arteries is seen in patients with CKD, termed calciphy-
milieu and as such they are a marker of disease, increas- laxis. Another term for this is calcific uremic arterio-
ing early in the course of CKD-MBD [144] and predicting lopathy. This results in skin necrosis and can progress
adverse cardiovascular events [145]. to large areas of soft-tissue necrosis (Fig. 58.15). This
is a serious condition with high mortality. The exact
etiology is uncertain, but high PTH is often seen.
Tumoral Calcinosis
Some patients respond to parathyroidectomy or to
One of the earliest problems discovered when hemo- the antioxidant c­ation chelator sodium thiosulphate
dialysis was first able to prolong the life of patients [146–148].

VII.  SPECTRUM OF SECONDARY OSTEOPOROSIS


Extraskeletal Calcifications 1403

FIGURE 58.13  Factors that are involved with vascular calcifications. The promoters of vascular calcification are shown above and the inhibi-
tors below. The principal factors are highlighted by color. Those in ovals are decreased in chronic kidney disease, while those in flags are increased,
and those in boxes are variable or normal. ANK: transmembrane protein; BMP: bone morphogenetic protein; IGF-1: insulin-like growth factor-1;
NPP-1: pyrophosphatase/phosphodiesterase; PTH: parathyroid hormone; PTHrP, parathyroid hormone-related peptide; TGF-β: transforming
growth factor-beta [41,152–160].

FIGURE 58.15  Calciphylaxis. (See color plate.)

frequently seen in dialysis patients in soft tissues of the


FIGURE 58.14  Tumoral calcinosis in a noncompliant dialysis internal organs, including lung and stomach, and their
­patient. pathology is not well defined [150]. They contain whit-
lockite crystal patterns, as opposed to the ­hydroxyapatite
seen in bones [151].
Metastatic Calcifications
Many tissues which have been damaged or inflamed
Vascular Calcifications
may develop calcifications. This happens in people with
normal renal function, for example, with tuberculosis or Vascular calcifications (Fig. 58.16) are associated
tendonitis, and the exact mechanism is unknown. Dys- with high mortality. They are common in patients with
trophic calcifications are not metabolically active, and CKD, and by stage 5 the majority of patients exhibit
do not take up tracers on nuclear medicine bone scans vascular calcifications. These can be intimal or medial.
[149]. Other, more metabolically active calcifications are The intimal calcifications represent calcium deposition

VII.  SPECTRUM OF SECONDARY OSTEOPOROSIS


1404 58.  OSTEOPOROSIS ASSOCIATED WITH CHRONIC KIDNEY DISEASE

osteoblast markers, and secrete collagen and control


mineralization. Degradation of elastin also contributes
to the ­calcification [152,153].
The vascular smooth muscle cells release apoptotic
particles or matrix vesicles that can form a nidus for
mineralization [152,154]. In severe cases there is literally
bone inside the arteries.
Phosphate is an important signal causing smooth
muscle cells to differentiate into osteoblasts [154–158]. In
CKD, the serum phosphate levels are correlated to the
extent of vascular calcifications; furthermore, phosphate
control can retard the progression of these calcifications.
The phosphate activates Pit-1, a phosphate transporter
[152,157].
Calcitriol can have different effects on vascular cal-
cification, depending on its concentration and experi-
mental conditions [40,41]. Low concentrations can
retard the differentiation of smooth muscle cells into
osteoblasts, but at high concentrations the calcitriol,
by increasing serum calcium and phosphate, contrib-
utes to vascular calcifications. In phosphate-enriched
cultures of mouse smooth muscle cells, calcitriol or
paricalcitol altered expression of many genes that are
involved in mineralization. The changes were complex,
with increases in factors that promote calcifications as
well as inhibitors, and decreases in both promoters and
inhibitors [159].
The vascular cells express CaRs and experimental
studies show reduction in vascular calcifications when
these are activated with calcimimetics. A phase II ran-
domized clinical trial of cinacalcet combined with low
doses of paricalcitol showed a trend towards reduction in
coronary artery calcification score [160], but the phase III
study, which involved 3883 hemodialysis patients with
secondary hyperparathyroidism, found that although
the patients in the cinacalcet arm experienced numeri-
cally fewer composite primary events, the results were
not statistically significant [161].
The wnt-signaling pathway is also involved in vascu-
lar calcification. Microarrays of vascular smooth muscle
FIGURE 58.16  Vascular calcifications in a patient with chronic cells cultured with calcitriol, which caused transforma-
­kidney disease. tion to osteoblastic cells, revealed over-expression of
Dkk-1 and reduced expression of Wnt5A [159]. In rats
with CKD who developed aortic calcifications, microar-
into atherosclerotic plaques caused by high cholesterol, ray analysis of the aortic tissue showed increased sFRPs
which is another abnormality seen in late stages of [158]. It is possible that these could circulate to bone and
CKD. inhibit bone formation. In 77 patients with CKD stage
The medial calcifications are actively formed by vas- 3 to 4, sclerostin (which inhibits wnt signaling) was
cular smooth muscle cells which have transformed into increased as GFR worsened. In addition, levels of Dkk-1
osteoblast-like cells. This causes stiffness of the arter- were inversely associated with the arterial s­tiffness
ies and abnormal pulse pressure waves, with increased index [162].
systolic and decreased diastolic pressures that may Other mechanisms involved in vascular calcifications
lead to left heart failure and reduced coronary perfu- include oxidative stress [163], hydrogen sulfide [164],
sion. In tissue cultures, these vascular smooth muscle decreased Fetuin [165], decreased BMP7 [166], increased
cells spontaneously differentiate into cells that express BMP2, lower pyrophosphate, lower matrix gla protein,

VII.  SPECTRUM OF SECONDARY OSTEOPOROSIS


Treatment 1405
abnormal serum lipids, and the hostile environment of TABLE 58.3  Management of CKD-MBD from KDIGO
the uremic milieu [155]. Guidelines [4]
Parathyroid hormone (PTH)
Relationship to Bone Turnover Monitor serum levels of calcium, phosphorus, PTH, and alkaline
phosphatase activity beginning in CKD stage 3
In bone biopsies from patients with CKD, the bone
formation rate is lower in those with arterial calcifica- In patients with CKD stages 3–5 not on dialysis, correct high PTH
levels with any or all of the following: reducing dietary phosphate
tions than in those without [167–169]. Both suppressed intake and administering phosphate binders, calcium supplements,
bone formation and excess bone resorption and forma- and/or native vitamin D
tion are seen with coronary artery calcification [170]. In
In patients with CKD stage 5D, maintain iPTH levels in the range of
stage 1 to 4 CKD the biochemical markers of bone forma-
approximately two to nine times the upper normal limit for the assay
tion are inversely correlated with the arterial wall stiff-
ness, whereas markers of bone resorption are directly In patients with CKD stage 5D and elevated or rising PTH, use
calcitriol, or vitamin D analogs, or calcimimetics, or a combination of
correlated [171].
calcimimetics and calcitriol or vitamin D analogs to lower PTH
Underlying mechanisms include excess bone resorp-
tion releasing calcium and phosphate from the bones, In patients with CKD stages 3–5D with severe hyperparathyroidism
(HPT) who fail to respond to medical/pharmacological therapy,
and thereby initiating the process of vascular smooth
perform. parathyroidectomy
muscle cell transformation. In contrast, low bone forma-
tion results in mineralizing bone surfaces being unable Phosphate
to buffer the extracellular phosphate concentration. In patients with CKD stages 3–5 maintain serum phosphorus in the
Thus, phosphate loads cannot be deposited into the bone normal range (2C)
and instead increase the serum concentration [172,173]. In patients with CKD stage 5D, lower elevated phosphorus levels
This can lead to a viscous cycle as shown in Figure 58.15, toward the normal range
which is one potential pathway in a network of factors
Use phosphate-binding agents to treat hyperphosphatemia
that regulate mineral metabolism. The phosphate stimu-
lates differentiation of the vascular smooth muscle cells Restrict the dose of calcium-based phosphate binders and/or the
dose of calcitriol or vitamin D analog in the presence of persistent or
into osteoblasts. This is also promoted by wnt signal-
recurrent hypercalcemia, arterial calcification, adynamic bone disease,
ing, and is inhibited by moderate levels of 1,25D. The or persistently low PTH levels
­osteoblasts then form bone inside the vascular walls.
Limit dietary phosphate intake in the treatment of
Bisphosphonates reduce excess bone resorption,
hyperphosphatemia
but they reduce bone formation and could potentially
exacerbate extraskeletal calcification. Clinical studies Calcium
of bisphosphonates are described later in this chapter. Maintain serum calcium in the normal range
It will be important for future studies to clarify the role
Vitamin D
of the wnt-signaling pathway in the vasculature because
medications which increase the activity of the pathway, Correct vitamin D deficiency using treatment strategies recommended
and which increase intracellular beta-catenin, are being for the general population
developed for treatment of osteoporosis.
dialysis, and medications. Although we would expect
the bones to become stronger when these abnormali-
ties are addressed, no clinical trials have had sufficient
TREATMENT
power to prove that fractures can be prevented. Since
the KDIGO report, several large randomized trials have
Correct Classic Abnormalities shown disappointing results when evaluating clinical
The first therapeutic steps in a patient with fractures outcomes.
or bone density in the “osteoporotic” range (hip T-score Cholecalciferol is frequently lower than 20 ng/mL (50
lower than –2.5) are to correct the known associated nmol/L) in patients with CKD, but clinical trials have
problems with the mineral and bone disorder seen in not shown whether supplementation is beneficial [174].
CKD-MBD. It is difficult but important to treat abnor- One study in predominantly African-American patients
malities in serum calcium, phosphate, vitamin D, and showed that 200,000 international units (IU) of chole-
PTH simultaneously. The 2009 KDIGO guidelines have calciferol given weekly for 3 weeks increased the serum
extensively reviewed the evidence about medications levels, but there was no change in the PTH levels [175].
that can help achieve this [4] (Table 58.3). Many patients Data from the large international Dialysis Outcomes and
with fractures have abnormalities that can be improved Practice Patterns Study found that the positive associa-
with greater attention to phosphate control using diet, tions between 25(OH)D levels and clinical mortality were

VII.  SPECTRUM OF SECONDARY OSTEOPOROSIS


1406 58.  OSTEOPOROSIS ASSOCIATED WITH CHRONIC KIDNEY DISEASE

not significant when adjusted for comorbidities. This individualized depending on the serum calcium, phos-
suggests that patients who were prescribed vitamin D phate, PTH, and presence of arterial calcifications, and
were healthier [176]. The KDIGO guidelines also sug- the bone turnover rate [182]. Calcium citrate should be
gested that cholecalciferol should be given according to avoided because it increases absorption of aluminum,
recommendations for the general population, because and patients with CKD are at increased risk of ­aluminum
there were not enough studies to provide evidence that toxicity [183].
supplementation should be different in patients with Treatment with the calcimimetic cinacalcet, reduces
CKD [4]. parathyroid gland size by about 50% in hemodialysis
There have been many studies of calcitriol or analogs patients [184]. Early clinical studies that were not large
in patients with CKD, but most were not large enough to enough to determine patient-level outcomes had posi-
determine patient-oriented outcomes. A meta-analysis of tive results that were not found in a large randomized
controlled trials of active vitamin D compounds showed clinical trial. First a randomized trial with 741 hemodi-
they did not consistently reduce PTH levels or have ben- alysis patients found reduced PTH levels after 26 weeks
eficial effects on patient-level outcomes. Hypercalcemia of treatment [185]. Then a cohort study of 19,186 patients
and hyperphosphatemia were frequent complications found lower cardiovascular and all-cause mortality in
[177]. In a large, multinational randomized controlled patients who had been treated with cinacalcet [186]. A
clinical trial that enrolled 227 patients with CKD 3 or pooled analysis of four trials suggested a reduced risk
4, paricalcitol did not reduce left ventricular volume of cardiovascular events and fracture [187]. Finally, the
compared with placebo. However, the group treated EValuation Of Cinacalcet HCl Therapy to Lower Cardio-
with paracalcitol experienced more frequent episodes of Vascular Events (EVOLVE) investigators enrolled 3883
hypercalcemia [178]. The appropriate timing, types, and hemodialysis patients in a randomized clinical trial of
doses of vitamin D, calcitriol, and active analogs in this cinacalcet or placebo, with a mean study duration of 21
population is still uncertain. However, in patients with months in the cinacalcet group. There was a 12% reduc-
CKD stage 3 to 5 with progressively rising PTH levels tion in the composite end point of death or cardiovascu-
despite correction of modifiable factors, treatment with lar event, but this was not statistically significant. There
calcitriol or activated vitamin D analogs is suggested by was no reduction in fracture rates with cinacalcet [160].
the KDIGO guidelines [4]. For patients with CKD stage In patients with earlier stages of CKD, cinacalcet also
5D, the benefits of treatment with both cholecalciferol reduces PTH but the studies did not evaluate cardio-
and physiological levels of calcitriol or an active analog vascular end points. In a large clinical trial patients had
remain unproven. Nevertheless, this approach appears frequent episodes of hypocalcemia, and increased serum
logical, providing clinicians carefully monitor levels of phosphate [188], likely to be related to a reduction in
serum calcium and phosphate. PTH-related phosphaturia.
When should clinicians start using phosphate bind-
ers? In earlier stages of CKD, a historical cohort study Life-Style
suggested that phosphate binders were associated with
lower mortality [179]. However, a pilot study in patients Nutrition
with CKD stages 3 to 5 produced unexpected results. Nutrition in the dialysis patient is challenging. The
Patients were randomized to placebo, calcium acetate, standard nutritional recommendations for patients with
lanthanum, or sevelamer. After 9 months the serum phos- ordinary osteoporosis do not apply to patients with
phate levels were decreased with each of the phosphate CKD-MBD, who must avoid foods with excess phos-
binders, but calcification of the coronary arteries and phate. And while protein deficiency can be harmful and
abdominal aortas was increased [180]. Although there cause malnutrition, extra protein contributes to acidosis,
is abundant evidence that high phosphate is associated hyperphosphatemia, uremia, and progression of renal
with cardiovascular disease, lowering the phosphate failure. Thus, high biologic value protein should be used.
using binders may not prevent disease progression. Consumption of dairy products is often limited
In CKD stage 5D phosphate must be controlled to because they contain high amounts of phosphate. Many
avoid the problems identified with the first dialysis patients are prescribed calcium-based phosphate bind-
patients; these were severe enough to make it unethical ers, which also act as calcium supplements. Calcium
to do a placebo-controlled trial [3]. To date, none of the citrate increases absorption of aluminum and should be
phosphate binders are clearly better than others in terms avoided in CKD stages 4 to 5D.
of patient outcomes such as mortality or cardiovascular Fruits and vegetables, which appear to be associated
adverse events [181]. with better bone density in the general population, con-
Calcium supplements are routinely prescribed for tain high levels of potassium, which can be dangerous to
patients with postmenopausal osteoporosis, but in patients as CKD becomes more advanced, so these foods
patients with CKD-MBD calcium intake must be must often be limited.

VII.  SPECTRUM OF SECONDARY OSTEOPOROSIS


Treatment 1407
The Institute of Medicine recommends 600 IU/day of were elderly, some of the patients did have CKD stage 3
vitamin D for adults younger than 70 years and 800 IU/ or 4. These patients respond to osteoporosis medications
day for older adults in the general population. There are and can be treated the same as other elderly people with
no studies specifically in kidney patients that suggest the osteoporosis. The KDIGO guidelines [4] recommend
dose should be different in those with CKD. that patients with early stages of CKD, who have normal
laboratory values and low bone density or high risk of
Exercise fracture, should be treated to prevent fractures as in the
Bones respond to mechanical forces, and many patients general population.
with CKD-stage 4 or 5 do not get much weight-bearing Patients with stage 4 or 5 CKD, who are developing
exercise. Walking, stair climbing, and dancing, as well as CKD-MBD, may have low bone density, poor bone qual-
back extension exercises, should be encouraged in these ity, and fractures that appear similar to those patients
patients. Exercise programs have not been shown to with ordinary postmenopausal osteoporosis, but the
reduce fractures in large trials, but they make a lot of sense physiological differences must be considered before
and can also improve cardiovascular and mental health. starting medications [190]. Unfortunately, there are still
no large randomized trials of osteoporosis medications
Fall Prevention in stage 4 or 5 CKD patients. The following review sum-
Patients with chronic diseases are at higher risk for marizes available evidence in these late stage patients.
falls [6,117,189], and fall prevention strategies are similar
in those with CKD as in those with ordinary osteoporosis. Estrogen
Many women with CKD have gonadal dysfunction and
thus it may seem logical to treat with estrogen. In post-
Medications used for Osteoporosis menopausal women from the general population, estro-
Osteoporosis is a disease of reduced bone mass and gen replacement therapy has been conclusively shown to
reduced bone strength. There is overlap between “ordi- reduce the incidence of hip, vertebral, and nonvertebral
nary” osteoporosis and CKD-MBD, but they are different fractures [191]. However, the administration of combined
diseases—just as hemolytic anemia and iron-deficiency estrogen–progestin may also increase the risk of breast
anemia have similar manifestations but different physi- cancer, thromboembolic events, and coronary and cerebro-
ology (Table 58.4). vascular disease, with risks dependent on age and years
Although clinical trials have included thousands of since menopause [192]. A current theory is that estrogen
patients with osteoporosis, a small minority of the sub- can help prevent coronary artery calcifications if given to
jects had CKD-MBD. The pivotal trials were designed to women who have normal coronary arteries, but can cause
exclude those with renal disease. Subjects had normal plaque rupture and myocardial infarctions in women who
serum PTH, alkaline phosphatase, calcium, and phos- already have coronary artery disease (CAD) [193]. Because
phate and thus did not have the biochemical abnormali- older women with CKD frequently have CAD, estrogen
ties that herald the advanced mineral and bone disorders should not be initiated in late postmenopausal women.
associated with CKD-MBD. Because they were excluded Estrogen treatment increases BMD in premenopausal
based on creatinine, and not estimated GFR, and they women with CKD. Amenorrheic women receiving dialy-
sis treated with transdermal estradiol for 1 year had bet-
TABLE 58.4  Differences between CKD-MBD and ter change in BMD than those not treated [194]. Because
Postmenopausal Osteoporosis of abnormalities in protein binding, the free estradiol
CKD-MBD Postmenopausal osteoporosis levels in CKD stage 5D women are twice as high as in
women with normal kidney function when given the
Increased PTH and alkaline Normal PTH and alkaline
phosphatase phosphatase same dose [80], so if estrogen is used in this population
the dose should be lowered.
BMD weakly associated with BMD predicts fractures
fractures Testosterone
Bone loss mostly in cortical bone Bone loss in cancellous and Similarly, men with advanced CKD may also have
cortical bone
reduced testosterone levels [195,196], which also may
High prevalence of adynamic or Bone formation generally contribute to abnormal bone. However, there are no
very high bone formation rates normal to slightly high studies that have specifically evaluated the effect of
Abnormal calcium, phosphate, Laboratory results normal ­testosterone therapy on bone in CKD patients.
FGF23, BMP7, Klotho, 1,25D, iron,
bicarbonate, and various cytokines Bisphosphonates
1,25D: 1,25-dihydroxyvitamin D; BMD: bone mineral density; PTH: parathyroid Bisphosphonates are effective medications that
hormone. are widely used to prevent fractures in patients with

VII.  SPECTRUM OF SECONDARY OSTEOPOROSIS


1408 58.  OSTEOPOROSIS ASSOCIATED WITH CHRONIC KIDNEY DISEASE

“ordinary” postmenopausal osteoporosis, in whom the calcifications. One important and unexplored ques-
fracture rates are decreased by about 50%. These medi- tion is whether there are any adverse effects related to
cations decrease both the bone resorption rate and bone ­bisphosphonate deposition in these nonskeletal sites.
formation rate by 70% to 90%. They increase the bone Etidronate reduces vascular calcifications in ani-
density, but not the bone volume, and it is likely that they mal studies and in some case reports [205]. This first-
prevent fractures because they prevent the microarchi- generation bisphosphonate inhibits mineralization because
tectural deterioration that occurs with untreated osteo- it is an analog of pyrophosphate. The newer bisphos-
porosis. Women with the highest tertile of baseline bone phonates are given at substantially lower doses which
turnover markers show significant reduction in fracture do not inhibit mineralization. In women with ordinary
rates with alendronate, but there is no difference in frac- osteoporosis radiographic imaging studies of coronary
ture rate in those with baseline low markers of bone arteries have shown increases in calcifications with both
turnover [197]. The safety and efficacy in patients with ibandronate and alendronate that were similar to the
CKD-MBD remain uncertain [198]. ­Bisphosphonates increases seen in control subjects [206,207]. In a random-
decrease serum calcium and increase PTH. ized, placebo-controlled clinical trial using alendronate
in 51 patients with CKD stage 3 to 5, alendronate did
PHARMACOKINETICS not decrease the progression of vascular calcification
Several features about bisphosphonate actions and ­compared with placebo [208].
pharmacokinetics are important in the context of CKD. A retrospective cohort study included 9604 women
They bind very tightly to mineral, with a half-life of over with CKD stages 3 or 4 who did not have a previ-
10 years [199]. In patients with normal kidney function, ous serious cardiovascular event. They were followed
about half of the administered dose is bound to the bone for 3.9 years to ascertain mortality and rate of serious
and the rest excreted within several hours by the kidney, cardiac events. In the 3234 women who had been
so most of the tissues have only brief exposure to the exposed to bisphosphonates the all-cause mortality
drugs [199]. They are cleared by hemodialysis [200]. was significantly better than in the patients who had
not taken bisphosphonates (adjusted hazard ratio 0.78).
ACUTE RENAL FAILURE However, there was no benefit for cardiovascular events
Acute renal failure has been reported with intravenous (hazard ratio 1.14). Of note, in those patients who ini-
bisphosphonates, especially when given with rapid infu- tially had some cardiac disease, the hazard ratios were
sion and frequent doses to patients with malignancy [201]. 1.20 for death and 1.24 for a serious cardiovascular end
Annual infusions given to osteoporotic patients resulted point [209].
in modest changes in creatinine in 0.4% of placebo- A theoretical concern about bisphosphonate toxicity
treated patients and 1.3% of patients treated with zole- in CKD patients relates to the low bone turnover which
dronic acid [202]. In older patients with a hip fracture, an is a direct effect of the drugs. This could increase vascular
increase in serum creatinine of greater than 0.5 mg/dL calcifications (see Fig. 58.17). In contrast, high turnover
was seen in a similar percentage of patients in both pla- can also increase serum phosphate by directly releasing
cebo (5.6%) and zoledronic acid (6.2%) groups [203]. Spo- it from the bone, and in this case the bisphosphonates
radic cases of acute kidney diseases have been reported, could be helpful.
and the kidney biopsies show tubular injury with low
Na+, K+-ATPase expression [204]. The mechanism of CLINICAL TRIALS IN CHRONIC KIDNEY DISEASE
toxicity is uncertain but may relate to known proper- STAGE 3
ties of the drugs. The proximal tubular cells internalize Miller et al. [14] reported a pooled analysis of nine tri-
bisphosphonates, as do the osteoclasts, and inhibition of als using risedronate for treatment of osteoporosis. The
farnestyl pyrophosphate synthase results in inhibition primary trials were designed to exclude patients with
of prenylation of small GTPases which are necessary for significant systemic disease, so subjects with serum cre-
maintenance of the cytoskeleton and r­ uffled border. This atinine greater than 1.1 times the upper limit of normal
may lead to renal damage. were excluded. There were 4643 women with an eGFR
below 60 mL/minute, most of whom had stage 3 CKD.
CARDIOVASCULAR CALCIFICATIONS These women did not demonstrate the classical bone
Other safety issues must be considered in CKD abnormalities seen in later stages of kidney failure. They
patients. Soft tissue calcifications can also bind bisphos- were excluded if serum PTH or alkaline phosphatase
phonates, as seen with bone scanning agents containing values were higher than normal. These patients showed
bisphosphonates. In patients with ordinary osteopo- reduction in vertebral fracture rates and improvements
rosis, the nonskeletal tissues have short and low expo- in bone density that were similar to those with normal
sure to the drugs. In patients with late stages of CKD renal function, except that in those with the highest cre-
the bisphosphonates may concentrate within soft-tissue atinine, the femoral neck bone density did not respond

VII.  SPECTRUM OF SECONDARY OSTEOPOROSIS


Treatment 1409

Vascular smooth 1,25 (OH)2 vitamin D


muscle cells HO moderate levels

Osteoblasts

wnt
High PO4
sFRP
1,25 (OH)2 vitamin D
high levels
wnt
H Low bone turnover

HO

Bisphosphonate

FIGURE 58.17  Viscous cycle of bone inhibition and vascular calcifications. Vascular smooth muscle cells differentiate into osteoblasts under
the influence of vitamin D and Wnt signaling. The osteoblasts secrete frizzled-related protein (sFRP), which blocks the Wnt signaling but also
causes low bone formation. Bisphosphonates could exacerbate this. The low bone formation leads to increased serum phosphate, which stimulates
the changes in the smooth muscle cells. Source: from Ott (2012) [198], with permission.

to risedronate. An important limitation of this study is after increasing the calcitriol dose the PTH returned to
that nonvertebral fracture rates were not mentioned. values similar to baseline. There were no controls or bone
A similar posthoc analysis of an alendronate trial density measurements at cortical sites [212].
was reported by Jamal et al. [15]. In that study as well, The largest randomized trial in later stages of CKD
the intent of the original trial was to exclude women enrolled 51 patients with CKD stage 3 or 4, and treated
with renal disease, but because of their age, many sub- with alendronate 70 mg/week or placebo for 18 months.
jects did have mild to moderate decreases in the eGFR. The lumbar spine BMD T-score was normal at baseline
Data extrapolated from a figure in the paper shows that (+0.40) and with treatment increased to +0.7. The femo-
fewer than 20 subjects had CKD stage 4, and those with ral neck T-score at baseline was –1.27 but the change in
abnormal serum calcium, PTH or alkaline phosphatase femoral BMD was similar between groups. There was a
values were excluded. This makes it unlikely that many significant increase in PTH levels and decrease in alkaline
patients had CKD-MBD. The authors found that the phosphatase with alendronate compared to placebo [208].
women with eGFR less than 45 mL/minute/1.73 m2 had In patients without known renal disease the bisphos-
similar improvements in bone density and decreases in phonates lead to decreased bone formation. This also
relative fracture risk to those with higher eGFR. occurs in patients with CKD. Bone biopsies from 13 CKD
patients (stages 2–4) who were referred to a nephrology
USE IN LATER STAGES OF CHRONIC KIDNEY DISEASE clinic and who had taken bisphosphonates all demon-
A report of 12 dialysis patients given a single dose strated adynamic bone [213].
of pamidronate found reduced serum calcium and There is no physiological rationale to treat patients
increased PTH [210]. who have low bone turnover with medications whose
Pamidronate every 2 months for 1 year was given to mechanism of action is to inhibit bone resorption. How-
13 hemodialysis patients with high PTH in an open-label ever, the bisphosphonates could potentially be use-
observational study. The PTH initially increased and then ful in patients with low bone density and increased
decreased when the calcitriol dose was increased. The bone resorption. Mitsopoulos and colleagues therefore
bone density increased [211]. A similar study used iban- designed a pilot study to tailor the medicine to the type
dronate for 1 year in 16 dialysis patients who had increased of ROD and value of iPTH. They measured bone den-
PTH and low bone density. Eleven of them completed the sity in 102 hemodialysis patients and 66 of them had
study, and they showed a 5% increase in spinal bone den- osteoporosis. Of these, 38 consented to a bone biopsy.
sity. They also found that the PTH increased at first, but The biopsy revealed high turnover in 22, and in 11 of

VII.  SPECTRUM OF SECONDARY OSTEOPOROSIS


1410 58.  OSTEOPOROSIS ASSOCIATED WITH CHRONIC KIDNEY DISEASE

these the iPTH was less than 300 pg/mL. This group was the hip. Levels of serum pyridinoline (a marker of bone
treated with ibandronate, and 1 year later the bone den- resorption) and of LDL cholesterol decreased after 6
sity had decreased by 3% at the spine and by 2.2% at the months in the raloxifene-treated patients, compared to
hip. The other 11 patients with high turnover had iPTH placebo. There were no side effects noted.
greater than 300 pg/mL and were treated with cina- A prospective study in 17 postmenopausal Japanese
calcet, and their bone density decreased by 7.7% at the women on dialysis found an increase in spine bone den-
spine and 4.2% at the hip. Five patients whose biopsies sity after 1 year, compared with a decrease in nonran-
showed only mild lesions were not treated and they lost domized controls. At the radius, the loss was less than
3% at the hip without change at the spine [214]. in controls. The PTH concentrations increased with ral-
In summary, there is no definite evidence that bisphos- oxifene, and the authors suggested that calcium and/or
phonates are helpful to patients with CKD stages 4 or vitamin D should be added to treatment [217].
5. They are not as effective at improving bone density In another study from Japan, 22 women were treated
as they are in patients with normal kidneys. Long-term with raloxifene and 23 similar women were controls who
effects of low bone formation in these patients pose a chose not to be on the medication. In those with PTH lev-
theoretical concern but clinical evidence is scarce. els lower than 250 pg/mL, the bone density was better
with raloxifene than control. Abdominal aortic calcifica-
Raloxifene tions increased to a similar extent in those taking or not
On a physiologic level, raloxifene is expected to be taking the raloxifene [215].
beneficial to the bone in postmenopausal women with
CKD-MBD because it acts as an agonist to the estrogen Teriparatide
receptors in the skeleton. In women with normal renal It seems counterintuitive to administer the active part
function there is approximately at 50% reduction in the of PTH to patients in whom hyperparathyroidism is a
incidence of breast cancer, which is an important addi- prominent part of the skeletal pathology. However, daily
tional benefit. Raloxifene increases the risk of thrombo- subcutaneous injections are strongly anabolic in women
embolism, similar to estrogen. The drug is excreted via with postmenopausal and glucocorticoid-associated
hepatic metabolism, and drug concentrations are about osteoporosis.
1.4 times higher in patients with CKD [215].
PATIENTS WITH STAGE 3 CHRONIC KIDNEY DISEASE
CLINICAL TRIAL IN CHRONIC KIDNEY DISEASE Miller et al. [13] reported a post hoc analysis which
STAGE 3 utilized data from a clinical trial that excluded subjects
A posthoc study utilized data from the MORE trial with a serum creatinine greater than 2 mg/dL. The study
(Multiple Outcomes of Raloxifene Evaluation) to evalu- excluded individuals with elevations in serum calcium,
ate the efficacy of raloxifene in patients with reduced phosphorus, or PTH, or with vitamin D deficiency. The
kidney function [17]. The original trial included 7705 number of patients was small enough that none had non-
postmenopausal women aged 31 to 80 years old. Women vertebral fractures during the study period. However, the
were randomly assigned to receive placebo, raloxifene vertebral fracture incidence was reduced in those using
60 mg/day or raloxifene 120 mg/day in addition to teriparatide. In addition, teriparatide improved lumbar
daily calcium and vitamin D. Importantly, the study spine BMD, femoral neck BMD and PINP similarly in
excluded individuals with elevations in serum PTH, or patients with normal, mild, and moderately impaired
with vitamin D deficiency. The femoral neck and spine GFR. The treatment increased serum calcium and uric
BMD increased with raloxifene compared to placebo. acid in all subgroups, but the percentage of patients with
The odds ratio for vertebral fracture was 0.60 for those hypercalcemia and hyperuricemia was greater in the
with normal kidney function, 0.54 with eGFR 45–59 moderately impaired GFR group.
mL/minute/1.73 m2, and 0.74 if eGFR was less than
45 mL/minute/1.73 m2. There was no difference in the PATIENTS WITH LATER STAGES OF CHRONIC KIDNEY
nonvertebral fracture incidence. The adverse events
­ DISEASE
were not increased in raloxifene compared with placebo. A pilot study in seven hemodialysis patients with low
bone density, low serum PTH, and bone biopsy evidence
CLINICAL TRIALS IN LATER STAGES OF CHRONIC of adynamic bone disease found that 6 months of terip-
KIDNEY DISEASE aratide increased the bone density of the spine from a
Hernandez and colleagues [216] conducted a random- mean of 0.88 ± 0.08 to 0.91 ± 0.09 g/cm (p < 0.02). The
ized trial in 50 dialysis patients who were at least 2 years femoral neck increased from 0.66 to 0.71, which was not
postmenopausal with a bone density T-score below –2.0. significant [218].
After 1 year, raloxifene-treated patients had a significant Mitsopoulos et al. treated nine patients with biopsy-
improvement in BMD at the lumbar spine, but not at confirmed adynamic bone disease with teriparatide for

VII.  SPECTRUM OF SECONDARY OSTEOPOROSIS


Hypercalciuria 1411
a mean of 16 months. Serum calcium and phosphate Summary
were unchanged, alkaline phosphatase and serum PTH
increased. Bone density of the spine had decreased by In patients with early stages of CKD, in whom kidney
5.4% the year prior to treatment, but increased 4.9% at disease has not been recognized and whose PTH and
the spine and 2.7% at the hip, which, given the limited alkaline phosphatase are normal, the drugs approved
number of participants, did not quite achieve statistical for ordinary osteoporosis are effective and appear safe.
significance [214]. In selected patients with CKD stage 4 to 5, medica-
tions used in the standard management of osteoporosis
Denosumab may be helpful, but they should be used cautiously. In
Denosumab is a new medication that works by inhib- other cases, none of the current osteoporosis medications
iting RANKL, thereby reducing the number of osteo- are appropriate, even when patients are at high risk of
clasts and substantially decreasing bone formation fracture. If we are uncertain about efficacy and safety, we
rates. This monoclonal antibody is not metabolized by should follow the principle of primum non nocere (above
the kidney. all, do no harm). Pilot studies of new medications, such
In a randomized clinical trial of denosumab, 2890 as sclerostin inhibitors, offer hope that in the future there
of 7801 women had an estimated GFR below 59 mL/ will be better choices for these patients. However, any
minute/1.73 m2 of safety and efficacy after 3 years studies must also consider effects on cardiovascular
showed that the subjects with CKD had a similar frac- calcifications.
ture reduction to those with normal renal function, and
they did not have more adverse events. As in the other
studies, patients with hyperparathyroidism were not HYPERCALCIURIA
included. Only 17 had stage 4 CKD and none had stage
5 [16]. A pilot study of patients with later stages of CKD
Pathophysiology
showed a higher incidence of severe hypocalcemia after
an ­injection of denosumab [219]. Hypercalciuria can be extrinsic to the kidney when
there is excess filtered calcium. This does not necessarily
Calcitonin mean there is hypercalcemia, because the kidneys will
Intranasal calcitonin has been considered a safe medi- excrete the excess load to prevent hypercalcemia from
cation that is approved for treatment of osteoporosis. The occurring. Other causes of hypercalciuria are intrin-
BMD does not show large increases on this drug, and the sic to the kidney and can be found with abnormalities
fracture reduction is significant but modest compared to throughout the nephron [224]. This allows an anatomic
other drugs. There is minimal information about using organization of the etiology of hypercalciuria, from the
this antiresorptive medication in patients with stage 4 glomerulus to the collecting tubule (Fig. 58.18). Several
or 5 CKD. One small study from Poland did show that of these causes are genetic (Table 58.5 [225–227]).
control patients lost bone density whereas those taking
Glomerulus: Filtered Load
calcitonin did not. There were no serious side effects or
changes in the phosphate or PTH [220]. Similar results INCREASED CALCIUM LOAD
were seen in a study of 148 Chinese patients who The kidney filters about 8000 mg of calcium a day.
received subcutaneous calcitonin three times a week The filtered load depends on the serum ionized calcium
for 1 year. Their bone density was stable, whereas the and the GFR. The causes of increased calcium load are
­control patients lost bone [221]. similar to the causes of hypercalcemia. Increased bone
One study in dogs given calcitonin for 16 weeks resorption, caused by space flight [228], spinal cord
found protection against bone loss but abnormalities injury, strict bed rest, malignancy, hyperparathyroidism,
in the bone quality [222]. However, bone biopsies from hyperthyroidism, vitamin A toxicity, and glucocorticoid
humans show decreased erosion depth without change use can all cause an increased filtered load.
in the activation frequency, and no problems with Increased intestinal absorption is directly related
­mineralization lag time [223]. to the dietary calcium intake and the vitamin D level.
Recently, the European Union medical regulatory In a clinical trial of calcitriol, hypercalciuria was seen
agency, based on analyses showing that calcitonin is before a noticeable increase in the serum calcium [229].
associated with an increased risk for malignancy, rec- High endogenous levels of 1,25D are also seen in sar-
ommended that calcitonin nasal spray preparations be coidosis, some lymphomas, granulomatous diseases,
removed from the market. At the time of writing, the and hyperparathyroidism. Some studies have found
US Food and Drug Administration (FDA) had not yet an increased incidence of calcium-containing kidney
released any statements or recommendations, and data stones in patients with polymorphisms in the vitamin D
review is still ongoing. receptor which may lead to increased intestinal calcium

VII.  SPECTRUM OF SECONDARY OSTEOPOROSIS


1412 58.  OSTEOPOROSIS ASSOCIATED WITH CHRONIC KIDNEY DISEASE

FIGURE 58.18  Calcium reabsorption through the nephron.

absorption [230]. A newly defined mutation in the INCREASED GLOMERULAR FILTRATION


CYP24A1 gene also causes hypercalcemia, idiopathic Increased delivery of calcium to the renal tubules can
­hypercalciuria and elevated 1,25D [231,232]. also be caused by increased GFR. This can cause mild
It is worth noting here that low dietary calcium can hypercalciuria, and is seen with caffeine, amino acids,
decrease the urine calcium but increase the risk of kid- and pregnancy.
ney stones [233]. This seemingly paradoxical situation
is due to oxalate, which is less soluble when complexed Proximal Tubular Reabsorption
in the gastrointestinal tract to calcium than to sodium. Seventy percent of the filtered calcium is reabsorbed
A low calcium intake, may increase the absorption of in the proximal tubule. This is mainly by passive reab-
sodium oxalate, which must be excreted by the kidney. sorption and is paracellular. Normally the lumen has
The increased urine oxalate may then form calcium oxa- a net positive charge which favors reabsorption. There
late kidney stones. In the past, physicians who did not is also some cellular transport via calcium/hydrogen
understand this instructed patients to decrease their exchangers, which require luminal bicarbonate [236].
calcium intake, which only exacerbated the problem This is one mechanism by which patients with acidosis
of oxalate stones and potentially increased the risk of have decreased calcium reabsorption [237].
osteoporosis. Sodium excess and mannitol increase calcium excre-
Phosphate wasting can indirectly lead to hypercalci- tion by increasing the flow and solvent drag.
uria, because low phosphate stimulates renal produc- Dent disease is an x-linked recessive disease caused
tion of 1,25D. In turn, this increased vitamin D increases by a mutation in the gene for the chloride channel
intestinal calcium absorption and the filtered calcium (CLCN5), which leads to proximal tubular dysfunc-
load. The PTH concentration may be low, which will tion. Patients have hypercalciuria, phosphaturia with
also contribute to hypercalciuria by an effect on tubular hypophosphatemia, normal to low serum PTH, micro-
reabsorption (discussed below). One study reported a globulinemia, rickets, and progressive renal failure.
shift in the distribution of phosphate tubular resorption The mechanism is probably decreased proximal tubular
in hypercalciuric patients with calcium nephrolithiasis, CLCN5-dependent endocytosis of filtered PTH. This
all with normal PTH, and in some patients it was due to increases the luminal PTH and activates PTH receptors
a mutation in the sodium phosphate cotransporter [234]. in the late proximal tubule, leading to increased levels of
However, another study did not find a similar high inci- 1,25D and phosphaturia. Thiazide diuretics reduce the
dence of ­hypophosphatemia [235]. urine calcium to normal [238].

VII.  SPECTRUM OF SECONDARY OSTEOPOROSIS


Hypercalciuria 1413
TABLE 58.5 Some Genetic Causes of Hypercalciuria
Gene Name Location Mechanism

CYP24A1 Vitamin D 24-hydroxylase Filtered load Mutated enzyme in proximal tubule causes
accumulation of 1,25D

SLC34A3 Sodium-phosphate cotransporter Filtered load Mutated protein in the proximal tubule causes
(hereditary hypophosphatemic rickets phosphate loss which stimulates 1,25D and
with hypercalciuria) increases gastrointestinal calcium absorption

CLCN5 Chloride channel (Dent disease) Proximal tubule Abnormal proximal transport, excess PTH in
urine leads to calcium loss

ROMK1, NKCC2, Potassium channel or sodium chloride Thick ascending Mutations block potassium recycling which
ClC-Kb cotransporter or chloride channel (Bartter limb causes loss of voltage gradient
syndrome type I, II, III)

CASR Calcium-sensing receptor (autosomal- Thick ascending Activating mutation blocks potassium
dominant hypocalcemia with limb recycling which causes loss of voltage gradient
hypercalciuria)

Claudin-16 Familial hypomagnesemia Thick ascending Tight junction do not function properly, so
limb voltage dissipates

WNK-4 Pseudohypoaldosteronsim Distal tubule Loss of inhibition of sodium chloride


cotransporter which results in closed calcium
channels

ATP6B1, SLC4A1, Distal renal tubular acidosis Distal tubule Acidosis blocks calcium channels
SLC4A1

1,25D: 1,25-dihydroxyvitamin D; PTH: parathyroid hormone.

Reabsorption from the Loop hypercalcemia and elevated PTH, which is “benign” in
There is no calcium transport in the descending loop heterozygotes (familial hypocalciuric hypercalcemia).
of Henle, but the thick ascending loop is an important Parathyroidectomy should be avoided in heterozygotes.
segment which reabsorbs 20% of the calcium. This Homozygotes, however, have severe neonatal hyperpara-
is also passive and paracellular and is driven by the thyroidism and need both parathyroidectomy and loop
lumen-­positive voltage, which depends on the Na/K/Cl diuretics. Rarely, these diseases are seen in patients who
triporter and associated potassium channel. This develop autoantibodies to the calcium sensor, which can
­segment also has abundant CaRs [239,240]. be either activating [243] or inactivating [244].

LOOP DIURETICS MUTATIONS IN CLAUDIN-16


Loop diuretics increase calcium excretion by inhibi- Claudin-16 (formerly called paracellin) is found in
tion of the Na/K/Cl triporter. This triporter creates a the tight junctions. It allows sodium to pass through the
transepithelial voltage gradient that is positive in the junction, but not chloride. This segment of the nephron
lumen, which drives paracellular calcium reabsorption. is impermeable to water, so the sodium concentration
in the tubule in much lower than in the extracellular
ACTIVATING MUTATIONS OF CaRs space; therefore, sodium will move down the concen-
The CaRs, when activated by extracellular calcium, tration gradient, leaving chloride behind, and keeps the
signal intracellular inhibition of the potassium channels. transepithelial voltage positive on the lumen side. With
This inhibition of potassium recycling will inhibit the homozygous mutations, the voltage dissipates and both
ability of the Na/K/Cl triporter to maintain the voltage calcium and magnesium remain inside the tubule. Famil-
gradient, and therefore calcium remains in the urine [239]. ial hypomagnesemia with hypercalciuria is an autoso-
Patients can have either activating or inactivating muta- mal-recessive disease, presents in children with urinary
tions of the CaR, with resulting mineral abnormalities that tract infections, polyuria, hematuria and magnesium
are mirror images of each other. With activating mutations and calcium wasting [245]. They develop nephrocalcino-
(autosomal-dominant hypocalcemia with hypercalciuria), sis and renal failure, may have rickets, ocular abnormali-
they have hypercalciuria, hypocalcemia, and suppressed ties and hearing impairment. Heterozygotes may have
PTH. Overtreatment with calcium or vitamin D leads to hypercalciuria. Treatment with thiazide partially helps
nephrocalcinosis [241], and thiazides are beneficial [242]. the hypercalciuria, but definitive cure requires renal
In contrast, inactivating mutations cause hypocalciuria, transplantation [246].

VII.  SPECTRUM OF SECONDARY OSTEOPOROSIS


1414 58.  OSTEOPOROSIS ASSOCIATED WITH CHRONIC KIDNEY DISEASE

Distal Tubular Reabsorption and urine. Klotho hydrolyzes sugar residues on TRPV5,
The distal nephron provides the fine regulation of which traps this calcium channel in the membrane,
electrolyte excretion or reabsorption [247]. About 5% prolongs its activity and increases calcium reabsorp-
to 10% of the filtered load is reabsorbed, and this is an tion [252]. Klotho also contributes to the inhibition of
active process that is transcellular, against both electrical 1-alpha hydroxylase and the synthesis of 1,25D. Mice
and concentration gradients. Epithelial calcium channels with Klotho knockout have hypercalciuria. A patient
on the luminal side (TRPV5; transient receptor potential with a missense mutation in the Klotho gene developed
cation channel vanilloid 5) are induced by 1,25D and tumoral calcinosis with high serum calcium, high 1,25D,
activated by membrane hyperpolarization agents such and hypercalciuria [253], but in other people with klotho
as PTH, amiloride, and calcitonin. These channels are gene mutations, the urine calcium has not been elevated
inhibited by acidosis. The calcium is carried through [254,255].
the cell by calbinden (which is transcribed when vita-
PSEUDOHYPOALDOSTERONISM II
min D receptors are activated), and then exported from
the basolateral side by magnesium-dependent calcium- Pseudohypoaldosteronism type II is caused by
ATPase with cotransport of H+ or by a Na/Ca exchanger mutations in WNK-4 [256,257]. This is an autosomal-
which can be stimulated by PTH [248]. dominant disease. WNK-4 normally inhibits the Na/
On the luminal cell membrane, the sodium-chloride Cl cotransporter, so when it is mutated the transporter
cotransporter regulates membrane polarization. When this is unregulated and patients develop hypercalciuria,
transporter is inhibited, the membrane becomes hyperpo- hypertension, hypermagnesemia, hyperkalemia, acido-
larized and the voltage-sensitive calcium channels (TRPV5) sis, and low bone mass. Thiazide diuretics correct all the
are opened. This transporter is sensitive to thiazide diuret- abnormalities.
ics. The cell membranes are also hyperpolarized with The “mirror image” disease is an inactivating mutation
amiloride, which also reduces urine calcium loss [249]. of the Na/Cl cotransporter (Gitelman’s syndrome) which
results in hypocalciuria, hypotension, hypomagnesemia,
HYPOPARATHYROIDISM hypokalemia, alkalosis, and high bone mass. These find-
Both hyperparathyroidism and hypoparathyroidism ings could also be seen with overdoses of t­ hiazide diuretics.
cause hypercalciuria, but the mechanisms are different.
Collecting Duct
In hyperparathyroidism the serum calcium is elevated
and this causes the high urine calcium. In hypoparathy- There is not much calcium transport, but this seg-
roidism there is increased tubular reabsorption of cal- ment plays an important role in regulating the concen-
cium. If serum calcium is markedly low then patients tration of the urine, and CaRs are located here [239].
will not have hypercalciuria. However, if well-meaning When activated by calcium, they block the message
health care providers treat with calcium and vitamin D from anti-diuretic hormone (on the basolateral side) and
in an attempt to “normalize” the serum calcium, then inhibit aquaporin (on the luminal side) which makes the
serious hypercalciuria or nephrocalcinosis may develop urine more dilute. This mechanism is one of the body’s
[250]. These patients should be maintained at a serum defenses against kidney stones. It also contributes to
calcium slightly below the normal range. hypernatremia in patients with hypercalcemia.
PTH acts to increase calcium passage into the cells by
the epithelial calcium channels. When the PTH receptor Hypercalciuria and Relationship to Osteoporosis
is activated, intracellular signaling causes transcription of
Hypercalciuria as a Factor in Osteoporosis
TRPV5, hyperpolarization of cell membranes which opens
the TRPV5 channels, and inhibition of caveolae-mediated Hypercalciuria is variably defined [258], but most
endocytosis of the TRPV5. In addition PTH receptors are agree that urine calcium should be lower than 300
on the basolateral side of the cells, and when activated mg/day when the dietary intake is within the recom-
they probably activate the nearby Na/Ca exchanger [251]. mended range (1200 mg/day). Fasting and postpran-
dial spot urinary calcium:creatinine ratios may not
ACIDOSIS detect hypercalciuria accurately, so a 24-hour urine is
Hydrogen ions within the lumen will close the TRPV5 the best test [259]. The prevalence of idiopathic hyper-
channels and reduce calcium reabsorption [237]. Sev- calciuria in patients with o
­ steoporosis varies from 10%
eral hereditary forms of distal renal tubular acidosis are to 20% [258,260–262].
­associated with hypercalciuria [227].
Osteoporosis in Stone Formers
KLOTHO DEFICIENCY Not only is hypercalciuria common in patients with
Klotho is expressed in the distal tubules, and the osteoporosis, but patients with hypercalciuria manifest
extracellular portion of the molecule is shed into blood decreased bone density [224,263,264]. BMD is about one

VII.  SPECTRUM OF SECONDARY OSTEOPOROSIS


Hypercalciuria 1415
standard deviation below expected in patients with idio- Treatment of Bone in Patients with
pathic hypercalciuria [265]. The bone loss is seen par- Hypercalciuria
ticularly when patients also avoid dairy products [266].
Vertebral fracture risk was increased nearly fourfold in Response to Alkali
patients with urolithiasis [267]. The NHANES study Several studies have shown increase in BMD in patients
also showed increased fracture incidence in patients with kidney stones who have been treated with potas-
with kidney stones [266]. An analysis of multiple studies sium citrate [268], which also decreases the urine calcium.
showed bone loss at all skeletal sites [268].
Hydrochlorthiazide
Bone Biopsy Data It is important to remember that different diuretics
Three studies have measured bone biopsies in have opposite effects on urine calcium. Loop diuretics
patients with hypercalciuria. There is decreased bone such as furosimide will cause calcium loss, whereas thia-
volume, increased eroded surfaces without fibrosis, and zides and amiloride will reduce calcium loss. Studies of
decreased bone formation rates. Some mineralization bone density in thiazide users are shown in Table 58.6
defects have been described as well [269–271]. [272]. Most show a positive effect on the bone density.

TABLE 58.6 Studies that Relate Thiazide to Bone Density


Author Year Design n Subjects BMD difference

Cauley 2005 xs 5995 Healthy men Hip 2.4% higher, spine 3.0% higher, not significant in
multivariate analysis

Bakhireva 2004 cohort 507 Elderly men No significant difference in bone loss during 4 years

Tanaka 2001 xs 325 Older men Higher in thiazide

Sigurdsson 2001 xs 248 Spine 9.6% higher, whole body 5.4% higher

LaCroix 2000 RCT 320 Healthy men and Better hip (0.8–0.9% difference) and spine
3 years women

Reid 2000 RCT 185 pm women Better at total body, arms (0.8% between groups),
2 years not spine or hip

Orwoll 1996 xs 7963 Women > 65 years Higher

Glynn 1995 xs 523 Men > 50 years Higher

Wasnich 1995 RCT 119 Hypertensive women Calcaneus, radius increased (+1.5%/year vs. –0.9%/year
2.6 years in placebo)

Morton 1994 xs 1696 Community > 44 years Higher with current use

Jergas 1994 xs 89 Hypertensive Spine BMD 10% higher in men, 19% higher in women

Cauley 1993 xs 9704 Women > 65 years Higher 7–10%, 4–5% adjusted

Bauer 1993 xs 9704 Women > 65 years Higher BMD about 6%

Sowers 1993 cohort 435 Women 55–80 years Loss less (5% vs. 7.4% in 5 years)

Ooms 1993 xs 348 Women > 70 years Thiazide not better

Dawson Hughes 1993 obs 246 pm women Bone loss during winter less in those on thiazide

Wasnich 1990 obs 1017 Japanese men Bone loss less with thiazide

Wasnich 1986 retro 993 Japanese women Higher

Sowers 1985 xs 324 pm women Higher

Adland-Davenport 1985 retro 54 pm women Not different than control

Wasnich 1983 xs 1368 Elderly men SPA 1.157 vs. 1.11 nonusers vs. 1.103 hypertensive nonusers

Transbol 1982 RCT 63 Early pm No loss in SPA for 6 months, then decline similar to placebo;
2 years after 3 years placebo –5.9%, thiazide –4.8%

pm: postmenopause; RCT: randomized controlled trial; SPA: single photon asorptiometry; xs: cross-sectional.
Source: table from Ott, SM. (2012) [272] “Hypercalciuria” Retrieved 12/12/12, from http://courses.washington.edu/bonephys/hypercalU/hypercal.html.

VII.  SPECTRUM OF SECONDARY OSTEOPOROSIS


1416 58.  OSTEOPOROSIS ASSOCIATED WITH CHRONIC KIDNEY DISEASE

FIGURE 58.19  Forest plot of comparison: current thiazide users versus nonusers. Source: from Aung and Htay (2011) [281], with permission.

Hydrochlorthiazide (HCTZ) reduced bone resorption treatment for the bone disease depends on the etiology
in a bone histological study [273]. Pilot studies showed of the renal disease. In idiopathic hypercalciuria, thia-
increase in bone density in patients with hypercalciuria zide diuretics are often helpful for both the kidney and
[274]. A 1-year randomized clinical trial of thiazides in the bone.
24 men with hypercalciuria showed increased of 2.5% of
bone density at the radius in the treatment group com-
References
pared to a loss of 2.1% in the placebo group [275].
Several randomized trials in normal subjects have [1] 
Moe S, Drueke T, Cunningham J, Goodman W, Martin K,
Olgaard K, et al. Definition, evaluation, and classification
shown a modest but significant benefit on BMD [276–
of renal osteodystrophy: a position statement from Kidney
278]. One trial showed that HCTZ lowered urine cal- Disease: Improving Global Outcomes (KDIGO). Kidney Int
cium, but in many people the effect waned, whereas 2006;69(11):1945–53.
there was a consistent mild increase in bicarbonate, [2] 
Zhang QL, Rothenbacher D. Prevalence of chronic kidney dis-
which reverted to normal upon discontinuation of the ease in population-based studies: systematic review. BMC Public
Health 2008;8(1):117.
thiazide. Thus, thiazides also could act by reducing the
[3] 
Pendras JP, Erickson RV. Hemodialysis: a successful therapy for
acid load [279]. chronic uremia. Ann Intern Med 1966;64(2):293–311.
A meta-analysis of fracture risk with thiazide use [4] 
KDIGO. KDIGO clinical practice guideline for the diagno-
found an overall benefit with a relative risk of 0.86 (95% sis, evaluation, prevention, and treatment of Chronic Kidney
CI 0.81 to 0.92) [280]. A Cochrane review of thiazide Disease-Mineral and Bone Disorder (CKD-MBD). Kidney Int
diuretic use and the risk of hip fracture showed a benefit 2009(Suppl. 113):S1–30.
[5] 
Alem AM, Sherrard DJ, Gillen DL, Weiss NS, Beresford SA,
in the large cohort studies, with a summary relative risk Heckbert SR, et al. Increased risk of hip fracture among patients
of 0.76 (95% CI 0.64 to 0.89) (Fig. 58.19) [281]. with end-stage renal disease. Kidney Int 2000;58:396–9.
[6] 
Dukas L, Schacht E, Stahelin HB. In elderly men and women
treated for osteoporosis a low creatinine clearance of <65 ml/
Summary min is a ssf risk factor for falls and fractures. Osteoporos Int
2005;16(12):1683–90.
The renal handling of calcium is complex and abnor- [7] 
Dooley AC, Weiss NS, Kestenbaum B. Increased risk of hip
malities can occur in many steps along the way. These fracture among men with CKD. Am J Kidney Dis 2008;51(1):
often are associated with bone loss or osteomalacia. The 38–44.

VII.  SPECTRUM OF SECONDARY OSTEOPOROSIS


Hypercalciuria 1417
[8] 
Ensrud KE, Lui LY, Taylor BC, Ishani A, Shlipak MG, Stone KL, [25] 
Ishimura E, Okuno S, Yamakawa T, Inaba M, Nishizawa Y.
et al. Renal function and risk of hip and vertebral fractures in Serum magnesium concentration is a significant predictor of
older women. Arch Intern Med 2007;167(2):133–9. mortality in maintenance hemodialysis patients. Magnes Res
[9] 
LaCroix AZ, Lee JS, Wu L, Cauley JA, Shlipak MG, Ott SM, et al. 2007;20(4):237–44.
Cystatin-C, renal function and incidence of hip fracture in post- [26] 
Spiegel DM, Farmer B. Long-term effects of magnesium car-
menopausal women. J Am Soc Geriat 2008;56(8):1434–41. bonate on coronary artery calcification and bone mineral
[10] 
Ensrud KE, Parimi N, Fink HA, Ishani A, Taylor BC, Steffes M, density in hemodialysis patients: a pilot study. Hemodial Int
et al. Renal function and risk of hip fracture in older men. Am J 2009;13(4):453–9.
Kidney Dis 2013; In press. [27] 
Malluche HH, Mawad H, Trueba D, Monier-Faugere MC. Para-
[11] 
Nitsch D, Mylne A, Roderick PJ, Smeeth L, Hubbard R, Fletcher A. thyroid hormone assays—evolution and revolutions in the care
Chronic kidney disease and hip fracture-related mortality in older of dialysis patients. Clin Nephrol 2003;59:313–8.
people in the UK. Nephrol Dial Transplant 2009;24(5):1539–44. [28] 
Naveh-Many T, Rahamimov R, Livn N, Silver J. Parathyroid
[12] 
Klawansky S, Komaroff E, Cavanaugh Jr PF, Mitchell DY, cell proliferation in normal and chronic renal failure rats. The
Gordon MJ, Connelly JE, et al. Relationship between age, renal effects of calcium, phosphate, and vitamin D. J Clin Invest
function and bone mineral density in the US population. Osteo- 1995;96(4):1786–93.
poros Int 2003;14(7):570–6. [29] 
Lewin E, Rodriguez M. Abnormal parathyroid gland function in
[13] 
Miller PD, Schwartz EN, Chen P, Misurski DA, Krege JH. Teripa- CKD. In: Olgaard K, Salusky IB, Silver J, editors. The Spectrum
ratide in postmenopausal women with osteoporosis and mild or of Mineral and Bone Disorders in Chronic Kidney DIsease. 2nd
moderate renal impairment. Osteoporos Int 2007;18(1):59–68. ed. Oxford: Oxford University Press; 2010. p. 77–107.
[14] 
Miller PD, Roux C, Boonen S, Barton IP, Dunlap LE, Burgio DE. [30] 
Silver J, Naveh-Many T. FGF23 and the parathyroid glands.
Safety and efficacy of risedronate in patients with age-related Pediatr Nephrol 2010;25(11):2241–5.
reduced renal function as estimated by the Cockcroft and Gault [31] 
Canaff L, Hendy GN. Human calcium-sensing receptor gene.
method: a pooled analysis of nine clinical trials. J Bone Miner Res Vitamin D response elements in promoters P1 and P2 confer
2005;20(12):2105–15. transcriptional responsiveness to 1,25-dihydroxyvitamin D. J
[15] 
Jamal SA, Bauer DC, Ensrud KE, Cauley JA, Hochberg M, Biol Chem 2002;277(33):30337–50.
Ishani A, et al. Alendronate treatment in women with normal [32] 
Tfelt-Hansen J, Brown EM. The calcium-sensing receptor in nor-
to severely impaired renal function: an analysis of the Fracture mal physiology and pathophysiology: a review. Crit Rev Clin
Intervention Trial. J Bone Miner Res 2007;22(4):503–8. Lab Sci 2005;42(1):35–70.
[16] 
Jamal SA, Ljunggren O, Stehman-Breen C, Cummings SR, [33] 
Dusso A, Arcidiacono MV, Yang J, Tokumoto M. Vitamin D
McClung MR, Goemaere S, et al. Effects of denosumab on frac- inhibition of TACE and prevention of renal osteodystrophy and
ture and bone mineral density by level of kidney function. J Bone cardiovascular mortality. J Steroid Biochem Mol Biol 2010;121(1–
Miner Res 2011;26(8):1829–35. 2):193–8.
[17] 
Ishani A, Blackwell T, Jamal SA, Cummings SR, Ensrud KE. The [34] 
Cozzolino M, Lu Y, Finch J, Slatopolsky E, Dusso AS. p21WAF1
effect of raloxifene treatment in postmenopausal women with and TGF-alpha mediate parathyroid growth arrest by vitamin D
CKD. J Am Soc Nephrol 2008;19(7):1430–8. and high calcium. Kidney Int 2001;60(6):2109–17.
[18] 
Andrukhova O, Zeitz U, Goetz R, Mohammadi M, Lanske B, [35] 
Krajisnik T, Olauson H, Mirza MA, Hellman P, Akerstrom G,
Erben RG. FGF23 acts directly on renal proximal tubules to Westin G, et al. Parathyroid Klotho and FGF-receptor 1 expres-
induce phosphaturia through activation of the ERK1/2-SGK1 sion decline with renal function in hyperparathyroid patients
signaling pathway. Bone 2012;51(3):621–8. with chronic kidney disease and kidney transplant recipients.
[19] 
Weinman EJ, Steplock D, Shenolikar S, Biswas R. Fibroblast Kidney Int 2010;78(10):1024–32.
growth factor-23-mediated inhibition of renal phosphate trans- [36] 
Kalkwarf HJ, Denburg MR, Strife CF, Zemel BS, Foerster DL,
port in mice requires sodium–hydrogen exchanger regulatory Wetzsteon RJ, et al. Vitamin D deficiency is common in chil-
factor-1 (NHERF-1) and synergizes with parathyroid hormone. dren and adolescents with chronic kidney disease. Kidney Int
J Biol Chem 2011;286(43):37216–21. 2012;81(7):690–7.
[20] 
Palmer SC, Hayen A, Macaskill P, Pellegrini F, Craig JC, [37] 
LaClair RE, Hellman RN, Karp SL, Kraus M, Ofner S, Li Q,
Elder GJ, et al. Serum levels of phosphorus, parathyroid hor- et al. Prevalence of calcidiol deficiency in CKD: a cross-sectional
mone, and calcium and risks of death and cardiovascular disease study across latitudes in the United States. Am J Kidney Dis
in individuals with chronic kidney disease: a systematic review 2005;45(6):1026–1033.
and meta-analysis. JAMA 2011;305(11):1119–27. [38] 
Patel S, Barron JL, Mirzazedeh M, Gallagher H, Hyer S,
[21] 
Alfrey AC, Miller NL, Butkus D. Evaluation of body magnesium Cantor T, et al. Changes in bone mineral parameters, vitamin D
stores. J Lab Clin Med 1974;84(2):153–62. metabolites, and PTH measurements with varying chronic kid-
[22] 
Gonella M, Ballanti P, Della Rocca C, Calabrese G, Pratesi G, ney disease stages. J Bone Miner Metab 2011;29(1):71–9.
Vagelli G, et al. Improved bone morphology by normalizing [39] 
Malluche HH, Mawad H, Monier-Faugere MC. Effects of treat-
serum magnesium in chronically hemodialyzed patients. Miner ment of renal osteodystrophy on bone histology. Clin J Am Soc
Electrolyte Metab 1988;14(4):240–5. Nephrol 2008;3(Suppl. 3):S157–63.
[23] 
D’Haese PC, Couttenye MM, Lamberts LV, Elseviers MM, [40] 
Lau WL, Leaf EM, Hu MC, Takeno MM, Kuro OM, Moe OW,
Goodman WG, Schrooten I, et al. Aluminum, iron, lead, cad- et al. Vitamin D receptor agonists increase Klotho and osteopon-
mium, copper, zinc, chromium, magnesium, strontium, and tin while decreasing aortic calcification in mice with chronic kid-
calcium content in bone of end-stage renal failure patients. Clin ney disease fed a high phosphate diet. Kidney Int 2012;82(12):
Chem 1999;45(9):1548–56. 1261–70.
[24] 
de Francisco AL, Leidig M, Covic AC, Ketteler M, Benedyk-­ [41] 
Razzaque MS. The dualistic role of vitamin D in vascular calcifi-
Lorens E, et al. Evaluation of calcium acetate/magnesium cations. Kidney Int 2011;79(7):708–14.
carbonate as a phosphate binder compared with sevelamer [42] 
Liu S, Gupta A, Quarles LD. Emerging role of fibroblast growth
hydrochloride in haemodialysis patients: a controlled random- factor 23 in a bone–kidney axis regulating systemic phosphate
ized study (CALMAG study) assessing efficacy and tolerability. homeostasis and extracellular matrix mineralization. Curr Opin
Nephrol Dial Transplant 2010;25(11):3707–17. Nephrol Hypertens 2007;16(4):329–35.

VII.  SPECTRUM OF SECONDARY OSTEOPOROSIS


1418 58.  OSTEOPOROSIS ASSOCIATED WITH CHRONIC KIDNEY DISEASE

[43] 
Perwad F, Portale AA. Vitamin D metabolism in the kidney: reg- [60] 
Isakova T, Wahl P, Vargas GS, Gutierrez OM, Scialla J, Xie H,
ulation by phosphorus and fibroblast growth factor 23. Mol Cell et al. Fibroblast growth factor 23 is elevated before parathyroid
Endocrinol 2011;347(1–2):17–24. hormone and phosphate in chronic kidney disease. Kidney Int
[44] 
Inoue Y, Segawa H, Kaneko I, Yamanaka S, Kusano K, 2011;79(12):1370–8.
Kawakami E, et al. Role of the vitamin D receptor in FGF23 ac- [61] 
Wesseling-Perry K, Pereira RC, Wang H, Elashoff RM, Sahney
tion on phosphate metabolism. Biochem J 2005;390(Pt 1):325–31. S, Gales B, et al. Relationship between plasma fibroblast growth
[45] 
Martin A, David V, Quarles LD. Regulation and function of the factor-23 concentration and bone mineralization in children with
FGF23/Klotho endocrine pathways. Physiol Rev 2012;92(1): renal failure on peritoneal dialysis. J Clin Endocrinol Metab
131–55. 2009;94(2):511–7.
[46] 
Zisman AL, Wolf M. Recent advances in the rapidly evolving [62] 
van Husen M, Fischer AK, Lehnhardt A, Klaassen I, Moller K,
field of fibroblast growth factor 23 in chronic kidney disease. Muller-Wiefel DE, et al. Fibroblast growth factor 23 and bone
Curr Opin Nephrol Hypertens 2010;19(4):335–42. metabolism in children with chronic kidney disease. Kidney Int
[47] 
Chefetz I, Kohno K, Izumi H, Uitto J, Richard G, Sprecher E. 2010;78(2):200–6.
GALNT3, a gene associated with hyperphosphatemic familial [63] 
Stubbs JR, He N, Idiculla A, Gillihan R, Liu S, David V, et al.
tumoral calcinosis, is transcriptionally regulated by extracellular Longitudinal evaluation of FGF23 changes and mineral metabo-
phosphate and modulates matrix metalloproteinase activity. Bio- lism abnormalities in a mouse model of chronic kidney disease.
chim Biophys Acta 2009;1792(1):61–7. J Bone Miner Res 2011;27(1):38–46.
[48] 
Bhattacharyya N, Chong WH, Gafni RI, Collins MT. Fibroblast [64] 
Shimada T, Urakawa I, Isakova T, Yamazaki Y, Epstein M,
growth factor 23: state of the field and future directions. Trends Wesseling-Perry K, et al. Circulating fibroblast growth factor 23
Endocrinol Metab 2012;23(12):610–8. in patients with end-stage renal disease treated by peritoneal
[49] 
Masuyama R, Stockmans I, Torrekens S, Van Looveren R, dialysis is intact and biologically active. J Clin Endocrinol Metab
Maes C, Carmeliet P, et al. Vitamin D receptor in chondrocytes 2010;95(2):578–85.
promotes osteoclastogenesis and regulates FGF23 production in [65] 
Gutierrez OM, Mannstadt M, Isakova T, Rauh-Hain JA,
osteoblasts. J Clin Invest 2006;116(12):3150–9. Tamez H, Shah A, et al. Fibroblast growth factor 23 and mor-
[50] 
Liu S, Tang W, Zhou J, Stubbs JR, Luo Q, Pi M, et al. Fibroblast tality among patients undergoing hemodialysis. N Engl J Med
growth factor 23 is a counter-regulatory phosphaturic hormone 2008;359(6):584–92.
for vitamin D. J Am Soc Nephrol 2006;17(5):1305–15. [66] 
Kendrick J, Cheung AK, Kaufman JS, Greene T, Roberts WL,
[51] 
Lopez I, Rodriguez-Ortiz ME, Almaden Y, Guerrero F, Smits G, et al. FGF-23 associates with death, cardiovascular
de Oca AM, et al. Direct and indirect effects of parathyroid hor- events, and initiation of chronic dialysis. J Am Soc Nephrol
mone on circulating levels of fibroblast growth factor 23 in vivo. 2011;22(10):1913–22.
Kidney Int 2011;80(5):475–82. [67] 
Wolf M, Molnar MZ, Amaral AP, Czira ME, Rudas A,
[52] 
Rhee Y, Bivi N, Farrow E, Lezcano V, Plotkin LI, White KE, et al. Ujszaszi A, et al. Elevated fibroblast growth factor 23 is a risk
Parathyroid hormone receptor signaling in osteocytes increases factor for kidney transplant loss and mortality. J Am Soc Nephrol
the expression of fibroblast growth factor-23 in vitro and in vivo. 2011;22(5):956–66.
Bone 2011;49(4):636–43. [68] 
Ix JH, Katz R, Kestenbaum BR, de Boer IH, Chonchol M, et al.
[53] 
Koizumi M, Komaba H, Nakanishi S, Fujimori A, Fukagawa M. Fibroblast growth factor-23 and death, heart failure, and cardio-
Cinacalcet treatment and serum FGF23 levels in haemodialysis vascular events in community-living individuals: CHS (Cardio-
patients with secondary hyperparathyroidism. Nephrol Dial vascular Health Study). J Am Coll Cardiol 2012;60(3):200–7.
Transplant 2012;27(2):784–90. [69] 
Shalhoub V, Ward SC, Sun B, Stevens J, Renshaw L, Hawkins
[54] 
Lavi-Moshayoff V, Wasserman G, Meir T, Silver J, Naveh- N, et al. Fibroblast growth factor 23 (FGF23) and alpha-Klotho
Many T. PTH increases FGF23 gene expression and medi- stimulate osteoblastic MC3T3.E1 cell proliferation and inhibit
ates the high-FGF23 levels of experimental kidney failure: a mineralization. Calcif Tissue Int 2011;89(2):140–50.
bone parathyroid feedback loop. Am J Physiol Renal Physiol [70] 
Urena Torres P, Friedlander G, de Vernejoul MC, Silve C, Prie D.
2010;299(4):F882–9. Bone mass does not correlate with the serum fibroblast growth
[55] 
Atkins GJ, Rowe PS, Lim HP, Welldon KJ, Ormsby R, factor 23 in hemodialysis patients. Kidney Int 2008;73(1):102–7.
Wijenayaka AR, et al. Sclerostin is a locally acting regulator of [71] 
Kanda E, Yoshida M, Sasaki S. Applicability of fibroblast growth
late-osteoblast/preosteocyte differentiation and regulates min- factor 23 for evaluation of risk of vertebral fracture and chronic
eralization through a MEPE-ASARM-dependent mechanism. kidney disease–mineral bone disease in elderly chronic kidney
J Bone Miner Res 2011;26(7):1425–36. disease patients. BMC Nephrol 2012;13:122.
[56] 
Imel EA, Peacock M, Gray AK, Padgett LR, Hui SL, Econs MJ. [72] 
Kuro-o M, Matsumura Y, Aizawa H, Kawaguchi H, Suga T,
Iron modifies plasma FGF23 differently in autosomal dominant Utsugi T, et al. Mutation of the mouse Klotho gene leads to a
hypophosphatemic rickets and healthy humans. J Clin Endocri- syndrome resembling ageing. Nature 1997;390(6655):45–51.
nol Metab 2011;96(11):3541–9. [73] 
Kuro-o M. Klotho as a regulator of fibroblast growth factor sig-
[57] 
Manghat P, Fraser WD, Wierzbicki AS, Fogelman I, Goldsmith naling and phosphate/calcium metabolism. Curr Opin Nephrol
DJ, Hampson G. Fibroblast growth factor-23 is associated Hypertens 2006;15(4):437–41.
with C-reactive protein, serum phosphate and bone mineral [74] 
Urakawa I, Yamazaki Y, Shimada T, Iijima K, Hasegawa H,
density in chronic kidney disease. Osteoporos Int 2010;21(11): Okawa K, et al. Klotho converts canonical FGF receptor into a
1853–61. specific receptor for FGF23. Nature 2006;444(7120):770–4.
[58] 
Takeda Y, Komaba H, Goto S, Fujii H, Umezu M, Hasegawa H, [75] 
Bushinsky DA. Acidosis and renal bone disease. In: Olgaard K,
et al. Effect of intravenous saccharated ferric oxide on serum Salusky IB, Silver J, editors. The Spectrum of Mineral and Bone
FGF23 and mineral metabolism in hemodialysis patients. Am J Disorders in Chronic Kidney Disease. Oxford: Oxford University
Nephrol 2011;33(5):421–6. Press; 2010. p. 253–65.
[59] 
Young A, Hodsman AB, Boudville N, Geddes C, Gill J, Goltzman [76] 
Jehle S, Hulter HN, Krapf R. Effect of potassium citrate on bone
D, et al. Bone and mineral metabolism and fibroblast growth fac- density, microarchitecture, and fracture risk in healthy older
tor 23 levels after kidney donation. Am J Kidney Dis 2011;59(6): adults without osteoporosis: a randomized controlled trial. J Clin
761–9. Endocrinol Metab 2012;98(1):207–17.

VII.  SPECTRUM OF SECONDARY OSTEOPOROSIS


Hypercalciuria 1419
[77] 
Frassetto L, Morris Jr RC, Sebastian A. Long-term persis- [94]  Robling AG, Niziolek PJ, Baldridge LA, Condon KW,
tence of the urine calcium-lowering effect of potassium bicar- Allen MR, Alam I, et al. Mechanical stimulation of bone in vivo
bonate in postmenopausal women. J Clin Endocrinol Metab reduces osteocyte expression of sost/sclerostin. J Biol Chem
2005;90(2):831–4. 2008;283(9):5866–75.
[78] 
Roderick P, Willis NS, Blakeley S, Jones C, Tomson C. Correction [95]  Cejka D, Herberth J, Branscum AJ, Fardo DW, Monier-Faugere
of chronic metabolic acidosis for chronic kidney disease patients. MC, Diarra D, et al. Sclerostin and Dickkopf-1 in renal osteodys-
Cochrane Database Syst Rev 2007;1; CD001890. trophy. Clin J Am Soc Nephrol 2011;6(4):877–82.
[79] 
Doumouchtsis KK, Kostakis AI, Doumouchtsis SK, Grapsa EI, [96]  Sabbagh Y, Graciolli FG, O’Brien S, Tang W, dos Reis LM, et al.
Passalidou IA, Tziamalis MP, et al. The effect of sexual hormone Repression of osteocyte Wnt/beta-catenin signaling is an early
abnormalities on proximal femur bone mineral density in hemo- event in the progression of renal osteodystrophy. J Bone Miner
dialysis patients and the possible role of RANKL. Hemodial Int Res 2012;27(8):1757–72.
2008;12(1):100–7. [97]  Wang D, Dai C, Li Y, Liu Y. Canonical Wnt/beta-catenin signal-
[80] 
Stehman-Breen C, Anderson G, Gibson D, Kausz AT, Ott S. ing mediates transforming growth factor-beta1-driven podocyte
Pharmacokinetics of oral micronized beta-estradiol in postmeno- injury and proteinuria. Kidney Int 2011;80(11):1159–69.
pausal women receiving maintenance hemodialysis. Kidney Int [98]  He W, Dai C, Li Y, Zeng G, Monga SP, Liu Y. Wnt/beta-catenin
2003;64:290–4. signaling promotes renal interstitial fibrosis. J Am Soc Nephrol
[81] 
Meng J, Ohlsson C, Laughlin GA, Chonchol M, Wassel CL, 2009;20(4):765–76.
Ljunggren O, et al. Osteoporotic Fractures in Men Study G. [99]  Chaudhary LR, Hofmeister AM, Hruska KA. Differential growth
Associations of estradiol and testosterone with serum phospho- factor control of bone formation through osteoprogenitor differ-
rus in older men: the Osteoporotic Fractures in Men study. Kid- entiation. Bone 2004;34(3):402–11.
ney Int 2010;78(4):415–22. [100] Gonzalez EA, Lund RJ, Martin KJ, McCartney JE, Tondravi
[82] 
Carrillo-Lopez N, Roman-Garcia P, Rodriguez-Rebollar A, MM, Sampath TK, et al. Treatment of a murine model of high-
Fernandez-Martin JL, Naves-Diaz M, Cannata-Andia JB. Indirect turnover renal osteodystrophy by exogenous BMP-7. Kidney Int
regulation of PTH by estrogens may require FGF23. J Am Soc 2002;61(4):1322–31.
Nephrol 2009;20(9):2009–17. [101] Lund RJ, Davies MR, Brown AJ, Hruska KA. Successful treat-
[83] 
Ix JH, Chonchol M, Laughlin GA, Shlipak MG, Whooley MA. ment of an adynamic bone disorder with bone morphoge-
Relation of sex and estrogen therapy to serum fibroblast growth netic protein-7 in a renal ablation model. J Am Soc Nephrol
factor 23, serum phosphorus, and urine phosphorus: the Heart 2004;15(2):359–69.
and Soul Study. Am J Kidney Dis 2011;58(5):737–45. [102] Hruska KA, Saab G, Mathew S, Lund R. Renal osteodystrophy,
[84] 
Fulzele K, Clemens TL. Novel functions for insulin in bone. Bone phosphate homeostasis, and vascular calcification. Semin Dial
2012;50(2):452–6. 2007;20(4):309–15.
[85] 
Ferron M, Hinoi E, Karsenty G, Ducy P. Osteocalcin differentially [103] Ott SM, Maloney NA, Coburn JW, Alfrey AC, Sherrard DJ. The
regulates beta cell and adipocyte gene expression and affects the prevalence of bone aluminum deposition in renal osteodystro-
development of metabolic diseases in wild-type mice. Proc Natl phy and its relation to the response to calcitriol therapy. N Engl J
Acad Sci U S A 2008;105(13):5266–70. Med 1982;307:709–13.
[86] 
Brennan-Speranza TC, Henneicke H, Gasparini SJ, Blankenstein [104] McCarthy JT, Hodgson SF, Fairbanks VF, Moyer TP. Clinical
KI, Heinevetter U, Cogger VC, et al. Osteoblasts mediate the ad- and histologic features of iron-related bone disease in dialysis
verse effects of glucocorticoids on fuel metabolism. J Clin Invest patients. Am J Kidney Dis 1991;17(5):551–61.
2012;122(11):4172–89. [105] Valenti L, Varenna M, Fracanzani AL, Rossi V, Fargion S, Sin-
[87] 
Okuno S, Ishimura E, Tsuboniwa N, Norimine K, Yamakawa K, igaglia L. Association between iron overload and osteoporosis
Yamakawa T, et al. Significant inverse relationship between se- in patients with hereditary hemochromatosis. Osteoporos Int
rum undercarboxylated osteocalcin and glycemic control in main- 2009;20(4):549–55.
tenance hemodialysis patients. Osteoporos Int 2013;24(2):605–12. [106] Yang Q, Jian J, Abramson SB, Huang X. Inhibitory effects of iron
[88] 
Kiepe D, Tonshoff B. Insulin-like growth factors in nor- on bone morphogenetic protein 2-induced osteoblastogenesis. J
mal and diseased kidney. Endocrinol Metab Clin North Am Bone Miner Res 2011;26(6):1188–96.
2012;41(2):351–74. [107] Oste L, Bervoets AR, Behets GJ, Dams G, Marijnissen RL,
[89] 
Teppala S, Shankar A, Sabanayagam C. Association between Geryl H, et al. Time-evolution and reversibility of strontium-
IGF-1 and chronic kidney disease among US adults. Clin Exp induced osteomalacia in chronic renal failure rats. Kidney Int
Nephrol 2010;14(5):440–4. 2005;67(3):920–30.
[90] 
Jehle PM, Ostertag A, Schulten K, Schulz W, Jehle DR, Stracke [108] Schrooten I, Behets GJ, Cabrera WE, Vercauteren SR, Lamberts
S, et al. Insulin-like growth factor system components in LV, Verberckmoes SC, et al. Dose-dependent effects of strontium
hyperparathyroidism and renal osteodystrophy. Kidney Int on bone of chronic renal failure rats. Kidney Int 2003;63(3):927–35.
2000;57(2):423–36. [109] D’Haese PC, Schrooten I, Goodman WG, Cabrera WE, Lamberts
[91] 
Iglesias P, Diez JJ, Fernandez-Reyes MJ, Mendez J, Bajo MA, LV, Elseviers MM, et al. Increased bone strontium levels in
Aguilera A, et al. Growth hormone, IGF-I and its binding pro- hemodialysis patients with osteomalacia. Kidney Int 2000;57(3):
teins (IGFBP-1 and -3) in adult uraemic patients undergoing 1107–14.
peritoneal dialysis and haemodialysis. Clin Endocrinol (Oxf) [110] Cohen-Solal M. Strontium overload and toxicity: impact on
2004;60(6):741–9. renal osteodystrophy. Nephrol Dial Transplant 2002;17(Suppl.
[92] 
Han DS, Chen YM, Lin SY, Chang HH, Huang TM, Chi YC, et al. 2):30–4.
Serum myostatin levels and grip strength in normal subjects and [111] Doublier A, Farlay D, Khebbab MT, Jaurand X, Meunier PJ,
patients on maintenance haemodialysis. Clin Endocrinol (Oxf) Boivin G. Distribution of strontium and mineralization in iliac
2011;75(6):857–63. bone biopsies from osteoporotic women treated long-term with
[93] 
Little RD, Carulli JP, Del Mastro RG, Dupuis J, Osborne M, et al. strontium ranelate. Eur J Endocrinol 2011;165(3):469–76.
A mutation in the LDL receptor-related protein 5 gene results in [112] Bardin T, Kuntz D, Zingraff J, Voisin MC, Zelmar A, Lansaman J.
the autosomal dominant high-bone-mass trait. Am J Hum Genet Synovial amyloidosis in patients undergoing long-term hemodi-
2002;70(1):11–9. alysis. Arthritis Rheum 1985;28(9):1052–8.

VII.  SPECTRUM OF SECONDARY OSTEOPOROSIS


1420 58.  OSTEOPOROSIS ASSOCIATED WITH CHRONIC KIDNEY DISEASE

[113] Adams JE. Dialysis bone disease. Semin Dial 2002;15(4):277–89. [132] Lotinun S, Sibonga JD, Turner RT. Evidence that the cells respon-
[114] Palmer SC, Rabindranath KS, Craig JC, Roderick PJ, sible for marrow fibrosis in a rat model for hyperparathyroidism
Locatelli F, Strippoli GF. High-flux versus low-flux membranes are preosteoblasts. Endocrinology 2005;146:4074–81.
for end-stage kidney disease. Cochrane Database Syst Rev [133] Calvi LM, Sims NA, Hunzelman JL, Knight MC, Giovannetti A,
2012;9(9):CD005016. Saxton JM, et al. Activated parathyroid hormone/parathyroid
[115] Jamal SA, Cheung AM, West SL, Lok CE. Bone mineral density hormone-related protein receptor in osteoblastic cells differentially
by DXA and HR pQCT can discriminate fracture status in men affects cortical and trabecular bone. J Clin Invest 2001;107:277–86.
and women with stages 3 to 5 chronic kidney disease. Osteopo- [134] Rogers A, Eastell R. Circulating osteoprotegerin and receptor
ros Int 2012;23(12):2805–13. activator for nuclear factor kappaB ligand: clinical utility in
[116] West SL, Jamal SA, Lok CE. Tests of neuromuscular function are metabolic bone disease assessment. J Clin Endocrinol Metab
associated with fractures in patients with chronic kidney disease. 2005;90(11):6323–31.
Nephrol Dial Transplant 2011;27(6):2384–8. [135] Garnero P, Sornay-Rendu E, Claustrat B, Delmas PD. Biochemi-
[117] Li M, Tomlinson G, Naglie G, Cook WL, Jassal SV. Geriatric cal markers of bone turnover, endogenous hormones and the
comorbidities, such as falls, confer an independent mortal- risk of fractures in postmenopausal women: the OFELY study.
ity risk to elderly dialysis patients. Nephrol Dial Transplant J Bone Miner Res 2000;15:1526–36.
2008;23(4):1396–400. [136] Gerdhem P, Ivaska KK, Alatalo SL, Halleen JM, Hellman J,
[118] Workeneh BT, Mitch WE. Review of muscle wasting associ- Isaksson A, et al. Biochemical markers of bone metabolism
ated with chronic kidney disease. Am J Clin Nutr 2010;91(4): and prediction of fracture in elderly women. J Bone Miner Res
1128S–32S. 2004;19(3):386–93.
[119] Coco M, Rush H. Increased incidence of hip fractures in dialysis [137] Johnell O, Oden A, De Laet C, Garnero P, Delmas PD, Kanis JA.
patients with low serum parathyroid hormone. Am J Kidney Dis Biochemical indices of bone turnover and the assessment of frac-
2000;36:1115–21. ture probability. Osteoporos Int 2002;13(7):523–6.
[120] Mittalhenkle A, Gillen DL, Stehman-Breen CO. Increased risk of [138] Urena P, Bernard-Poenaru O, Ostertag A, Baudoin C, Cohen-
mortality associated with hip fracture in the dialysis population. Solal M, Cantor T, et al. Bone mineral density, biochemical mark-
Am J Kidney Dis 2004;44(4):672–9. ers and skeletal fractures in haemodialysis patients. Nephrol
[121] Danese MD, Kim J, Doan QV, Dylan M, Griffiths R, Chertow Dial Transplant 2003;18(11):2325–31.
GM. PTH and the risks for hip, vertebral, and pelvic fractures [139] Fahrleitner-Pammer A, Herberth J, Browning SR, Obermayer-
among patients on dialysis. Am. J. Kidney Dis 2006;47(1):149–56. Pietsch B, Wirnsberger G, Holzer H, et al. Bone markers predict
[122] Kaneko TM, Foley RN, Gilbertson DT, Collins AJ. Clinical epide- cardiovascular events in chronic kidney disease. J Bone Miner
miology of long-bone fractures in patients receiving hemodialy- Res 2008;23(11):1850–8.
sis. Clin Orthop Relat Res 2007;457:188–93. [140] Ott SM. Review article: bone density in patients with chronic
[123] Iimori S, Mori Y, Akita W, Kuyama T, Takada S, Asai T, et al. kidney disease stages 4–5. Nephrology (Carlton) 2009;14(4):395–
Diagnostic usefulness of bone mineral density and biochemical 403.
markers of bone turnover in predicting fracture in CKD stage 5D [141] Ott SM. When bone mass fails to predict bone failure. Calcif Tis-
patients—a single-center cohort study. Nephrol Dial Transplant sue Int 1993;53(Suppl. 1):S7–13.
2012;27(1):345–51. [142] Lindergard B, Johnell O, Nilsson BE, Wiklung PE. Studies of
[124] Rudser KD, de Boer IH, Dooley A, Young B, Kestenbaum B. Frac- bone morphology, bone densitometry, and laboratory data in
ture risk after parathyroidectomy among chronic hemodialysis patients on maintenance hemodialysis treatment. Nephron
patients. J Am Soc Nephrol 2007;18(8):2401–7. 1985;39:122–9.
[125] Blayney MJ, Pisoni RL, Bragg-Gresham JL, Bommer J, Piera L, [143] Nickolas TL, Stein E, Cohen A, Thomas V, Staron RB, McMahon
Saito A, et al. High alkaline phosphatase levels in hemodialysis DJ, et al. Bone mass and microarchitecture in CKD patients with
patients are associated with higher risk of hospitalization and fracture. J Am Soc Nephrol 2010;21(8):1371–80.
death. Kidney Int 2008;74(5):655–63. [144] Jiang JQ, Lin S, Xu PC, Zheng ZF, Jia JY. Serum osteoprotegerin
[126] Ott SM. Bone histomorphometry in renal osteodystrophy. Semin measurement for early diagnosis of chronic kidney disease-
Nephrol 2009;29(2):122–32. mineral and bone disorder. Nephrology (Carlton) 2011;16(6):588–94.
[127] Araujo SM, Ambrosoni P, Lobao RR, Caorsi H, Moyses RM, Bar- [145] Nishiura R, Fujimoto S, Sato Y, Yamada K, Hisanaga S, Hara
reto FC, et al. The renal osteodystrophy pattern in Brazil and S, et al. Elevated osteoprotegerin levels predict cardiovas-
Uruguay: an overview. Kidney Int Suppl 2003;85:S54–6. cular events in new hemodialysis patients. Am J Nephrol
[128] Gerakis A, Hadjidakis D, Kokkinakis E, Apostolou T, Raptis S, 2009;29(3):257–63.
Billis A. Correlation of bone mineral density with the histological [146] Rogers NM, Coates PT. Calcific uraemic arteriolopathy: an
findings of renal osteodystrophy in patients on hemodialysis. J update. Curr Opin Nephrol Hypertens 2008;17(6):629–34.
Nephrol 2000;13:437–43. [147] Raymond CB, Wazny LD. Sodium thiosulfate, bisphosphonates,
[129] Malluche HH, Mawad HW, Monier-Faugere MC. Renal osteo- and cinacalcet for treatment of calciphylaxis. Am J Health Syst
dystrophy in the first decade of the new millennium: analysis of Pharm 2008;65(15):1419–29.
630 bone biopsies in black and white patients. J Bone Miner Res [148] Mason D, Best SD. Calcific uremic arteriolopathy: contemporary
2011;26(6):1368–76. pharmacotherapy. Adv Chronic Kidney Dis 2010;17(5):428–38.
[130] Kalantar-Zadeh K, Miller JE, Kovesdy CP, Mehrotra R, [149] Kinuya S, Taki J, Fujioka M, Michigishi T, Tonami N. Metastatic
Lukowsky LR, Streja E, et al. Impact of race on hyperpara- calcification: accumulation of a bone tracer during dynamic data
thyroidism, mineral disarrays, administered vitamin D mi- acquisition. Ann Nucl Med 1996;10(4):433–5.
metic, and survival in hemodialysis patients. J Bone Miner Res [150] Breitz HB, Sirotta PS, Nelp WB, Ott S, Figley MM. Progressive
2010;25(12):2724–34. pulmonary calcification complicating successful renal transplan-
[131] Lowry MB, Lotinun S, Leontovich AA, Zhang M, Maran A, tation. Am Rev Respir Dis 1987;136(6):1480–2.
Shogren KL, et al. Osteitis fibrosa is mediated by platelet-derived [151] Conger JD, Hammond WS, Alfrey AC, Contiguglia SR, Stan-
growth factor-a via a phosphoinositide 3-kinase-dependent sig- ford RE, Huffer WE. Pulmonary calcification in chronic dialy-
naling pathway in a rat model for chronic hyperparathyroidism. sis patients. Clinical and pathologic studies. Ann Intern Med
Endocrinology 2008;149:5735–46. 1975;83(3):330–6.

VII.  SPECTRUM OF SECONDARY OSTEOPOROSIS


Hypercalciuria 1421
[152] Mizobuchi M, Towler D, Slatopolsky E. Vascular calcification: the [171] Manghat P, Souleimanova I, Cheung J, Wierzbicki AS,
killer of patients with chronic kidney disease. J Am Soc Nephrol Harrington DJ, Shearer MJ, et al. Association of bone turnover
2009;20(7):1453–64. markers and arterial stiffness in pre-dialysis chronic kidney dis-
[153] Pai AS, Giachelli CM. Matrix remodeling in vascular calcifica- ease (CKD). Bone 2011;48(5):1127–32.
tion associated with chronic kidney disease. J Am Soc Nephrol [172] Cannata-Andia JB, Roman-Garcia P, Hruska K. The connections
2010;21(10):1637–40. between vascular calcification and bone health. Nephrol Dial
[154] Shanahan CM, Crouthamel MH, Kapustin A, Giachelli CM. Transplant 2011;26(11):3429–36.
Arterial calcification in chronic kidney disease: key roles for [173] Hruska KA, Mathew S. The roles of the skeleton and phospho-
calcium and phosphate. Circ Res 2011;109(6):697–711. rus in the CKD mineral bone disorder. Adv Chronic Kidney Dis
[155] Shroff R, Long DA, Shanahan C. Mechanistic insights into vascu- 2011;18(2):98–104.
lar calcification in CKD. J Am Soc Nephrol 2013;24(2):179–89. [174] Moorthi RN, Kandula P, Moe SM. Optimal vitamin D, calcitriol,
[156] El-Abbadi MM, Pai AS, Leaf EM, Yang HY, Bartley BA, Quan and vitamin D analog replacement in chronic kidney disease: to
KK, et al. Phosphate feeding induces arterial medial calcification D or not to D: that is the question. Curr Opin Nephrol Hypertens
in uremic mice: role of serum phosphorus, fibroblast growth 2011;20(4):354–9.
factor-23, and osteopontin. Kidney Int 2009;75(12):1297–307. [175] Wasse H, Huang R, Long Q, Singapuri S, Raggi P, Tangpricha V.
[157] Giachelli CM. The emerging role of phosphate in vascular calci- Efficacy and safety of a short course of very-high-dose cholecal-
fication. Kidney Int 2009;75(9):890–7. ciferol in hemodialysis. Am J Clin Nutr 2012;95(2):522–8.
[158] Roman-Garcia P, Carrillo-Lopez N, Fernandez-Martin JL, Naves- [176] Tentori F, Hunt WC, Stidley CA, Rohrscheib MR, Bedrick EJ, Mey-
Diaz M, Ruiz-Torres MP, Cannata-Andia JB. High phosphorus er KB, et al. Mortality risk among hemodialysis patients receiving
diet induces vascular calcification, a related decrease in bone mass different vitamin D analogs. Kidney Int 2006;70(10):1858–65.
and changes in the aortic gene expression. Bone 2010;46(1):121–8. [177] Palmer SC, McGregor DO, Macaskill P, Craig JC, Elder GJ, Strip-
[159] Shalhoub V, Shatzen EM, Ward SC, Young JI, Boedigheimer M, poli GF. Meta-analysis: vitamin D compounds in chronic kidney
Twehues L, et al. Chondro/osteoblastic and cardiovascular gene disease. Ann Intern Med 2007;147(12):840–53.
modulation in human artery smooth muscle cells that calcify in [178] Thadhani R, Appelbaum E, Pritchett Y, Chang Y, Wenger J,
the presence of phosphate and calcitriol or paricalcitol. J Cell Bio- Tamez H, et al. Vitamin D therapy and cardiac structure and
chem 2010;111(4):911–21. function in patients with chronic kidney disease: the PRIMO
[160] Torres PA, De Broe M. Calcium-sensing receptor, calcimimetics, randomized controlled trial. JAMA 2012;307(7):674–84.
and cardiovascular calcifications in chronic kidney disease. Kid- [179] Kovesdy CP, Kuchmak O, Lu JL, Kalantar-Zadeh K. Outcomes
ney Int 2012;82(1):19–25. associated with phosphorus binders in men with non-dialysis-
[161] Investigators Evolve. Effect of cinacalcet on cardiovascu- dependent CKD. Am J Kidney Dis 2010;56(5):842–51.
lar disease in patients undergoing dialysis. N Engl J Med [180] Block GA, Wheeler DC, Persky MS, Kestenbaum B, Ketteler
2012;367(26):2482–94. M, Spiegel DM, et al. Effects of phosphate binders in moderate
[162] Thambiah S, Roplekar R, Manghat P, Fogelman I, Fraser WD, CKD. J Am Soc Nephrol 2012;23(8):1407–15.
Goldsmith D, et al. Circulating sclerostin and dickkopf-1 [181] Navaneethan SD, Palmer SC, Vecchio M, Craig JC, Elder GJ,
(DKK1) in predialysis chronic kidney disease (CKD), relation- Strippoli GF. Phosphate binders for preventing and treating
ship with bone density and arterial stiffness. Calcif Tissue Int bone disease in chronic kidney disease patients. Cochrane Data-
2012;90(6):473–80. base Syst Rev 2011;2; CD006023.
[163] Al-Aly Z. Phosphate, oxidative stress, and nuclear factor-kap- [182] Elder GJ. Calcium supplementation: lessons from the general
paB activation in vascular calcification. Kidney Int 2011;79(10): population for chronic kidney disease and back. Curr Opin
1044–7. Nephrol Hypertens 2011;20(4):369–75.
[164] Zavaczki E, Jeney V, Agarwal A, Zarjou A, Oros M, Katko M, [183] Coburn JW, Mischel MG, Goodman WG, Salusky IB. Calcium
et al. Hydrogen sulfide inhibits the calcification and osteoblas- citrate markedly enhances aluminum absorption from alumi-
tic differentiation of vascular smooth muscle cells. Kidney Int num hydroxide. Am J Kidney Dis 1991;17(6):708–11.
2011;80(7):731–9. [184] Komaba H, Nakanishi S, Fujimori A, Tanaka M, Shin J, Shibuya
[165] Hamano T, Matsui I, Mikami S, Tomida K, Fujii N, Imai E, et al. K, et al. Cinacalcet effectively reduces parathyroid hormone
Fetuin-mineral complex reflects extraosseous calcification stress secretion and gland volume regardless of pretreatment gland
in CKD. J Am Soc Nephrol 2010;21(11):1998–2007. size in patients with secondary hyperparathyroidism. Clin J Am
[166] Mathew S, Davies M, Lund R, Saab G, Hruska KA. Function and Soc Nephrol 2010;5(12):2305–14.
effect of bone morphogenetic protein-7 in kidney bone and the [185] Block GA, Martin KJ, de Francisco AL, Turner SA, Avram MM,
bone-vascular links in chronic kidney disease. Eur J Clin Invest et al. Cinacalcet for secondary hyperparathyroidism in patients
2006;36(Suppl. 2):43–50. receiving hemodialysis. N Engl J Med 2004;350(15):1516–25.
[167] Tomiyama C, Carvalho AB, Higa A, Jorgetti V, Draibe SA, Can- [186] Block GA, Zaun D, Smits G, Persky M, Brillhart S, Nieman K,
ziani ME. Coronary calcification is associated with lower bone et al. Cinacalcet hydrochloride treatment significantly improves
formation rate in CKD patients not yet in dialysis treatment. J all-cause and cardiovascular survival in a large cohort of hemo-
Bone Miner Res 2010;25(3):499–504. dialysis patients. Kidney Int 2010;78(6):578–89.
[168] Barreto DV, Barreto FC, Carvalho AB, Cuppari L, Cendoroglo M, [187] Cunningham J, Danese M, Olson K, Klassen P, Chertow GM.
Draibe SA, et al. Coronary calcification in hemodialysis patients: Effects of the calcimimetic cinacalcet HCl on cardiovascular
the contribution of traditional and uremia-related risk factors. disease, fracture, and health-related quality of life in secondary
Kidney Int 2005;67(4):1576–82. hyperparathyroidism. Kidney Int 2005;68(4):1793–800.
[169] London GM, Marty C, Marchais SJ, Guerin AP, Metivier F, de [188] Chonchol M, Locatelli F, Abboud HE, Charytan C, de Francisco
Vernejoul MC. Arterial calcifications and bone histomorphom- AL, et al. A randomized, double-blind, placebo-controlled study to
etry in end-stage renal disease. J Am Soc Nephrol 2004;15(7): assess the efficacy and safety of cinacalcet HCl in participants with
1943–51. CKD not receiving dialysis. Am J Kidney Dis 2009;53(2):197–207.
[170] Asci G, Ok E, Savas R, Ozkahya M, Duman S, Toz H, et al. The [189] Desmet C, Beguin C, Swine C, Jadoul M. Falls in hemodialysis
link between bone and coronary calcifications in CKD-5 patients patients: prospective study of incidence, risk factors, and com-
on haemodialysis. Nephrol Dial Transplant 2011;26(3):1010–5. plications. Am J Kidney Dis 2005;45(1):148–53.

VII.  SPECTRUM OF SECONDARY OSTEOPOROSIS


1422 58.  OSTEOPOROSIS ASSOCIATED WITH CHRONIC KIDNEY DISEASE

[190] Cunningham J, Sprague SM, Cannata-Andia J, Coco M, Cohen- [209] Hartle JE, Tang X, Kirchner HL, Bucaloiu ID, Sartorius JA,
Solal M, Fitzpatrick L, et al. Osteoporosis in chronic kidney Pogrebnaya ZV, et al. Bisphosphonate therapy, death, and car-
disease. Am J Kidney Dis 2004;43:566–71. diovascular events among female patients with CKD: a retro-
[191] Anderson GL, Limacher M, Assaf AR, Bassford T, Beresford spective cohort study. Am J Kidney Dis 2012;59(5):636–44.
SA, Black H, et al. Effects of conjugated equine estrogen in post- [210] Lu KC, Yeung LK, Lin SH, Lin YF, Chu P. Acute effect of pami-
menopausal women with hysterectomy: the Women’s Health dronate on PTH secretion in postmenopausal hemodialysis
Initiative randomized controlled trial. JAMA 2004;291:1701–12. patients with secondary hyperparathyroidism. Am J Kidney Dis
[192] Rossouw JE, Anderson GL, Prentice RL, LaCroix AZ, Kooper- 2003;42(6):1221–7.
berg C, Stefanick ML, et al. Risks and benefits of estrogen plus [211] Torregrosa JV, Moreno A, Mas M, Ybarra J, Fuster D. Useful-
progestin in healthy postmenopausal women: principal results ness of pamidronate in severe secondary hyperparathyroidism
from the Women’s Health Initiative randomized controlled trial. in patients undergoing hemodialysis. Kidney Int Suppl 2003:
JAMA 2002;288:321–33. S88–90.
[193] Manson JE, Allison MA, Rossouw JE, Carr JJ, Langer RD, Hsia J, [212] Bergner R, Henrich D, Hoffmann M, Schmidt-Gayk H, Lenz T,
et al. Estrogen therapy and coronary-artery calcification. N Engl Upperkamp M. Treatment of reduced bone density with iban-
J Med 2007;356(25):2591–602. dronate in dialysis patients. J Nephrol 2008;21(4):510–6.
[194] Matuszkiewicz-Rowinska J, Skorzewska K, Radowicki S, [213] Amerling R, Harbord NB, Pullman J, Feinfeld DA. Bisphos-
Sokalski A, Przedlacki J, Niemczyk S, et al. The benefits of hor- phonate use in chronic kidney disease: association with ady-
mone replacement therapy in pre-menopausal women with oes- namic bone disease in a bone histology series. Blood Purif
trogen deficiency on haemodialysis. Nephrol Dial Transplant 2010;29(3):293–9.
1999;14(5):1238–43. [214] Mitsopoulos E, Ginikopoulou E, Economidou D, Zanos S,
[195] Guevara A, Vidt D, Hallberg MC, Zorn EM, Pohlman C, Wieland Pateinakis P, Minasidis E, et al. Impact of long-term cina-
RG. Serum gonadotropin and testosterone levels in uremic males calcet, ibandronate or teriparatide therapy on bone mineral
undergoing intermittent dialysis. Metabolism 1969;18(12):1062–6. density of hemodialysis patients: a pilot study. Am J Nephrol
[196] Albaaj F, Sivalingham M, Haynes P, McKinnon G, Foley RN, 2012;36(3):238–44.
Waldek S, et al. Prevalence of hypogonadism in male patients [215] Eriguchi R, Umakoshi J, Miura S, Sato Y. Raloxifene amelio-
with renal failure. Postgrad Med J 2006;82(972):693–6. rates progressive bone loss in postmenopausal dialysis patients
[197] Bauer DC, Black DM, Garnero P, Hochberg M, Ott S, Orloff J, with controlled parathyroid hormone levels. Clin Nephrol
et al. Change in bone turnover and hip, non-spine, and vertebral 2009;72(6):423–9.
fracture in alendronate-treated women: the fracture intervention [216] Hernandez E, Valera R, Alonzo E, Bajares-Lilue M, Carlini R,
trial. J Bone Miner Res 2004;19(8):1250–8. Capriles F, et al. Effects of raloxifene on bone metabolism and
[198] Ott SM. Bisphosphonate safety and efficacy in chronic kidney serum lipids in postmenopausal women on chronic hemodialysis.
disease. Kidney Int 2012;82(8):833–5. Kidney Int 2003;63:2269–74.
[199] Lin JH. Bisphosphonates: a review of their pharmacokinetic [217] Tanaka M, Itoh K, Matsushita K, Moriishi M, Kawanishi H,
properties. Bone 1996;18(2):75–85. Fukagawa M. Effects of raloxifene on bone mineral metabolism
[200] Buttazzoni M, Rosa Diez GJ, Jager V, Crucelegui MS, Algranati in postmenopausal Japanese women on hemodialysis. Ther
SL, et al. Elimination and clearance of pamidronate by haemodi- Apher Dial 2011;15(Suppl. 1):62–6.
alysis. Nephrology (Carlton) 2006;11(3):197–200. [218] Cejka D, Kodras K, Bader T, Haas M. Treatment of hemodialysis-
[201] Perazella MA, Markowitz GS. Bisphosphonate nephrotoxicity. associated adynamic bone disease with teriparatide (PTH(1-34)):
Kidney Int 2008;74(11):1385–93. a pilot study. Kidney Blood Press Res 2010;33(3):221–6.
[202] Boonen S, Sellmeyer DE, Lippuner K, Orlov-Morozov A, Abrams [219] Block GA, Bone HG, Fang L, Lee E, Padhi D. A single-dose study
K, Mesenbrink P, et al. Renal safety of annual zoledronic acid of denosumab in patients with various degrees of renal impair-
infusions in osteoporotic postmenopausal women. Kidney Int ment. J Bone Miner Res 2012;27(7):1471–9.
2008;74(5):641–8. [220] Matuszkiewicz-Rowinska J, Niemczyk S, Przedlacki J, Puka J,
[203] Lyles KW, Colon-Emeric CS, Magaziner JS, Adachi JD, Pieper CF, Switalski M, Ostrowski K. Effect of salmon calcitonin on bone
Mautalen C, et al. Zoledronic acid in reducing clinical fracture and mineral density and calcium-phosphate metabolism in chronic
mortality after hip fracture. N Engl J Med 2007; 357, nihpa40967. hemodialysis patients with secondary hyperparathyroidism. Pol
[204] Markowitz GS, Appel GB, Fine PL, Fenves AZ, Loon NR, Jag- Arch Med Wewn 2004;112(1):797–803.
annath S, et al. Collapsing focal segmental glomerulosclerosis [221] Wang SX, Li H. Salmon calcitonin in prevention of osteo-
following treatment with high-dose pamidronate. J Am Soc porosis in maintenance dialysis patients. Chin Med J (Engl)
Nephrol 2001;12(6):1164–72. 2008;121(14):1280–4.
[205] Ariyoshi T, Eishi K, Sakamoto I, Matsukuma S, Odate T. Effect [222] Pienkowski D, Doers TM, Monier-Faugere MC, Geng Z, Cama-
of etidronic acid on arterial calcification in dialysis patients. Clin cho NP, Boskey AL, et al. Calcitonin alters bone quality in beagle
Drug Investig 2006;26(4):215–22. dogs. J Bone Miner Res 1997;12(11):1936–43.
[206] Hill JA, Goldin JG, Gjertson D, Emerick AM, Greaser LD, [223] Thamsborg G, Jensen JE, Kollerup G, Hauge EM, Melsen F,
Yoon HC, et al. Progression of coronary artery calcification Sorensen OH. Effect of nasal salmon calcitonin on bone remod-
in patients taking alendronate for osteoporosis. Acad Radiol eling and bone mass in postmenopausal osteoporosis. Bone
2002;9(10):1148–52. 1996;18(2):207–12.
[207] Tanko LB, Qin G, Alexandersen P, Bagger YZ, Christiansen C. [224] Frick KK, Bushinsky DA. Molecular mechanisms of primary
Effective doses of ibandronate do not influence the 3-year pro- hypercalciuria. J Am Soc Nephrol 2003;14(4):1082–95.
gression of aortic calcification in elderly osteoporotic women. [225] Devuyst O, Pirson Y. Genetics of hypercalciuric stone forming
Osteoporos Int 2004;10:10. diseases. Kidney Int 2007;72(9):1065–72.
[208] Toussaint ND, Lau KK, Strauss BJ, Polkinghorne KR, Kerr PG. [226] Stechman MJ, Loh NY, Thakker RV. Genetic causes of hypercal-
Effect of alendronate on vascular calcification in CKD stages ciuric nephrolithiasis. Pediatr Nephrol 2009;24(12):2321–32.
3 and 4: a pilot randomized controlled trial. Am J Kidney Dis [227] Monico CG, Milliner DS. Genetic determinants of urolithiasis.
2010;56(1):57–68. Nat Rev Nephrol 2012;8(3):151–62.

VII.  SPECTRUM OF SECONDARY OSTEOPOROSIS


Hypercalciuria 1423
[228] Whitson PA, Pietrzyk RA, Morukov BV, Sams CF. The risk of [246] Zimmermann B, Plank C, Konrad M, Stohr W, Gravou-Apostol-
renal stone formation during and after long duration space atou C, Rascher W, et al. Hydrochlorothiazide in CLDN16 muta-
flight. Nephron 2001;89(3):264–70. tion. Nephrol Dial Transplant 2006;21(8):2127–32.
[229] Ott SM, Chesnut CH. Calcitriol treatment is not effective in post- [247] Lambers TT, Bindels RJ, Hoenderop JG. Coordinated control of
menopausal osteoporosis. Ann Intern Med 1989;110:267–74. renal Ca2+ handling. Kidney Int 2006;69(4):650–4.
[230] Favus MJ, Karnauskas AJ, Parks JH, Coe FL. Peripheral blood [248] Magyar CE, White KE, Rojas R, Apodaca G, Friedman PA. Plas-
monocyte vitamin D receptor levels are elevated in patients ma membrane Ca2+-ATPase and NCX1 Na+/Ca2+ exchanger
with idiopathic hypercalciuria. J Clin Endocrinol Metab expression in distal convoluted tubule cells. Am J Physiol Renal
2004;89(10):4937–43. Physiol 2002;283:F29–40.
[231] Carpenter TO. Take another CYP: confirming a novel mecha- [249] Friedman PA, Gesek FA. Stimulation of calcium transport by
nism for “idiopathic” hypercalcemia. J Clin Endocrinol Metab amiloride in mouse distal convoluted tubule cells. Kidney Int
2012;97(3):768–71. 1995;48(5):1427–34.
[232] Tebben PJ, Milliner DS, Horst RL, Harris PC, Singh RJ, Wu Y, [250] Mitchell DM, Regan S, Cooley MR, Lauter KB, Vrla MC, Becker
et al. Hypercalcemia, hypercalciuria, and elevated calcitriol CB, et al. Long-term follow-up of patients with hypoparathy-
concentrations with autosomal dominant transmission due roidism. J Clin Endocrinol Metab 2012;97:4507–14.
to CYP24A1 mutations: effects of ketoconazole therapy. J Clin [251] van Abel M, Hoenderop JG, van der Kemp AW, Friedlaender
Endocrinol Metab 2012;97(3):E423–7. MM, van Leeuwen JP, et al. Coordinated control of renal
[233] Curhan GC, Willett WC, Speizer FE, Spiegelman D, Stampfer MJ. Ca(2+) transport proteins by parathyroid hormone. Kidney Int
Comparison of dietary calcium with supplemental calcium and 2005;68(4):1708–21.
other nutrients as factors affecting the risk for kidney stones in [252] Huang CL. Regulation of ion channels by secreted Klotho: mech-
women. Ann Intern Med 1997;126:497–504. anisms and implications. Kidney Int 2010;77(10):855–60.
[234] Prie D, Huart V, Bakouh N, Planelles G, Dellis O, Gerard B, et al. [253] Tsuruoka S, Nishiki K, Ioka T, Ando H, Saito Y, Kurabayashi M,
Nephrolithiasis and osteoporosis associated with hypophospha- et al. Defect in parathyroid-hormone-induced luminal calcium
temia caused by mutations in the type 2a sodium-phosphate absorption in connecting tubules of Klotho mice. Nephrol Dial
cotransporter. N Engl J Med 2002;347:983–91. Transplant 2006;21(10):2762–7.
[235] Lapointe JY, Tessier J, Paquette Y, Wallendorff B, Coady MJ, [254] Brownstein CA, Adler F, Nelson-Williams C, Iijima J, Li P, Imura
Pichette V, et al. NPT2a gene variation in calcium nephrolithi- A, et al. A translocation causing increased alpha-Klotho level
asis with renal phosphate leak. Kidney Int 2006;69(12):2261–7. results in hypophosphatemic rickets and hyperparathyroidism.
[236] White KE, Gesek FA, Nesbitt T, Drezner MK, Friedman PA. Proc Natl Acad Sci U S A 2008;105(9):3455–60.
Molecular dissection of Ca2+ efflux in immortalized proximal [255] Ichikawa S, Imel EA, Kreiter ML, Yu X, Mackenzie DS, Sorenson
tubule cells. J Gen Physiol 1997;109(2):217–28. AH, et al. A homozygous missense mutation in human Klotho
[237] Moe OW, Huang CL. Hypercalciuria from acid load: renal mech- causes severe tumoral calcinosis. J Clin Invest 2007;117(9):
anisms. J Nephrol 2006;19(Suppl. 9):S53–61. 2684–91.
[238] Raja KA, Schurman S, D’Mello RG, Blowey D, Goodyer P, Van [256] Mayan H, Munter G, Shaharabany M, Mouallem M, Pauzner
Why S, et al. Responsiveness of hypercalciuria to thiazide in R, Holtzman EJ, et al. Hypercalciuria in familial hyperkalemia
Dent’s disease. J Am Soc Nephrol 2002;13:2938–44. and hypertension accompanies hyperkalemia and precedes
[239] Brown EM, Pollak M, Hebert SC. The extracellular calcium- hypertension: description of a large family with the Q565E
sensing receptor: its role in health and disease. Annu Rev Med WNK4 mutation. J Clin Endocrinol Metab 2004;89(8):4025–30.
1998;49:15–29. [257] Wilson FH, Kahle KT, Sabath E, Lalioti MD, Rapson AK, Hoover
[240] D’Souza-Li L, Yang B, Canaff L, Bai M, Hanley DA, Bastepe RS, et al. Molecular pathogenesis of inherited hypertension with
M, et al. Identification and functional characterization of novel hyperkalemia: the Na-Cl cotransporter is inhibited by wild-
calcium-sensing receptor mutations in familial hypocalciuric type but not mutant WNK4. Proc Natl Acad Sci U S A 2003;100:
hypercalcemia and autosomal dominant hypocalcemia. J Clin 680–4.
Endocrinol Metab 2002;87:1309–18. [258] Giannini S, Nobile M, Dalle Carbonare L, Lodetti MG, Sella S,
[241] Lienhardt A, Bai M, Lagarde JP, Rigaud M, Zhang Z, Jiang Y, Vittadello G, et al. Hypercalciuria is a common and important
et al. Activating mutations of the calcium-sensing receptor: finding in postmenopausal women with osteoporosis. Eur J
management of hypocalcemia. J Clin Endocrinol Metab 2001;86: Endocrinol 2003;149:209–13.
5313–23. [259] Jones AN, Shafer MM, Keuler NS, Crone EM, Hansen KE.
[242] Sato K, Hasegawa Y, Nakae J, Nanao K, Takahashi I, Tajima T, Fasting and postprandial spot urine calcium-to-creatinine
et al. Hydrochlorothiazide effectively reduces urinary calcium ratios do not detect hypercalciuria. Osteoporos Int 2012;23(2):
excretion in two Japanese patients with gain-of-function muta- 553–62.
tions of the calcium-sensing receptor gene. J Clin Endocrinol [260] Tannenbaum C, Clark J, Schwartzman K, Wallenstein S, Lapinski
Metab 2002;87:3068–73. R, Meier D, et al. Yield of laboratory testing to identify secondary
[243] Gavalas NG, Kemp EH, Krohn KJ, Brown EM, Watson PF, Weet- contributors to osteoporosis in otherwise healthy women. J Clin
man AP. The calcium-sensing receptor is a target of autoantibod- Endocrinol Metab 2002;87:4431–7.
ies in patients with autoimmune polyendocrine syndrome type 1. [261] Deutschmann HA, Weger M, Weger W, Kotanko P, Deutschmann
J Clin Endocrinol Metab 2007;92(6):2107–14. MJ, Skrabal F. Search for occult secondary osteoporosis: impact
[244] Kifor O, Moore Jr FD, Delaney M, Garber J, Hendy GN, Butters of identified possible risk factors on bone mineral density. J
R, et al. A syndrome of hypocalciuric hypercalcemia caused by Intern Med 2002;252:389–97.
autoantibodies directed at the calcium-sensing receptor. J Clin [262] Peris P, Guanabens N, Martinez de Osaba MJ, Monegal A, Alva-
Endocrinol Metab 2003;88(1):60–72. rez L, Pons F, et al. Clinical characteristics and etiologic factors
[245] Kausalya PJ, Amasheh S, Gunzel D, Wurps H, Muller D, Fromm of premenopausal osteoporosis in a group of Spanish women.
M, et al. Disease-associated mutations affect intracellular traffic Semin Arthritis Rheum 2002;32(1):64–70.
and paracellular Mg2+ transport function of Claudin-16. J Clin [263] Heilberg IP, Weisinger JR. Bone disease in idiopathic hypercalci-
Invest 2006;116(4):878–91. uria. Curr Opin Nephrol Hypertens 2006;15(4):394–402.

VII.  SPECTRUM OF SECONDARY OSTEOPOROSIS


1424 58.  OSTEOPOROSIS ASSOCIATED WITH CHRONIC KIDNEY DISEASE

[264] Weisinger JR. New insights into the pathogenesis of idiopathic [274] Adams JS, Song CF, Kantorovich V. Rapid recovery of bone mass
hypercalciuria: the role of bone. Kidney Int 1996;49:1507–18. in hypercalciuric, osteoporotic men treated with hydrochlorothi-
[265] Ghazali A, Fuentes V, Desaint C, Bataille P, Westeel A, Brazier azide. Ann Intern Med 1999;130:658–60.
M, et al. Low bone mineral density and peripheral blood mono- [275] Rico H, Revilla M, Villa LF, Arribas I, de Buergo MA. A longitu-
cyte activation profile in calcium stone formers with idiopathic dinal study of total and regional bone mineral content and bio-
hypercalciuria. J Clin Endocrinol Metab 1997;82:32–8. chemical markers of bone resorption in patients with idiopathic
[266] Lauderdale DS, Thisted RA, Wen M, Favus MJ. Bone mineral hypercalciuria on thiazide treatment. Miner Electrolyte Metab
density and fracture among prevalent kidney stone cases in the 1993;19(6):337–42.
Third National Health and Nutrition Examination Survey. J Bone [276] LaCroix AZ, Ott SM, Ichikawa L, Scholes D, Barlow WE. Low-
Miner Res 2001;16:1893–8. dose hydrochlorothiazide and preservation of bone mineral
[267] Melton 3rd LJ, Crowson CS, Khosla S, Wilson DM, O’Fallon WM. density in older adults. A randomized, double-blind, placebo-
Fracture risk among patients with urolithiasis: a population- controlled trial. Ann Intern Med 2000;133(7):516–26.
based cohort study. Kidney Int 1998;53(2):459–64. [277] Reid IR, Ames RW, Orr-Walker BJ, Clearwater JM, Horne AM,
[268] Sakhaee K, Maalouf NM, Kumar R, Pasch A, Moe OW. Neph- Evans MC, et al. Hydrochlorothiazide reduces loss of cortical
rolithiasis-associated bone disease: pathogenesis and treatment bone in normal postmenopausal women: a randomized con-
options. Kidney Int 2011;79(4):393–403. trolled trial. Am J Med 2000;109:362–70.
[269] Malluche HH, Tschoepe W, Ritz E, Meyer-Sabellek W, Massry [278] Wasnich RD, Davis JW, He YF, Petrovich H, Ross PD. A random-
SG. Abnormal bone histology in idiopathic hypercalciuria. J Clin ized, double-masked, placebo-controlled trial of chlorthalidone
Endocrinol Metab 1980;50:654–8. and bone loss in elderly women. Osteoporos Int 1995;5(4):247–51.
[270] Steiniche T, Mosekilde L, Christensen MS, Melsen F. A histomor- [279] Ott SM, LaCroix AZ, Scholes D, Ichikawa LE, Wu K. Effects
phometric determination of iliac bone remodeling in patients of three years of low-dose thiazides on mineral metabolism in
with recurrent renal stone formation and idiopathic hypercalci- healthy elderly persons. Osteoporos Int 2008;19(9):1315–22.
uria. Apmis 1989;97:309–16. [280] Wiens M, Etminan M, Gill SS, Takkouche B. Effects of antihyper-
[271] Misael da Silva AM, dos Reis LM, Pereira RC, Futata E, Branco- tensive drug treatments on fracture outcomes: a meta-analysis of
Martins CT, Noronha IL, et al. Bone involvement in idiopathic observational studies. J Intern Med 2006;260(4):350–62.
hypercalciuria. Clin Nephrol 2002;57:183–91. [281] Aung K, Htay T. Thiazide diuretics and the risk of hip fracture.
[272] Ott SM. 2012; Hypercalciuria courses.washington.edu/bonephys/ Cochrane Database Syst Rev 2011;10;CD005185.
hypercalU/hypercal.html; (accessed 28 March 2013).
[273] Steiniche T, Mosekilde L, Christensen MS, Melsen F. Histomor-
phometric analysis of bone in idiopathic hypercalciuria before
and after treatment with thiazide. Apmis 1989;97(4):302–8.

VII.  SPECTRUM OF SECONDARY OSTEOPOROSIS

Anda mungkin juga menyukai