Alexander 2014

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

467

INVITED REVIEW
Paracellular calcium transport across renal and intestinal
epithelia1
R. Todd Alexander, Juraj Rievaj, and Henrik Dimke
Biochem. Cell Biol. Downloaded from www.nrcresearchpress.com by San Francisco (UCSF) on 12/02/14

Abstract: Calcium (Ca2+) is a key constituent in a myriad of physiological processes from intracellular signalling to the miner-
alization of bone. As a consequence, Ca2+ is maintained within narrow limits when circulating in plasma. This is accomplished
via regulated interplay between intestinal absorption, renal tubular reabsorption, and exchange with bone. Many studies have
focused on the highly regulated active transcellular transport pathways for Ca2+ from the duodenum of the intestine and the
distal nephron of the kidney. However, comparatively little work has examined the molecular constituents creating the
paracellular shunt across intestinal and renal epithelium, the transport pathway responsible for the majority of transepithelial
Ca2+ flux. More specifically, passive paracellular Ca2+ absorption occurs across the majority of the intestine in addition to the
renal proximal tubule and thick ascending limb of Henle's loop. Importantly, recent studies demonstrated that Ca2+ transport
through the paracellular shunt is significantly regulated. Therefore, we have summarized the evidence for different modes of
paracellular Ca2+ flux across renal and intestinal epithelia and highlighted recent molecular insights into both the mechanism
of secondarily active paracellular Ca2+ movement and the identity of claudins that permit the passage of Ca2+ through the tight
junction of these epithelia.

Key words: kidney, NHE3, claudins, solvent drag.

Résumé : Le calcium (Ca2+) est un important constituant d’une myriade de processus physiologiques allant de la signalisation
For personal use only.

intracellulaire à la minéralisation de l’os. Conséquemment, le Ca2+ est maintenu à l’intérieur de limites étroites lorsqu’il circule
dans le plasma. Cela s’accomplit grâce à l’influence réciproque régulée entre l’absorption intestinale, la réabsorption rénale et
l’échange avec l’os. Plusieurs études se sont concentrées sur la voie hautement régulée du transport actif trans-cellulaire du Ca2+
du duodénum et du néphron distal. Cependant, en comparaison, peu d’études ont examiné les constituants moléculaires qui
créent une dérivation para-cellulaire à travers l’épithélium intestinal et rénal, la voie de transport responsable de la majorité du
flux de Ca2+ trans-épithélial. Plus spécifiquement, l’absorption para-cellulaire passive de Ca2+ survient le long de la plus grande
partie de l’intestin en plus du tubule proximal du rein et de la branche ascendante large de l’anse de Henle. Fait important, des
études récentes démontrent que le transport de Ca2+ à travers la dérivation para-cellulaire est sujet à une importante régulation.
Les auteurs ont donc fait la synthèse des indices qui suggèrent qu’il existe différents modes de flux para-cellulaires de Ca2+ à
travers l’épithélium du rein et de l’intestin, et souligné les récentes percées sur les plans du mécanisme moléculaire du
mouvement para-cellulaire de Ca2+ secondairement actif et de l’identité des claudines qui permettent le passage du Ca2+ à travers
les jonctions serrées de ces épithéliums. [Traduit par la Rédaction]

Mots-clés : rein, NHE3, claudines, déplacement de solvant.

Introduction Ca2+ from the diet, storage, and exchange of Ca2+ with bone, and
Calcium (Ca2+)participates in a diverse array of physiological changes in the renal reabsorptive capacity for Ca2+. These pro-
processes, including neural transmission, muscle contraction, cesses effectively stabilize systemic Ca2+ concentrations within
blood coagulation, intracellular signal transduction as a second normal limits. Parathyroid glands play a central role in sensing
messenger, and the mineralization of bone as part of the hy- changes in serum Ca2+ concentrations and respond by amending
droxylapatite crystal. Consequently, the concentration of free secretion of parathyroid hormone (PTH). Increased intestinal ab-
ionized Ca2+ is tightly regulated in plasma. Hypocalcemic or hy- sorption or augmented resorption from bone increases serum
percalcemic states can cause severe neurological and cardiovascu- Ca2+ typically leading to increased levels of Ca2+ in the urine be-
lar sequelae. In fact, disturbances in Ca2+ balance may in some cause of compensatory reductions in renal transport capacity (Pak
instances produce tetany, carpopedal spasms, life-threatening et al. 1975). Similarly, dysregulation of select renal transporters
arrhythmias, coma, or cardiac arrest (Bilezikian 1993; Guise and within the kidney results in hypercalciuria. Importantly, hyper-
Mundy 1995). Globally, Ca2+ homeostasis is maintained via regu- calciuria remains the single greatest risk factor for developing a
lated interplay between 3 organ systems: intestinal absorption of Ca2+-based kidney stone (Levy et al. 1995), a disease with high and

Received 9 May 2014. Revision received 30 August 2014. Accepted 14 September 2014.
R.T. Alexander. Department of Pediatrics, The University of Alberta, 4-585 Edmonton Clinic Health Academy, 11405 – 87 Ave, Edmonton, AB T6G 2R7,
Canada; Membrane Protein Disease Research Group, The University of Alberta, Edmonton, AB T6G 2H7, Canada.
J. Rievaj. Membrane Protein Disease Research Group, The University of Alberta, Edmonton, AB T6G 2H7, Canada; Department of Physiology, The
University of Alberta, Edmonton, AB T6G 2H7, Canada.
H. Dimke. Department of Cardiovascular and Renal Research, Institute of Molecular Medicine, University of Southern Denmark, Odense, Denmark.
Corresponding author: R. Todd Alexander (e-mail: todd2@ualberta.ca).
1This review is part of a Special Issue commemorating The Canadian Society for Molecular Biosciences 57th Annual Meeting – Membrane Proteins in Health

and Disease, held in Banff, Alberta, 9–13 April 2014.

Biochem. Cell Biol. 92: 467–480 (2014) dx.doi.org/10.1139/bcb-2014-0061 Published at www.nrcresearchpress.com/bcb on 17 September 2014.
468 Biochem. Cell Biol. Vol. 92, 2014

Fig. 1. Modes of transepithelial Ca2+ flux. Ca2+ is absorbed down its electrochemical gradient via passive diffusion (top tight junction) or with
water following the osmotic gradient created by solute absorption (often sodium, bottom junction). Transcellular Ca2+ flux (bottom cell)
occurs via apical entry through a channel down the electrochemical gradient. Ca2+ then moves to the basolateral membrane bound to a
calcium-binding protein and then is extruded across the basolateral membrane via a calcium pump or in exchange for sodium.

VTE

Paracellular Diffusion
Biochem. Cell Biol. Downloaded from www.nrcresearchpress.com by San Francisco (UCSF) on 12/02/14

[Ca2+] [Ca2+]

Na+ Na+

Na+

Ca2+
For personal use only.

H2O H2O Solvent Drag

Ca2+ Ca2+
Calb
Na+
Ca2+
Ca2+ Ca 2+
Ca2+

Transcellular Flux

Lumen

increasing prevalence (Stamatelou et al. 2003; Coe et al. 2005). in exchange for Na+ by a Na+/Ca2+ exchanger (Hoenderop et al.
Therefore, it is important to establish the physiological regulation 2005; Dimke et al. 2011). Transcellular Ca2+ absorption from the
of these epithelial Ca2+ transport pathways and the molecular renal tubule and the intestine is subject to intense regulation,
constituents driving these processes in an effort to treat or pre- which has been the focus of a large amount of work and the topic
vent these diseases. of a number of reviews (Hoenderop et al. 2005; Dimke et al. 2010,
Movement of Ca2+ from the lumen of the intestine or renal 2011).
tubule into the blood can proceed via 1 of 2 pathways: transcellu- Paracellular Ca2+ flux purportedly accounts for the majority of
lar (i.e., the ion moves through the cell) or paracellular (i.e., the Ca2+ absorption from the intestine and the renal tubule. Although
ion is transported between epithelial cells; Fig. 1) (Dimke et al. once thought to be a static unregulated process, emerging evi-
2010, 2011). Transcellular Ca2+ absorption has been reported from dence suggests otherwise. Similar to ion channels, tight junctions
the duodenum, jejunum, cecum, and proximal colon in the intes- are characterized by differing size and charge selectivity, which
tine as well as the renal distal convoluted and connecting tubules controls the permeation of ions. Recent evidence has greatly
(Table 1). This process occurs via apical influx of Ca2+ through expanded our knowledge about the constituents of the tight
selective Ca2+ channels, largely predicted to be of the transient junction, which is composed of a family of proteins called the
receptor potential of the vanilloid subtype (either TRPV5 or 6). claudins. These proteins determine the paracellular permeability
Ca2+ is then shuttled to the basolateral membrane typically bound characteristics of a given epithelial type (Simon et al. 1999; Wilcox
to a Ca2+ binding protein, such as a calbindin, and then secreted et al. 2001; Konrad et al. 2006; Hou et al. 2008b; Markov et al. 2010)
across the basolateral membrane via a Ca2+-dependent ATPase or and regulation of these pore-forming claudins can dramatically

Published by NRC Research Press


Alexander et al. 469

Table 1. Type of transintestinal Ca2+ absorptiona reported by segment.


Transcellular Paracellular Paracellular flux,
Segment flux flux, passive secondarily active References
Duodenum Yes Yes Yes, secretion b (Charoenphandhu et al. 2001, 2006; Karbach 1992;
Pansu et al. 1983; Tudpor et al. 2008)
Jejunum Yesc Yes Yes, secretion (Karbach 1992; Karbach and Feldmeier 1993;
Pansu et al. 1983; Rievaj et al. 2013;
Walling and Kimberg 1973)
Ileum No Yes Yes, secretion (Hu et al. 1993; Karbach 1992; Karbach and Feldmeier 1993;
Nellans and Kimberg 1979; Pansu et al. 1983;
Walling and Kimberg 1973)
Cecum and proximal colon Yes Yes Yes, absorption (Karbach 1992; Karbach and Feldmeier 1993;
Biochem. Cell Biol. Downloaded from www.nrcresearchpress.com by San Francisco (UCSF) on 12/02/14

Karbach and Rummel 1987; Kraidith et al. 2009;


Nellans and Goldsmith 1981; Rievaj et al. 2013)
Distal colon Yes Yes Yes, secretion (Karbach et al. 1986)
aTranscellular Ca2+ flux is always absorption, the direction of passive paracellular flux is dependent on the electrochemical gradient, and the direction of
secondarily active paracellular Ca2+ flux is segment dependent.
bThere is some debate as to whether there is secondarily active Ca2+ absorption from the duodenum. Decreased flux after the apical removal of glucose supports

absorption; however, both voltage clamp experiments and analysis of concentration dependence supports a slight asymmetry mediating secretion.
cTranscellular flux has been observed across the proximal jejunum of rats.

affect the paracellular permeability characteristics of the afore- Lengemann 1962a, 1962b; Cramer 1965; Wasserman 2004). Some
mentioned epithelia (Gong et al. 2012; Dimke et al. 2013; Gong and of these studies measured the disappearance of radiotracers from
Hou 2014). This review aims to summarize current knowledge of the intestinal lumen and did not consider Ca2+ flux in the opposite
Ca2+ transport occurring via the paracellular pathway in intestine direction (i.e., secretion). Consequently, they likely underestimated
and kidney. In particular, we highlight the molecular constitu- the relative contribution of more distal segments; radiotracer ab-
ents involved in setting the driving forces and permeability char- sorption from distal segments may be falsely underrepresented
acteristics that allow passive paracellular Ca2+ flux across renal due to proximal secretion. Significant distal intestinal Ca2+ ab-
For personal use only.

and intestinal epithelia. Moreover, we describe how regulated sorption has also been observed in humans (Birge et al. 1969).
changes in specific tight junction proteins alter paracellular Ca2+
absorption from these epithelia. Passive paracellular intestinal Ca2+ absorption
Most intestinal Ca2+ absorption, on a normal Ca2+-containing
Intestine diet, is reported to occur via passive, or secondarily active, para-
Daily dietary intake of Ca2+ in the Western diet should be ap- cellular transport (Bronner 2003). This assumption is largely
proximately 1 g (Bailey et al. 2010), although a minority of this based on in vivo experiments employing the ligated loop tech-
(20%–40% of intake) is actually absorbed (Brine and Johnston 1955; nique. The relationship between luminal Ca2+ concentration and
Nordin et al. 1979). Uptake of dietary Ca2+ from the intestinal rate of Ca2+ absorption obtained by this method can be conceptu-
lumen can occur via active transcellular or passive paracellular alized as the sum of 2 processes: a saturable component, which is
transport (Bronner 1998). Active transcellular Ca2+ flux is satura- observed in the duodenum and to a lesser extent the jejunum, and
ble and predominates when dietary Ca2+ is low. This process an unsaturable process with linear dependency observed in all
permits the uptake of Ca2+ into the enterocyte when low intralu- segments of the small intestine (Pansu et al. 1983). The saturable
minal Ca2+ concentrations are present (Bronner and Pansu 1999). process is assumed to represent active transcellular absorption,
In contrast, absorption of Ca2+ via the passive paracellular path- whereas the unsaturable process is interpreted as passive paracel-
way requires a sufficiently high luminal Ca2+ concentration and, lular absorption (Bronner et al. 1986; Bronner 2003). However, this
hence, adequate dietary Ca2+ intake (Bronner and Pansu 1999); the interpretation is limited by a paucity of data supporting the as-
molecular details of both pathways are depicted in Fig. 2. The sumption that the entire unsaturable component occurs in a
small intestine is thought to be responsible for most Ca2+ uptake,
paracellular fashion; that is, one may envisage transcellular
with the colon absorbing less than 10% under normal conditions
mechanisms that may not display saturable kinetics under the
(Marcus and Lengemann 1962a, 1962b; Cramer 1965). The relative
experimental conditions (McCormick 2002; Kellett 2011). Further,
amounts of Ca2+ taken up paracellularly in different intestinal
as the relative permeability for Ca2+ (per unit of intestinal length)
segments relate to several variables. The most important are the
is similar for all segments of the small intestine (Pansu et al. 1983),
intraluminal Ca2+ concentration, the solubility of Ca2+, the per-
meability of the tight junction, and sojourn time (Duflos et al. the amount and direction of passive paracellular Ca2+ flux would
1995). Sojourn time is the time that the intestinal content spends be strongly dependent on the electrochemical gradient for Ca2+
transiting each segment. Combined, these factors determine the across intestinal epithelium. The potential difference across the
paracellular transport rate of Ca2+ across a given intestinal seg- whole length of the intestine is lumen negative (Geall and
ment. The sojourn time varies considerably between individual Summerskill 1969). Measurements of the transepithelial potential
segments, with very short transit times in the duodenum and difference in vivo across the small intestine, regardless of seg-
much longer times in the distal small intestine, cecum, and colon ment, found a difference of approximately –5 mV (lumen nega-
(Marcus and Lengemann 1962b; Duflos et al. 1995). In rats fed a tive), whereas a transepithelial voltage potential of approximately
high Ca2+ diet, sojourn times were measured to be 3 min in duo- –20 mV was measured across the large intestine (although mea-
denum, 43 min in jejunum, 141 min in ileum, 92 min in cecum, surements of up to –60 mV have been reported) (Geall et al. 1969;
and 92 min in colon (Bronner and Pansu 1999). The dissimilarity Geall and Summerskill 1969). Similar values are reported from
between sojourn times in segments of the intestine is reflected in Ussing chamber studies, although these reports typically do not
the total Ca2+ absorption measured, with 7%–8% of Ca2+ uptake exceed –20 mV even for large bowel preparations (Clarke 2009).
occurring from the duodenum, 4%–17% from the jejunum, and Consequently, the energy for paracellular Ca2+ absorption must
65%–88% from the ileum (Cramer and Copp 1959; Marcus and be derived from the chemical concentration gradient.

Published by NRC Research Press


470 Biochem. Cell Biol. Vol. 92, 2014

Fig. 2. Intestinal transepithelial Ca2+ flux. The vectorial movement of Na+ creates an osmotic gradient for water reabsorption from the
intestine, which can either create a concentration gradient for Ca2+ to diffuse down (top junction) or drive paracellular Ca2+ flux via
convection/solvent drag (bottom junction). There is evidence supporting a role for NHE3 in paracellular Ca2+ flux across some intestinal
segments and sodium-glucose transport has also been implicated (Charoenphandhu et al. 2001). The claudins (CLDN) implicated in creating
paracellular Ca2+ permeability are depicted. In the duodenum and cecum, there is significant transcellular Ca2+ absorption. The players here
include apical entry via TRPV6, shuttling to the basolateral membrane via calbindin-D9K (Calb-D9K), and then efflux across the basolateral
membrane via the calcium pump, PMCA1b. Note, we have depicted a generic intestinal epithelium; however, in different parts of the
intestine, different pathways will function to a greater or lesser extent (please refer to the text and Table 1 for details).

paracellular diffusion
Biochem. Cell Biol. Downloaded from www.nrcresearchpress.com by San Francisco (UCSF) on 12/02/14

CLDN CLDN
2 15?
Ca2+ Ca2+
CLDN CLDN
12 2
Glucose

SGLT1
+
Na Na+ a+
Na NaK
ATPase

Na+ K+
a+
Na
NHE3
H+
For personal use only.

CLDN CLDN
2+ 2 15?
Ca H2O H2O solvent drag
CLDN CLDN
12 2

Ca2+
Calb
D9K Ca2+
Ca2+
TRPV6 Ca2+ Ca2+ PMCA1b

transcellular flux

Employing the Nernst equation and assuming a free Ca2+ con- peaked at approximately 10–45 mmol/L by the terminal ileum
centration in the blood of 1.2 mmol/L, one can calculate that the (Cramer 1965; Duflos et al. 1995). This increase in the luminal Ca2+
concentration of free Ca2+ at the surface of the small bowel re- concentration along the course of the intestine occurs because
quired to overcome a –5 mV transepithelial potential difference of Na+ and, consequently, osmotically driven water absorption
must be at least 1.74 mmol/L. Yet, to our knowledge, the free Ca2+ (Cramer 1964, 1965). This secondary concentration gradient is one
concentration on the apical surface of intestinal epithelium has purported mechanism connecting paracellular Ca2+ flux to water
not been investigated. The free Ca2+ concentration in the whole
absorption. The second is solvent drag, which is the movement
lumen of different intestinal segments was found to be dependent
of Ca2+ together with paracellular water flow via convection
on the Ca2+ content of ingested food. When food with a “low” Ca2+
content (0.15% w/w) was ingested, the intraluminal free Ca2+ con- (Diamond and Bossert 1967; Pappenheimer and Reiss 1987; Larsen
centration was always observed lower than the concentration in et al. 2000). Distinguishing between a secondary concentration
blood throughout the intestine (Sernka and Borle 1969). However, gradient and solvent drag is not easily accomplished because both
when ingested Ca2+ content was increased to 1%–1.5%, an intralu- are considered secondarily active processes and are, in principle,
minal concentration of 1.5–8 mmol/L was observed. For food able to move Ca2+ ions against the observed electrochemical
containing approximately 3% Ca2+, the intraluminal free Ca2+ con- gradient. Further, the value of the transintestinal potential
centration in the small intestine increased along its course and difference and luminal concentration of Ca2+ detailed previously

Published by NRC Research Press


Alexander et al. 471

are influenced by the methods employed and model systems stud- ary concentration gradient for Ca2+ absorption, paracellular water
ied and, therefore, should be considered estimates. movement is necessary for solvent drag. However, the relative
contribution of each pathway to water absorption is unknown.
Evidence in support of secondary active paracellular Although several aquaporins have been identified in the epithelia
intestinal Ca2+ absorption of small and large intestines (Laforenza 2012), their role in trans-
Evidence for the existence of secondary active paracellular Ca2+ epithelial water flux has not been examined with the exception of
flux is largely derived from Ussing chamber studies. This in vitro aquaporin-4. This water channel plays a minor role in colonic
technique benefits from being able to control both the electrical water absorption, as evinced by a knockout animal that has
and chemical transepithelial gradients. The relationship between mildly increased water content in defecated stool (Wang et al.
extracellular Ca2+ concentration in the apical compartment and 2000). Similarly, little is known about paracellular intestinal wa-
rate of Ca2+ flux from apical to basolateral side of the epithelium ter movement. Although the tight junction protein claudin-2
reveals a saturable and a linear unsaturable component to total (Cldn2) has been implicated in mediating paracellular water flux
Biochem. Cell Biol. Downloaded from www.nrcresearchpress.com by San Francisco (UCSF) on 12/02/14

Ca2+ flux, likely reflecting transcellular and paracellular transport across a renal epithelial cell line (Rosenthal et al. 2010) and Cldn2 is
routes, respectively. The unsaturable component contributes sig- expressed in mouse intestine (Fujita et al. 2008), Cldn2 knockout
nificantly to total Ca2+ flux in all parts of the small and large mice do not display a phenotype consistent with decreased intes-
intestines (Karbach et al. 1986; Nellans 1990; Karbach 1992; Karbach tinal water absorption (Tamura et al. 2011).
and Feldmeier 1993). These data obtained in Ussings chambers are Regardless of the pathway, water movement passively follows
comparable to results obtained by the ligated loop technique de- the movement of ions (Curran and Macintosh 1962). Conse-
scribed previously. Ussing chamber experiments also permit the quently, water secretion has been attributed to electrogenic
measurement of Ca2+ flux in the presence of divergent voltage anion secretion through apical anion channels, whereas water
clamps, permitting the separation of Ca2+ flux into either voltage- absorption is coupled to epithelial Na+ uptake. The mecha-
dependent (presumably paracellular) or voltage-independent (pre- nisms mediating Na+ uptake can be divided into electroneutral
sumably transcellular flux) components (Frizzell and Schultz 1972). and electrogenic transport. Electrogenic transport occurs via api-
A detailed discussion of the merits of these different techniques is cal Na+ channels expressed mainly in the distal colon (Garty and
beyond the scope of this review. Palmer 1997) and Na+/nutrient-linked cotransporters, in particu-
Importantly, fluxes assumed to be “paracellular” by the afore- lar Na+-dependent glucose cotransporters (SGLTs) expressed in the
mentioned techniques are not always equivalent when measured small intestine. Electroneutral Na+ absorption occurs through the
in opposite directions (i.e., apical to basolateral vs. basolateral to apically expressed Na+/H+ exchangers, where NHE3 appears to
apical) even in the absence of an electrochemical gradient. This play the predominant role in Na+ absorption from both the small
For personal use only.

difference favors paracellular Ca2+ absorption (i.e., higher apical and large intestine (Schultheis et al. 1998; Gawenis et al. 2002). Na+
to basolateral than basolateral to apical flux) from the cecum efflux occurs across the basolateral membrane via the Na+/K+
and ascending colon and paracellular Ca2+ secretion from all ATPase for both pathways. The exact mechanism coupling water
segments of the small intestine and the descending colon. This net flux to sodium transport is still debated. Among multiple differ-
paracellular Ca2+ movement has been attributed to solvent drag ent theories, the revised standing gradient model (Diamond and
(Nellans 1990; Karbach and Feldmeier 1993; Rievaj et al. 2013). Bossert 1967) and the Na+ recirculation theory, which is summa-
Because water flux is assumed (but rarely measured under the rized in detail in Larsen and Mobjerg (2006) and Larsen et al.
same experimental conditions) to be in the apical to basolateral (2007), are probably the most developed. Both theories are based
direction, the term “anomalous solvent drag” was coined to ex- on the observation that Na+/K+ ATPase is predominantly localized
plain paracellular Ca2+ secretion from the small intestine (Nellans to the lateral membrane of intestinal epithelial cells (DiBona and
and Kimberg 1979). Mills 1979). This lateral localization of the Na+/K+ ATPase and the
However, data from Ussing chamber studies measure much vastly different resistance between the tight junction and the
higher fluxes in the basolateral to apical direction than what is basal junction would generate a hyperosmotic and hyperbaric
generally assumed from in vivo studies. This led to the reporting lateral intercellular space, which would provide the conditions
of net Ca2+ secretion in the jejunum and no net transport from the for fluid uptake in the absence of an electrochemical gradient or
duodenum when symmetrical Ca2+ concentrations were em- even against a concentration gradient. Additionally, the Na+ recir-
ployed in Ussing chamber experiments under voltage clamp (i.e., culation theory assumes influx of Na+ from the basolateral mem-
in the absence of an electrochemical gradient) (Walling and brane and hence “recirculation” of Na+. Although this would be
Kimberg 1973, 1975; Karbach and Feldmeier 1993; Rievaj et al. energetically unfavourable for a tissue, there is experimental
2013). Although net Ca2+ secretion has also been observed during evidence consistent with the existence of this phenomenon
in vivo studies, this always occurred down its electrochemical (Nedergaard et al. 1999).
gradient (Krawitt and Schedl 1968; Sernka and Borle 1969). We can Given the dependence of secondarily active paracellular intes-
only speculate about the reasons for the discrepancy in measure- tinal Ca2+ absorption on water flux, which is driven by intestinal
ment between methodologies, with possible explanations includ- Na+ absorption, it can be hypothesized that molecules implicit to
ing selective damage to water absorptive cells during prolonged in intestinal Na+ absorption would also play a role in intestinal Ca2+
vitro measurements (Inagaki et al. 2005; Inagaki-Tachibana et al. absorption. This was tested by removal of glucose from the muco-
2008a, 2008b), the presence of substances in vivo which prevent sal side of intestinal epithelium, thereby inhibiting Na+-coupled
anomalous solvent drag such as bile salts (Hu et al. 1993), or the glucose transport by depriving the cotransporter of substrate
presence of greater unstirred layers on the basolateral side of (Charoenphandhu et al. 2001). However, this method has been
epithelia leading to underestimates of apical to basolateral flux criticized due to its effect on the polarity of the apical membrane
in Ussing chamber studies, especially when performed on un- (Kellett 2011). To test whether electroneutral NHE3-mediated Na+
stripped epithelium. flux contributes to intestinal Ca2+ uptake, we measured intestinal
The other way to estimate the involvement of secondarily active 45Ca2+ uptake after gastric gavage in wild-type and Nhe3 knockout

paracellular Ca2+ absorption is by altering the driving force of mice (Pan et al. 2012). Importantly, Ca2+ uptake from the intestine
water movement. All intestinal segments demonstrate both water was reduced in Nhe3−/− mice. Direct measurements of Ca2+ flux
absorptive and water secretive properties. Under physiological across duodenum and cecum (the 2 parts of intestine with known
conditions, water absorption prevails. Water can move across an net Ca2+ absorption) in Ussing chambers confirmed decreased lu-
epithelium via either a transcellular or paracellular pathway. Al- minal to basolateral transepithelial Ca2+ flux (Pan et al. 2012;
though both mechanisms of water flux could generate a second- Rievaj et al. 2013), confirming that NHE3 contributes to paracellu-

Published by NRC Research Press


472 Biochem. Cell Biol. Vol. 92, 2014

lar Ca2+ uptake. Finally, to avoid potential compensatory changes ity of channels and transporters that govern transcellular Ca2+
resulting from an altered Ca2+ balance in Nhe3−/− mice, we con- reabsorption as recently reviewed in Fleet and Schoch (2010) and
firmed the findings using a NHE3 inhibitor in cecum of wild-type Christakos (2012). A role for 1,25-dihydroxyvitamin D3 in regulat-
mice (Rievaj et al. 2013). Together these data support a role for ing paracellular Ca2+ flux has been debated; however, recent evi-
NHE3 in transepithelial intestinal Ca2+ absorption, likely via driv- dence is consistent with 1,25-dihydroxyvitamin D3 also increasing
ing paracellular Ca2+ flux either via convection or by allowing paracellular Ca2+ flux (Wasserman 2004; Fujita et al. 2008). In-line
water removal and thereby generating a favourable concentration with a role for 1,25-dihydroxyvitamin D3 in altering paracellular
gradient for Ca2+ to follow. Ca2+ flux, the expression of Cldn2 and Cldn12 was reduced in
vitamin D–receptor knockout mice (Fujita et al. 2008). Moreover,
Molecules setting the paracellular permselectivity across treatment of the intestinal Caco-2 cell line with 1,25-dihydroxyvitamin
intestinal epithelia D3 increased Cldn2 and Cldn12 expression, and increased paracel-
For paracellular ion flux to occur, there must be both a driving lular Ca2+ flux (Fujita et al. 2008). Further, it has been speculated
Biochem. Cell Biol. Downloaded from www.nrcresearchpress.com by San Francisco (UCSF) on 12/02/14

force and a Ca2+-permeable pore. The discovery and investigation that increased 1,25-dihydroxyvitamin D3–responsive paracellular
of a family of tight junction proteins, termed claudins, has con- Ca2+ flux compensates, at least in part, for decreased transepithe-
tributed significantly to our understanding of paracellular pores. lial Ca2+ flux in TRPV6 and calbindin-D9K double knockout mice
Claudins are 4-pass transmembrane proteins with both the amino (Christakos et al. 2010). Although this discussion is consistent with
and carboxy termini resident in the cytosol. They form homo- the possibility of claudin pores contributing to regulated paracel-
meric and heteromeric interactions with one another in the same lular Ca2+ absorption from the intestine, much work still needs to
tight junction and in the tight junction of the adjacent epithelial be done to firmly establish this. This includes direct measure-
cell. Although the structure of this complex is poorly understood, ments of Ca2+ flux across the intestine of claudin-specific knock-
it clearly plays a role in determining the permeability of the tight out animals.
junction (for recent reviews see Gunzel and Fromm (2012) and
Gunzel and Yu (2013)). Active transcellular Ca2+ absorption occurs from both the
Evidence supports a role for Cldn2 and Cldn12 in creating duodenum and proximal large bowel
cation-permeable paracellular pores, hence permitting Ca2+ flux Active transcellular Ca2+ absorption is predominantly studied
between intestinal epithelial cells (Fujita et al. 2008). Cldn2, when in duodenum, whereas active paracellular Ca2+ absorption has
overexpressed in epithelial cell culture models, creates a paracel- only been observed in the cecum and proximal colon. The cecum
lular cation-selective pore permeable to Ca2+ (Furuse et al. 2001; and proximal colon are also sites of significant transcellular Ca2+
Amasheh et al. 2002; Fujita et al. 2008; Yu et al. 2009, 2010). Exam- absorption. As in the duodenum, the same transcellular Ca2+ ab-
For personal use only.

ination of the Cldn2 knockout mouse confirmed that it forms a sorption machinery is present and measurements in Ussing
paracellular cation-selective pore across the intestine and renal chambers have confirmed both significant transcellular and
proximal tubule (Muto et al. 2010; Tamura et al. 2011). However, secondarily active paracellular Ca2+ flux across these segments,
direct Ca2+ flux studies have not been reported from Cldn2−/− mice in rodents at least (Nellans and Goldsmith 1981; Karbach and
to date. Overexpression and siRNA-mediated knockdown of Cldn12 Feldmeier 1993; Kraidith et al. 2009; Zhang et al. 2010; Rievaj et al.
in intestinal epithelial cells demonstrate that this claudin also 2013). Consistent with a role of proximal large bowel in Ca2+ ho-
creates a cation-selective pore permeable to Ca2+ (Fujita et al. meostasis, resection of the cecum in rats leads to decreased bone
2008). However, the phenotype of a Cldn12 knockout animal has mineral density secondary to enhanced trabecular bone resorp-
not been reported. Overexpression studies also demonstrate the tion (Charoenphandhu et al. 2012; Jongwattanapisan et al. 2012).
ability for Cldn15 to form a paracellular cation-selective pore These results are not likely directly translatable to humans given
(Colegio et al. 2002; Van Itallie et al. 2003). Interestingly, deletion the relatively larger size of the cecum in rodents vs. the cecum in
of this gene in mice leads to the development of megaintestine humans (Casteleyn et al. 2010). The role of the proximal large
(Tamura et al. 2011). The knockout mice also display decreased bowel in Ca2+ homeostasis in humans has been relatively unex-
intestinal Na+ and water absorption, which appears exacerbated plored; we are aware of only a single study directly confirming
in the Cldn2 and Cldn15 double knockout (Wada et al. 2013). Given active transcellular flux from this segment after vitamin D3 ad-
these findings, it is tempting to speculate that Cldn15 also contrib- ministration (Grinstead et al. 1984). Kinetic studies suggest that
utes to the formation of cation-selective paracellular pores under physiological conditions, there is a small but measurable
permeable to Ca2+ along the course of the intestine. Again, direct contribution of colon/cecum to total Ca2+ absorption (Barger-Lux
measurements of Ca2+ flux across the intestine of Cldn15 knockout et al. 1989). Thus, the large intestine seems to provide a reserve,
mice have not been reported. which can be used in situations when absorption from the small
The localization of Cldn2, Cldn12, and Cldn15 has been exam- intestine is inadequate (Hylander et al. 1980, 1990; Abrams et al.
ined along the course of the mouse intestine (Fujita et al. 2006, 2007). Whether increased Ca2+ absorption from the cecum/colon
2008). Consistent with a role in paracellular transport, Cldn2 and plays a role in situations of increased Ca2+ requirements, such as
Cldn12 expression is greatest in the distal segments of the small childhood growth, pregnancy, lactation, or when Ca2+ intake is
bowel. This expression pattern differs slightly in rats; however, reduced, remains to be determined. Given the frequency of co-
significant expression of both isoforms is seen in the distal small lonic resection, it would be important to determine the effect of
bowel (Markov et al. 2010). Expression of Cldn15 in mice is greater this procedure on Ca2+ homeostasis in humans.
in more proximal intestinal segments, although there is signifi-
cant expression along the course of the bowel consistent with a
Kidney
possible role in paracellular Ca2+ reabsorption (Fujita et al. 2006). Ca2+ filtered by the kidney is reclaimed by the nephron to main-
The expression of CLDN12 and CLDN15 has been observed along the tain serum levels within normal limits. Like the intestine, the
course of small and large bowel in humans, although CLDN2 has majority of Ca2+ is reabsorbed via passive paracellular processes.
only been reported in the small bowel as summarized in Lu et al. The bulk, approximately 70% of filtered Ca2+, is reabsorbed in this
(2013). fashion from the proximal tubule and approximately 25% via pas-
sive paracellular fluxes from the thick ascending limb. The re-
Regulation of paracellular Ca2+ absorption in intestine mainder of Ca2+ reabsorption occurs via active transcellular flux
The calciotropic hormone 1,25-dihydroxyvitamin D3 acts to in- from the distal convoluted and connecting tubules. This latter
crease intestinal Ca2+ absorption. Much research has focused on process appears to occur via an analogous process in the duode-
how 1,25-dihydroxyvitamin D3 increases the expression and activ- num and cecum. Here, TRPV5 channels mediate apical uptake of

Published by NRC Research Press


Alexander et al. 473

Fig. 3. Proximal tubular Ca2+ reabsorption. The majority of Ca2+ flux across the proximal tubule occurs in a paracellular fashion, driven by
the transepithelial flux of sodium (Na+), which in turn drives water (H2O) flux. Water removal from the proximal tubular lumen could act to
concentrate Ca2+ providing a concentration gradient (bottom junction) or paracellular water movement may drive Ca2+ flux via convection
(i.e., solvent drag) (top junction). The epithelial Na+/H+ exchanger (NHE3) is the primary Na+ influx pathway, whose activity is driven by the
Na+/K+-ATPase (NaK ATPase). Note that two-thirds of water flux across the proximal tubule occurs in a transcellular fashion mediated by
aquaporin-1 pores (not depicted here). There is also significant paracellular Na+ flux.

CLDN
2
2+
Ca H O H2O
Biochem. Cell Biol. Downloaded from www.nrcresearchpress.com by San Francisco (UCSF) on 12/02/14

2
CLDN
2

Na+ Na+
NHE3
H+ Na+
NaK
ATPase
K+

CLDN
2
Ca2+ Ca2+
CLDN
For personal use only.

the ion, with calbindin-D28K buffering intracellular Ca2+ and ex- Dirks 1975). Recent studies have begun to dissect out the molecu-
trusion across the basolateral membrane occurs via the plasma lar components mediating passive paracellular flux of Ca2+ across
membrane Ca2+-dependent ATPase and the Na+/Ca2+ exchanger the proximal tubule and their findings are outlined subsequently.
(NCX) (Hoenderop et al. 2005; Dimke et al. 2010, 2011). The remain- The generation and characterization of Nhe3 knockout mice
der of the review will focus on the mechanisms mediating Ca2+ have implicated this transporter in the bulk of Na+ and, conse-
flux across the proximal tubule and thick ascending limb. quently, osmotically driven water flux across the proximal tubule
(and bowel as described previously) (Schultheis et al. 1998; Lorenz
Proximal tubule et al. 1999). Both microperfusion and micropuncture studies dem-
Most (approximately 55%–60%) Ca2+ filtered by the glomerulus
onstrate significantly reduced Na+ and water reabsorption from
is reabsorbed by the proximal convoluted tubule (Duarte and
the proximal tubule of Nhe3 knockout mice relative to wild-type
Watson 1967; Suki 1979). A further 10% of filtered Ca2+ is reab-
littermates (Schultheis et al. 1998). These data and the known
sorbed from the straight portion of the proximal tubule (Suki
dependence of proximal tubular Ca2+ reabsorption on Na+ and
1979; Seldin 1999). Therefore, most (i.e., at least two-thirds) of Ca2+
water reabsorption led us to examine whether NHE3 drives prox-
in the ultrafiltrate is reclaimed by the proximal tubule. Evidence
from micropuncture studies performed on multiple species is imal paracellular Ca2+ reabsorption. To this end, we looked at
consistent with Ca2+ uptake from this nephron segment occur- paracellular Ca2+ flux across a proximal tubular cell culture model
ring in a passive paracellular fashion (Duarte and Watson 1967; where NHE3 was overexpressed and found that this doubled Ca2+
Morel et al. 1969; Sutton and Dirks 1975). The ratio of Ca2+ concen- flux (Pan et al. 2012). That NHE3 activity per se was required for
tration in tubular fluid to the ultrafiltrate, [(TF/UF)Ca], is consis- this effect was confirmed by inhibiting Na+/H+ exchange with the
tently reported to be between 1.0 and 1.2. This ratio was also removal of Na+. Examination of the Nhe3 null mouse revealed
consistently similar, if not identical, to the ratio for Na+, inferring significantly increased fractional excretion of Ca2+ consistent
that active Na+ reabsorption provides the driving force for Ca2+ with decreased proximal Ca2+ reabsorption (Pan et al. 2012), al-
reabsorption. Consistent with this, attempts to uncouple proxi- though specific studies of Ca2+ flux across the proximal tubule
mal Na+ reabsorption from Ca2+ reabsorption by the administra- were not made. The proposed role of NHE3 in proximal tubular
tion of parathyroid hormone, furosemide, hydrochlorothiazide, Ca2+ reabsorption is depicted in Fig. 3. Ultimately, increased renal
or acetazolamide were unsuccessful (Agus et al. 1973; Beck and excretion of Ca2+ and decreased intestinal Ca2+ uptake led to
Goldberg 1973; Edwards et al. 1973; Sutton and Dirks 1975). That poorly mineralized bones in the knockout animals (Pan et al.
Ca2+ reabsorption from the proximal tubule occurs in a paracel- 2012).
lular fashion is consistent with microperfusion studies of the As observed in the intestine, passive paracellular Ca2+ flux de-
proximal tubule in the absence of an electrochemical gradient, pends not only on a driving force, but also on a “leaky” tight
which demonstrate predominant passive paracellular flux (Ng junction permeable to Ca2+. Given the known role and function of
et al. 1984). These authors did detect a small (<10%) active trans- the claudin family of tight junction proteins, some combination
cellular component as well. Consistent with this, the paracellular of claudins in the proximal tubule likely confer the leaky perme-
shunt in the proximal tubule is very permeable to Ca2+, thereby ability characteristics of this nephron segment permitting Ca2+
facilitating transport of the ion (Murayama et al. 1972; Sutton and flux. Thus far, only Cldn2, Cldn10a, and Cldn17 have been identi-

Published by NRC Research Press


474 Biochem. Cell Biol. Vol. 92, 2014

fied in the proximal tubule of adult animals (Enck et al. 2001; secretion of K+ by renal outer medullary potassium (ROMK) chan-
Kiuchi-Saishin et al. 2002; Gunzel et al. 2009b; Krug et al. 2012). nels and its subsequent recycling because K+ is limiting. The K+
Cldn2 has been studied extensively in cell culture models, where efflux back into the lumen provides part of the lumen-positive
it consistently forms a cation-selective pore (Furuse et al. 2001; potential difference. The inwardly transported Na+ and the 2 Cl–
Amasheh et al. 2002; Yu et al. 2009). Moreover, detailed microper- ions exit basolaterally via the Na+/K+-ATPase and CLC-Kb channels,
fusion studies of proximal tubules from Cldn2 knockout mice con- respectively. Collectively, these transport processes create a
firm the role of Cldn2 in generating a cation-selective pore across lumen-positive voltage gradient of approximately 5–10 mV
this tissue in vivo (Muto et al. 2010). Although direct measure- (Greger and Schlatter 1983; Greger and Velazquez 1987). In addi-
ments of Ca2+ flux across the proximal tubule of Cldn2 knockout tion to this, significant paracellular backflux of Na+ is thought to
mice have not been made, the lack of expression of this claudin occur when the transepithelial Na+ concentration is reversed be-
from more distal segments in kidney combined with data confirm- cause filtrate ascends toward the top of the TAL (i.e., when the
ing its selective permeability to cations and the presence of hyper- luminal Na+ concentration is lower than that of the interstitium).
Biochem. Cell Biol. Downloaded from www.nrcresearchpress.com by San Francisco (UCSF) on 12/02/14

calciuria in Cldn2 knockout mice (Muto et al. 2010) implicates this This backflux of interstitial Na+ generates a diffusion potential
isoform in mediating paracellular Ca2+ flux across the proximal tu- thought to amplify the voltage gradient to reach values approach-
bule. However, it should be noted that in Cldn2-deficient mice, the ing +30 mV (Rocha and Kokko 1973; Greger 1981; Mandon et al.
altered structure of the tight junction causes reduced reabsorption 1993; Hou et al. 2005). Consequently, the cortical segment of the
of Na+, Cl–, and water (Muto et al. 2010). Consequently it cannot be TAL appears to be the main site of paracellular Ca2+ transport
excluded that urinary Ca2+ loss in these animals is due to reduced (Suki et al. 1980; Di Stefano et al. 1990; De Rouffignac et al. 1991).
solvent drag and (or) diminished gradient formation because these The lumen-positive transepithelial voltage gradient is critical in
cannot be uncoupled. In fact, changes in Na+ and water transport establishing the driving force for Ca2+ transport across the TAL.
from the PT correlate with changes in Ca2+ transport (Agus et al. 1973; Furosemide, a pharmacological blocker of NKCC2, effectively dis-
Beck and Goldberg 1973; Edwards et al. 1973; Sutton and Dirks 1975). sipates the lumen-positive potential difference across the seg-
In cell culture, Cldn10a and Cldn17 form anion-selective pores ment, which potently inhibits transepithelial Ca2+ transport
(Van Itallie et al. 2006; Krug et al. 2012). Consequently, they are across the TAL segment (Burg 1976; Quamme 1981; Good et al.
unlikely to form Ca2+-selective pores in the proximal tubule. How- 1984; Di Stefano et al. 1993; Deschenes et al. 2001).
ever, altering the relative expression of these isoforms in propor- The paracellular pathway permitting transport of Ca2+ across
tion to the amount of Cldn2 or another cation-selective isoform the epithelium of the cortical TAL shows cation selectivity (Greger
may regulate paracellular Ca2+ flux across the proximal tubule. A 1981; Di Stefano et al. 1988). The selectivity of this shunt is likely
detailed examination of the expression of all claudins in the prox- determined by the combined interactions between specific clau-
For personal use only.

imal tubule has not been completed. Consequently, it is possible dins. Cldn16 (previously referred to as paracellin-1) was the first
that other cation-selective claudins may be expressed in this seg- claudin identified to play a role in determining the selectivity
ment and contribute to the formation of a paracellular Ca2+ pore. characteristics of the TAL tight junction. The CLDN16 gene was
Cldn12 expression in the proximal tubule has not been examined; identified as a causative gene responsible for the phenotype of
however, given that it is expressed in a proximal tubule cell cul- familial hypomagnesemia with hypercalciuria and nephrocalci-
ture model (Borovac et al. 2012) and its purported role in paracel- nosis (FHHNC) (Simon et al. 1999). Individuals with FHHNC pres-
lular Ca2+ flux across the intestine (Fujita et al. 2008), it is ent with severe hypomagnesemia due to renal wasting, which is
tempting to speculate that it may perform a similar function in not surprising because the TAL also takes up a large portion
the proximal tubule. (⬃60%) of filtered Mg2+ via the paracellular pathway (de
Early postnatally, the proximal tubule undergoes developmen- Rouffignac et al. 1983; Simon et al. 1999). In-line with a common
tal changes and the ability to reabsorb a number of substrates, paracellular shunt for Ca2+ and Mg2+ in the TAL, FHHNC patients
including Ca2+, increases with age (Lelievre-Pegorier et al. 1983). also display hypercalciuria and nephrocalcinosis, which likely
There is increasing NHE3 expression coincident with increased promotes renal insufficiency. However, FHHNC patients tend to
Ca2+ reabsorption (Becker et al. 2007), perhaps contributing to an maintain serum Ca2+ within normal limits (Michelis et al. 1972;
increased driving force. However, there are also changes in prox- Praga et al. 1995). Another group of FHHNC patients have been
imal tubule claudin expression. Specifically, there is significant described with a similar phenotype, but present with more severe
Cldn6 and Cldn9 expression postnatally, which decreases with age ocular symptoms (Konrad et al. 2006). Mutations in the CLDN19
and is absent in adult animals (Abuazza et al. 2006). Both of these gene were shown to be the molecular cause of their disease.
claudins form cation barriers (Sas et al. 2008). Perhaps their re- Importantly, both Cldn16 and Cldn19 localize to the TAL and
duced expression explains some of the increased ability of the distal convoluted tubule segments of the kidney (Simon et al.
proximal tubule to reabsorb Ca2+ with age. It is important to note 1999; Kiuchi-Saishin et al. 2002; Konrad et al. 2006). The majority
that proximal tubule Mg2+ reabsorption decreases concurrently of data on the ion selectivity of Cldn16 and Cldn19 are derived
during development, so the age-dependent alterations in perme- from cell model experiments. It should be noted that data ob-
ability characteristics likely reflect a more complicated change tained from these in vitro studies utilizing different cell model
within the paracellular pore (Lelievre-Pegorier et al. 1983). systems to express Cldn16 and Cldn19 differ substantially, as re-
viewed in detail by Gunzel and Yu (2008). These differences may
Thick ascending limb of Henle's loop be due to several factors, but are most likely explained by the cell
The thick ascending limb of Henle's loop (TAL) plays an impor- line utilized. Different cell types form monolayers with widely
tant role in the reabsorption of filtered Ca2+. Detailed micropunc- varying transepithelial resistances and differing cation/anion se-
ture measurements estimate that 20%–25% of Ca2+ in the filtrate is lectivity in their basal state (Ikari et al. 2004; Hou et al. 2005,
reclaimed by the TAL (Lassiter et al. 1963; Suki 1979). Ca2+ reab- Kausalya et al. 2006; Angelow et al. 2007; 2008a, 2008b; Gunzel
sorption across the TAL epithelium occurs mainly by passive para- et al. 2009a). Such variability is due, in part, to the endogenous
cellular transport. This is largely dependent on a lumen-positive expression of various claudin genes. Importantly, transfection of
transepithelial voltage gradient acting to drive Ca2+ through the an exogenous claudin can affect the expression of endogenous
paracellular shunt. This lumen-positive transepithelial voltage claudins (Gunzel et al. 2009b; Dimke et al. 2013). This may help
gradient is generated by 2 interdependent mechanisms as out- explain the conflicting data obtained when evaluating the effect
lined in Fig. 4. NKCC2-dependent furosemide-sensitive cotrans- of overexpression of Cldn16 or Cldn19 on divalent cation perme-
port drives 1 Na+, 1 K+, and 2 Cl– ions across the apical membrane. ation (Ikari et al. 2004; Hou et al. 2005, 2008a, 2008b; Kausalya
This inward transport of monovalent ions is reliant on the apical et al. 2006; Angelow et al. 2007; Gunzel et al. 2009a). In general, in

Published by NRC Research Press


Alexander et al. 475

Fig. 4. The Ca2+-sensing receptor claudin-14 axis. Top panel: Thick ascending limb Ca2+ reabsorption under conditions of normocalcemia.
Asymmetrical ion fluxes across the TAL generates a lumen-positive potential difference which drives paracellular Ca2+ flux. The paracellular
pore is created by claudin-16 (CLDN16) and claudin-19 (CLDN19). Bottom panel: When hypercalcemia ensues, more Ca2+ is available to activate
the Ca2+-sensing receptor (CaSR), which relays cellular signalling–increasing claudin-14 (CLDN14) expression. CLDN14 then localizes to the tight
junction in the TAL, where it blocks paracellular Ca2+ flux, thereby promoting calciuria. Importantly, CLDN14 also prevents the backflux of Na+.
NKCC2, furosemide-sensitive Na+, K+, 2Cl– cotransporter; ROMK, renal outer medulla K+ channel; ClC-Kb, Cl– channel Kb; NaK ATPase,
Na+/K+-ATPase.

Normocalcemia
CLDN CLDN
16 19
Na+
Biochem. Cell Biol. Downloaded from www.nrcresearchpress.com by San Francisco (UCSF) on 12/02/14

CLDN CLDN
Na+
16 19
Na+
NaK
ROMK
K+ ATPase
K+
CLCKB
Na+ Cl-
NKCC2
K+
2Cl- CaSR

CLDN CLDN
2+ 16 19
Ca
For personal use only.

CLDN CLDN

+ 16 19
Vte

Hypercalcemia
CLDN CLDN CLDN
16 19
2+ 14
Ca
CLDN CLDN CLDN Na+
14 16 19 +
Na
NaK
ATPase
ROMK
K+ K+

CLCKB
Na+ Cl-
NKCC2
K+
2Cl- CaSR Ca2+
CLDN CLDN CLDN
Ca2+ 14 16 19

CLDN CLDN CLDN


14 16 19
+ Vte

vitro overexpression studies found that divalent cation perme- ilar permeability characteristics to that of the TAL (Greger 1981; Di
ation was not as severely disturbed as would be expected from Stefano et al. 1988; Hou et al. 2008a). Based on these observations,
mutations causing FHHNC. Studies done utilizing the proximal it is likely that FHHNC mutations reduce either cation selectivity
tubular LLC-PK1 line found that Cldn16 expression increases the or increase anion permeation of the paracellular shunt in the TAL.
permeability for cations, especially for Na+ (Hou et al. 2005). In Both types of mutations would ultimately dissipate the backflux
contrast, the overexpression of Cldn19 leads to a reduction in Cl– of Na+ and, hence, prevent augmentation of the lumen positive
permeation, while leaving cation permeability unaltered (Hou diffusion voltage, thereby reducing the driving force for divalent
et al. 2008a). In LLC-PK1 cells, combined expression of the 2 clau- cation flux (Hou et al. 2005, 2008a). Critical to the validity of this
dins promotes the formation of a cation-selective pore, with sim- permeability model is the observation that FHHNC mutations in

Published by NRC Research Press


476 Biochem. Cell Biol. Vol. 92, 2014

Cldn16 or Cldn19, largely abolish the cation selectivity of Cldn16 in the lumen-positive voltage (Wittner et al. 1993). A phos-
or reduce the ability of Cldn19 to block anion permeation (Hou phorylation site in Cldn16 exists at Ser217. In Madin-Darby canine
et al. 2005, 2008a). kidney cells (MDCK), this site is a substrate for protein kinase A
Two transgenic mice models of Cldn16 deficiency have been (PKA). The phosphorylated form of Cldn16 localizes to the tight
generated, one utilized RNA interference to knockdown Cldn16 junction, whereas the dephosphorylated form is targeted for lys-
gene expression (Hou et al. 2007), whereas the other employed osomal degradation (Ikari et al. 2006). Thus, changes in cellular
standard targeted transgenic strategies for genomic deletion of cAMP production within the TAL may rapidly affect the localiza-
the gene (Will et al. 2010). Although variability exists between the tion and expression of CLDN16. Another key regulator of cAMP
2 animal models, both display hypercalciuria and hypomag- production in the TAL is the Ca2+-sensing receptor (CaSR), which
nesemia. It is of relevance to note that electrophysiological mea- can effectively inhibit up to 90% of hormone-stimulated cAMP
surements in microperfused tubules differ substantially between production (Desfleurs et al. 1998). The protein is located in the
the models. Knockdown of Cldn16 shows reduced cation selectivity basolateral membrane, where it responds to changes in systemic
Biochem. Cell Biol. Downloaded from www.nrcresearchpress.com by San Francisco (UCSF) on 12/02/14

of the paracellular shunt in the TAL segment (i.e., reduced PNa/PCl), Ca2+ concentrations and relays the signal to amend transport
whereas the permeability characteristics between different cat- across the TAL (Vargas-Poussou et al. 2002; Egbuna and Brown
ions remain unaffected (Hou et al. 2007). This is in contrast to 2008). Detailed physiological measurements of Ca2+ transport
Cldn16 knockout mice, which have an intact PNa/PCl ratio, but ex- across microperfused TAL tubules demonstrate that CaSR stimu-
hibit decreased PCa/PNa and PMg/PNa permeability ratios (Will et al. lation specifically inhibits Ca2+ flux across the segment, without
2010). In essence, the Cldn16 knockout mice have a reduced per- affecting the absorption of Na+ or Cl– and consequently the
meability to divalent cations over that of Na+, but in contrast to lumen-positive voltage (Desfleurs et al. 1998; Motoyama and
the Cldn16 knockdown mice, display no change in permeability of Friedman 2002; Loupy et al. 2012). In Cldn16-transfected MDCK
Na+ over that of Cl–. This could indicate that changes in these cells, CaSR activation decreases the amount of phosphorylated
claudins not only affect backflux of Na+ via the paracellular shunt Cldn16, diverting the protein to the lysosomal compartment,
in TAL, but may also directly affect the permeation of divalent likely due to reduced cAMP levels and decreased PKA activity
cations under certain conditions. In fact, the exact composition of (Ikari et al. 2008). Whether this occurs in vivo remains to be de-
claudins within an epithelium is critical in determining the per- termined. Other posttranslational mechanisms likely exist to
meability characteristics of the epithelium. With the realization acutely regulate TAL claudins and the paracellular permeability of
that the claudin composition of the TAL is difficult to mimic the segment.
within cell lines and that changes in the composition leads to Using thyroidectomized and parathyroidectomized rats re-
changes in the abundance of other claudins (Dimke et al. 2013; supplemented with l-thyroxine and PTH, Loupy et al. (2012) were
For personal use only.

Gunzel et al. 2009b), more studies are needed to definitively able to show that chronic CaSR antagonism lead to significantly
answer how the interaction between the different claudins in the higher renal tubular Ca2+ reabsorption, which was not dependent
TAL sets the permselectivity of the paracellular shunt in this seg- on circulating PTH levels. The effects were likely of renal origin
ment. because CaSR antagonism had the same effect on systemic Ca2+
Although it is likely that several other claudins in addition to concentrations after pretreatment with bisphosphonates. Simi-
Cldn16 and Cldn19 are expressed in the TAL, only the functions of larly, selective ablation of the Casr gene in kidney using the Six2-
Cldn10 and Cldn14 have been established (Breiderhoff et al. 2012; CRE driver line also showed significantly lower urinary excretion
Gong et al. 2012; Dimke et al. 2013; Gong and Hou 2014). Deletion of Ca2+ (Toka et al. 2012). Thus, CaSR expression in kidney plays an
of Cldn10 specifically from the TAL produces an interesting pheno- important role in regulating the urinary excretion of Ca2+. CaSR
type. The mice develop hypermagnesemia and increased Mg2+ activation has recently been shown to greatly stimulate the ex-
reabsorption, whereas Ca2+ balance is less affected. The mice pression of Cldn14 within the TAL (Gong et al. 2012; Dimke et al.
show lower serum Ca2+ concentrations and slightly reduced tubu- 2013; Gong and Hou 2014). A potential role for Cldn14 in regulat-
lar excretion of Ca2+. Notably, the mice develop nephrocalcinosis ing Ca2+ balance was originally indicated following the results of
in the form of extensive Ca2+ deposits in the renal medulla a large genome-wide association study. Here it was found that
(Breiderhoff et al. 2012). In isolated microperfused TAL segments, single nucleotide polymorphisms (SNPs) in the human CLDN14
a lack of Cldn10 promotes an increase in the lumen-positive volt- gene associated with kidney stones and lower bone mineral den-
age gradient, whereas the short circuit current remains unaf- sity (Thorleifsson et al. 2009). However, deleterious mutations in
fected, suggesting that only paracellular permeability is affected. the human CLDN14 gene were not associated with changes in Ca2+
In fact, a 4-fold decrease was observed in the permeability of Na+ balance; rather, these individuals presented with nonsyndromic
over that of Cl– (PNa/PCl). However, the permeability of divalent deafness (Wilcox et al. 2001). Thus, the mechanism responsible for
cations over that of Na+ was markedly increased, suggesting a the generation of kidney stones by SNPs in the CLDN14 gene re-
functional change within the permeability filter of the shunt mained elusive. Cldn14 was subsequently shown to localize to
(Breiderhoff et al. 2012). The reduced shunt permeability to Na+ the TAL and respond to changes in serum Ca2+ (Gong et al.
may be related to a 20-fold increase in the expression of the pore- 2012). In fact, both increased dietary Ca2+ intake and 1,25-
blocking Cldn14; however, this does not adequately explain the dihydroxyvitamin D3 stimulated hypercalcemia markedly increased
increased permeability to divalent cations over Na+ because expression of Cldn14 in kidney (Gong et al. 2012; Dimke et al. 2013).
Cldn14 blocks both divalent and monovalent cation flux (Gong This regulation was found to be dependent on the CaSR (Gong
et al. 2012; Dimke et al. 2013; Gong and Hou 2014). The regulation et al. 2012; Dimke et al. 2013). We found that chronic activation of
and function of Cldn14 is described in detail in the subsequent the CaSR using the calcimimetic, Cinacalcet, led to a 40-fold in-
sections. crease in Cldn14 gene expression (Dimke et al. 2013). However, the
Paracellular transport of divalent ions across the TAL is under effect of pharmacological activation of the CaSR is rapid, with
intense hormonal regulation. Thus, these mechanisms act to reg- 3-fold increased Cldn14 expression observed as early as 2 h after
ulate paracellular flux via acute and chronic pathways. Early stud- oral gavage (Gong and Hou 2014). Moreover, after genetic ablation
ies demonstrated that acute application of PTH and other of the Casr in kidney, the expression of cldn14 was significantly
calciotropic hormones elicited a cAMP-dependent increase in elec- reduced (Toka et al. 2012). In renal cell culture models, Cldn14
trolyte transport across the TAL (Di Stefano et al. 1990; Wittner physically interacts with Cldn16 (Gong et al. 2012). In LLC-PK1 cells,
et al. 1993). Notably, it was reported that PTH produced an acute opossum kidney cells, and MDCK cells, Cldn14/CLDN14 overexpres-
change in the permeability of the paracellular pathway because sion provided specific permeability characteristics to the monolayer.
the increased transport of divalent ions surpassed the change A decrease was observed in both the permeability of Na+ over that of

Published by NRC Research Press


Alexander et al. 477

Cl– (PNa/PCl) as well as the permeability of Ca2+ (PCa), suggesting that supports Dr. R.T. Alexander. The laboratory of H. Dimke is funded
Cldn14 acts as a nonselective cation barrier (Ben-Yosef et al. 2003; by the Novo Nordisk Foundation, the Carlsberg Foundation, and
Gong et al. 2012; Dimke et al. 2013). Based on these permeability the A.P. Møller Foundation for the Advancement of Medical
characteristics, it seems that Cldn14 functions to reduce both the Science.
backflux of Na+ and hence the lumen-positive voltage gradient in the
cortical TAL, in addition to blocking permeation of Ca2+ directly References
(Fig. 4). The regulation of Cldn14 by the CaSR might be critical in Abrams, S.A., Hawthorne, K.M., Aliu, O., Hicks, P.D., Chen, Z., and Griffin, I.J.
2007. An inulin-type fructan enhances calcium absorption primarily via an
conditions of hypercalcemia because the luminal Ca2+ concentration effect on colonic absorption in humans. J. Nutr. 137(10): 2208–2212. PMID:
in the TAL increases substantially and would otherwise allow unin- 17884999.
tended paracellular reabsorption of Ca2+, based on changes in the Abuazza, G., Becker, A., Williams, S.S., Chakravarty, S., Truong, H.T., Lin, F., and
chemical gradient (Dimke et al. 2013). Genetic models of altered Baum, M. 2006. Claudins 6, 9, and 13 are developmentally expressed renal
tight junction proteins. Am. J. Physiol. Renal Physiol. 291(6): F1132–F1141.
Cldn14 expression support these observations. In fact, when Cldn14- doi:10.1152/ajprenal.00063.2006. PMID:16774906.
Biochem. Cell Biol. Downloaded from www.nrcresearchpress.com by San Francisco (UCSF) on 12/02/14

deficient mice are placed on a high Ca2+ diet, the fractional excretion Agus, Z.S., Gardner, L.B., Beck, L.H., and Goldberg, M. 1973. Effects of parathyroid
of divalent cations are significantly reduced, in-line with a defect in hormone on renal tubular reabsorption of calcium, sodium, and phosphate.
blocking Ca2+ permeation of the shunt across the TAL (Gong et al. Am. J. Physiol. 224(5): 1143–1148. PMID:4349532.
Amasheh, S., Meiri, N., Gitter, A.H., Schoneberg, T., Mankertz, J., Schulzke, J.D.,
2012). Recently, a new mouse model was generated that overex- and Fromm, M. 2002. Claudin-2 expression induces cation-selective channels
pressed the Cldn14 protein specifically in the TAL driven by the in tight junctions of epithelial cells. J. Cell Sci. 115(Pt. 24): 4969–4976. doi:10.
Tamm-Horsfall promoter. In-line with previous publications, overex- 1242/jcs.00165. PMID:12432083.
pression of Cldn14 in the TAL markedly increased the fractional ex- Angelow, S., El-Husseini, R., Kanzawa, S.A., and Yu, A.S. 2007. Renal localization
and function of the tight junction protein, claudin-19. Am. J. Physiol. Renal
cretion of divalent cations (Gong and Hou 2014). Physiol. 293(1): F166–F177. doi:10.1152/ajprenal.00087.2007. PMID:17389678.
Bailey, R.L., Dodd, K.W., Goldman, J.A., Gahche, J.J., Dwyer, J.T., Moshfegh, A.J.,
Summary/future directions et al. 2010. Estimation of total usual calcium and vitamin D intakes in the
United States. J. Nutr. 140(4): 817–822. doi:10.3945/jn.109.118539. PMID:
Given that altered Ca2+ homeostasis can lead to cardiovascular 20181782.
and neurological complications in addition to changes in the den- Barger-Lux, M.J., Heaney, R.P., and Recker, R.R. 1989. Time course of calcium
sity of bone, understanding how Ca2+ is absorbed from the gut and absorption in humans: evidence for a colonic component. Calcif. Tissue Int.
renal tubule is prerequisite to adequately treating and preventing 44(5): 308–311. doi:10.1007/BF02556309. PMID:2496901.
Beck, L.H., and Goldberg, M. 1973. Effects of acetazolamide and parathyroidec-
these sequelae. A common manifestation of altered epithelial
tomy on renal transport of sodium, calcium, and phosphate. Am. J. Physiol.
Ca2+ handling is hypercalciuria and kidney stone formation. 224(5): 1136–1142. PMID:4700632.
Hypercalciuria may be caused by increased intestinal absorption
For personal use only.

Becker, A.M., Zhang, J., Goyal, S., Dwarakanath, V., Aronson, P.S., Moe, O.W., and
of Ca2+, an augmented release of Ca2+ from bone, by dysregulation Baum, M. 2007. Ontogeny of NHE8 in the rat proximal tubule. Am. J. Physiol.
of renal transport pathways, or a combination thereof. We have Renal Physiol. 293(1): F255–F261. PMID:17429030.
Ben-Yosef, T., Belyantseva, I.A., Saunders, T.L., Hughes, E.D., Kawamoto, K.,
recently acquired a preliminary understanding of how Ca2+ is Van Itallie, C.M., et al. 2003. Claudin 14 knockout mice, a model for auto-
reabsorbed from the TAL in normal physiological conditions; somal recessive deafness DFNB29, are deaf due to cochlear hair cell degener-
however, the extent to which this mechanism is perturbed in ation. Hum. Mol. Genet. 12(16): 2049–2061. doi:10.1093/hmg/ddg210. PMID:
kidney stone formers remains to be completely defined. Given the 12913076.
Bilezikian, J.P. 1993. Clinical review 51: Management of hypercalcemia. J. Clin.
genome-wide association study (GWAS) implicating CLDN14 in kid- Endocrinol. Metab. 77(6): 1445–1449. doi:10.1210/jcem.77.6.8263125. PMID:
ney stone formation, it seems reasonable to assume that those 8263125.
individuals somehow inappropriately upregulate CLDN14 expres- Birge, S.J., Peck, W.A., Berman, M., and Whedon, G.D. 1969. Study of calcium
sion. There is also much work to be done on the proximal tubule. absorption in man: a kinetic analysis and physiologic model. J. Clin. Invest.
48(9): 1705–1713. doi:10.1172/JCI106136. PMID:5822579.
Specifically, the claudins involved in Ca2+ reabsorption need to be Borovac, J., Barker, R.S., Rievaj, J., Rasmussen, A., Pan, W., Wevrick, R., and
identified and their potential regulation understood. To this end, Alexander, R.T. 2012. Claudin-4 forms a paracellular barrier, revealing the
micropuncture and (or) microperfusion studies examining the interdependence of claudin expression in the loose epithelial cell culture
flux of Ca2+ across the proximal tubule in claudin-specific knock- model opossum kidney cells. Am. J. Physiol. 303(12): C1278–C1291. doi:10.1152/
ajpcell.00434.2011. PMID:23076790.
out mice will be beneficial. Further, given the lack of direct mea- Breiderhoff, T., Himmerkus, N., Stuiver, M., Mutig, K., Will, C., Meij, I.C., et al.
surements of Ca2+ flux across the intestine of Cldn2 and Cldn15 2012. Deletion of claudin-10 (Cldn10) in the thick ascending limb impairs
knockout mice, this seems a logical area to explore in the short paracellular sodium permeability and leads to hypermagnesemia and
term. Moreover, the generation and characterization of a Cldn12 nephrocalcinosis. Proc. Natl. Acad. Sci. U.S.A. 109(35): 14241–14246. doi:10.
1073/pnas.1203834109. PMID:22891322.
knockout animal will provide clarity into its role in intestinal and Brine, C.L., and Johnston, F.A. 1955. Endogenous calcium in the feces of adult
renal tubular Ca2+ reabsorption. Finally, there remain many ques- man and the amount of calcium absorbed from food. Am. J. Clin. Nutr. 3(5):
tions with respect to paracellular Ca2+ flux across the bowel. Is 418–420. PMID:13258517.
there significant backflux (i.e., serosal to mucosal Ca2+ movement) Bronner, F. 1998. Calcium absorption–a paradigm for mineral absorption.
J. Nutr. 128(5): 917–920. PMID:9567004.
as part of the recirculation of Ca2+ across intestinal epithelium
Bronner, F. 2003. Mechanisms of intestinal calcium absorption. J. Cell. Biochem.
(e.g., between meals or when fasting)? If yes, what is the role of the 88(2): 387–393. doi:10.1002/jcb.10330. PMID:12520541.
cecum/colon in this process? How exactly is the paracellular path- Bronner, F., and Pansu, D. 1999. Nutritional aspects of calcium absorption.
way regulated? Is there any transcellular Ca2+ absorption from the J. Nutr. 129(1): 9–12. PMID:9915868.
ileum? What is the role of solvent drag in paracellular Ca2+ move- Bronner, F., Pansu, D., and Stein, W.D. 1986. An analysis of intestinal calcium
transport across the rat intestine. Am. J. Physiol. 250(5): G561–G569. PMID:
ment in vivo? What is the effect of proximal large bowel resection/ 2939728.
disease on Ca2+ homeostasis in humans? Answers to these questions Burg, M.B. 1976. Tubular chloride transport and the mode of action of some
will aid in maintaining a positive Ca2+ balance in both health and diuretics. Kidney Int. 9(2): 189–197. doi:10.1038/ki.1976.20. PMID:940261.
disease. Casteleyn, C., Rekecki, A., Van der Aa, A., Simoens, P., and Van den Broeck, W.
2010. Surface area assessment of the murine intestinal tract as a prerequisite
for oral dose translation from mouse to man. Lab. Anim. 44(3): 176–183.
Acknowledgements doi:10.1258/la.2009.009112. PMID:20007641.
The laboratory of R.T. Alexander is supported by funding from Charoenphandhu, N., Limlomwongse, L., and Krishnamra, N. 2001. Prolactin
the Kidney Foundation of Canada, the Women and Children's directly stimulates transcellular active calcium transport in the duodenum
of female rats. Can. J. Physiol. Pharmacol. 79(5): 430–438. doi:10.1139/y01-014.
Health Research Institute, and the Canadian Institute of Health PMID:11405247.
Research (CIHR). A Clinician Scientist Award from CIHR and an Charoenphandhu, N., Tudpor, K., Pulsook, N., and Krishnamra, N. 2006. Chronic
Alberta Innovates Health Solutions Clinical Investigator Award metabolic acidosis stimulated transcellular and solvent drag-induced cal-

Published by NRC Research Press


478 Biochem. Cell Biol. Vol. 92, 2014

cium transport in the duodenum of female rats. Am. J. Physiol. Gastrointest. Egbuna, O.I., and Brown, E.M. 2008. Hypercalcaemic and hypocalcaemic condi-
Liver Physiol. 291(3): G446–G455. doi:10.1152/ajpgi.00108.2006. PMID:16675746. tions due to calcium-sensing receptor mutations. Best Pract. Res. Clin. Rheu-
Charoenphandhu, N., Suntornsaratoon, P., Jongwattanapisan, P., Wongdee, K., matol. 22(1): 129–148. doi:10.1016/j.berh.2007.11.006. PMID:18328986.
and Krishnamra, N. 2012. Enhanced trabecular bone resorption and micro- Enck, A.H., Berger, U.V., and Yu, A.S. 2001. Claudin-2 is selectively expressed in
structural bone changes in rats after removal of the cecum. Am. J. Physiol. proximal nephron in mouse kidney. Am. J. Physiol. 281(5): F966–F974. PMID:
Endocrinol. Metab. 303(8): E1069–E1075. doi:10.1152/ajpendo.00242.2012. 11592954.
PMID:22912366. Fleet, J.C., and Schoch, R.D. 2010. Molecular mechanisms for regulation of intes-
Christakos, S. 2012. Recent advances in our understanding of 1,25-dihy- tinal calcium absorption by vitamin D and other factors. Crit. Rev. Clin. Lab.
droxyvitamin D3 regulation of intestinal calcium absorption. Arch. Biochem. Sci. 47(4): 181–195. doi:10.3109/10408363.2010.536429. PMID:21182397.
Biophys. 523(1): 73–76. doi:10.1016/j.abb.2011.12.020. PMID:22230327. Frizzell, R.A., and Schultz, S.G. 1972. Ionic conductances of extracellular shunt
Christakos, S., Dhawan, P., Ajibade, D., Benn, B.S., Feng, J., and Joshi, S.S. 2010. pathway in rabbit ileum. Influence of shunt on transmural sodium transport
Mechanisms involved in vitamin D mediated intestinal calcium absorption and electrical potential differences. J. Gen. Physiol. 59(3): 318–346. doi:10.
and in non-classical actions of vitamin D. J. Steroid Biochem. Mol. Biol. 121(1–2): 1085/jgp.59.3.318. PMID:5058963.
183–187. doi:10.1016/j.jsbmb.2010.03.005. PMID:20214989. Fujita, H., Chiba, H., Yokozaki, H., Sakai, N., Sugimoto, K., Wada, T., et al. 2006.
Clarke, L.L. 2009. A guide to Ussing chamber studies of mouse intestine. Am. J. Differential expression and subcellular localization of claudin-7, -8, -12, -13,
Biochem. Cell Biol. Downloaded from www.nrcresearchpress.com by San Francisco (UCSF) on 12/02/14

Physiol. Gastrointest. Liver Physiol. 296(6): G1151–G1166. doi:10.1152/ajpgi. and -15 along the mouse intestine. J. Histochem. Cytochem. 54(8): 933–944.
90649.2008. PMID:19342508. doi:10.1369/jhc.6A6944.2006. PMID:16651389.
Coe, F.L., Evan, A., and Worcester, E. 2005. Kidney stone disease. J. Clin. Invest. Fujita, H., Sugimoto, K., Inatomi, S., Maeda, T., Osanai, M., Uchiyama, Y., et al.
115(10): 2598–2608. doi:10.1172/JCI26662. PMID:16200192. 2008. Tight junction proteins claudin-2 and -12 are critical for vitamin
Colegio, O.R., Van Itallie, C.M., McCrea, H.J., Rahner, C., and Anderson, J.M. 2002. D-dependent Ca2+ absorption between enterocytes. Mol. Biol. Cell, 19(5):
Claudins create charge-selective channels in the paracellular pathway be- 1912–1921. doi:10.1091/mbc.E07-09-0973. PMID:18287530.
tween epithelial cells. Am. J. Physiol. Cell Physiol. 283(1): C142–C147. doi:10. Furuse, M., Furuse, K., Sasaki, H., and Tsukita, S. 2001. Conversion of zonulae
1152/ajpcell.00038.2002. PMID:12055082. occludentes from tight to leaky strand type by introducing claudin-2 into
Cramer, C.F. 1964. In Vivo Intestinal Transport of Calcium and Water from Madin-Darby canine kidney I cells. J. Cell Biol. 153(2): 263–272. doi:10.1083/
Solutions Recycled through Healed Cut Loops in Dogs. J. Nutr. 84: 118–124. jcb.153.2.263. PMID:11309408.
PMID:14236909. Garty, H., and Palmer, L.G. 1997. Epithelial sodium channels: function, struc-
Cramer, C.F. 1965. Sites of Calcium Absorption and the Calcium Concentration ture, and regulation. Physiol. Rev. 77(2): 359–396. PMID:9114818.
of Gut Contents in the Dog. Can. J. Physiol. Pharmacol. 43: 75–78. doi:10.1139/
Gawenis, L.R., Stien, X., Shull, G.E., Schultheis, P.J., Woo, A.L., Walker, N.M., and
y65-009. PMID:14324234.
Clarke, L.L. 2002. Intestinal NaCl transport in NHE2 and NHE3 knockout
Cramer, C.F., and Copp, D.H. 1959. Progress and rate of absorption of radiostron- mice. Am. J. Physiol. Gastrointest. Liver Physiol. 282(5): G776–G784. doi:10.
tium through intestinal tracts of rats. Proc. Soc. Exp. Biol. Med. 102: 514–517. 1152/ajpgi.00297.2001. PMID:11960774.
doi:10.3181/00379727-102-25301. PMID:13812617.
Geall, M.G., and Summerskill, W.H. 1969. Electric-potential difference–a ne-
Curran, P.F., and Macintosh, J.R. 1962. A model system for biological water glected parameter of gut integrity and function? Gut, 10(6): 418–421. doi:10.
transport. Nature, 193: 347–348. doi:10.1038/193347a0. PMID:13882705. 1136/gut.10.6.418. PMID:4891737.
de Rouffignac, C., Corman, B., and Roinel, N. 1983. Stimulation by antidiuretic Geall, M.G., Spencer, R.J., and Phillips, S.F. 1969. Transmural electrical potential
For personal use only.

hormone of electrolyte tubular reabsorption in rat kidney. Am. J. Physiol. difference of the human colon. Gut, 10(11): 921–923. doi:10.1136/gut.10.11.921.
Renal Physiol. 244(2): F156–F164. PMID:6824079. PMID:5358583.
De Rouffignac, C., Di Stefano, A., Wittner, M., Roinel, N., and Elalouf, J.M. 1991. Gong, Y., and Hou, J. 2014. Claudin-14 Underlies Ca++-Sensing Receptor-Mediated
Consequences of differential effects of ADH and other peptide hormones on
Ca++ Metabolism via NFAT-microRNA-Based Mechanisms. J. Am. Soc. Neph-
thick ascending limb of mammalian kidney. Am. J. Physiol. Regul. Integr.
rol. 25(4): 745–760. doi:10.1681/ASN.2013050553. PMID:24335970.
Comp. Physiol. 260(6): R1023–R1035. PMID:2058731.
Gong, Y., Renigunta, V., Himmerkus, N., Zhang, J., Renigunta, A., Bleich, M., and
Deschenes, G., Wittner, M., Stefano, A.D., Jounier, S., and Doucet, A. 2001. Col-
Hou, J. 2012. Claudin-14 regulates renal Ca++ transport in response to CaSR
lecting Duct Is a Site of Sodium Retention in PAN Nephrosis: A Rationale for
signalling via a novel microRNA pathway. EMBO J. 31(8): 1999–2012. doi:10.
Amiloride Therapy. J. Am. Soc. Nephrol. 12(3): 598–601. PMID:11181809.
1038/emboj.2012.49. PMID:22373575.
Desfleurs, E., Wittner, M., Simeone, S., Pajaud, S., Moine, G., Rajerison, R., and
Good, D.W., Knepper, M.A., and Burg, M.B. 1984. Ammonia and bicarbonate
Di Stefano, A. 1998. Calcium-sensing receptor: regulation of electrolyte trans-
transport by thick ascending limb of rat kidney. Am. J. Physiol. 247(1): F35–
port in the thick ascending limb of Henle's loop. Kidney Blood Press. Res.
F44. PMID:6742203.
21(6): 401–412. doi:10.1159/000025892. PMID:9933824.
Greger, R. 1981. Cation selectivity of the isolated perfused cortical thick ascend-
Di Stefano, A., Wittner, M., Gebler, B., and Greger, R. 1988. Increased Ca++ or
ing limb of Henle's loop of rabbit kidney. Pflugers Arch. 390(1): 30–37. doi:
Mg++ concentration reduces relative tight-junction permeability to Na+ in
10.1007/BF00582707. PMID:7195550.
the cortical thick ascending limb of Henle's loop of rabbit kidney. Ren.
Greger, R., and Schlatter, E. 1983. Properties of the lumen membrane of the
Physiol. Biochem. 11(1–2): 70–79. PMID:3249835.
cortical thick ascending limb of Henle's loop of rabbit kidney. Pflugers Arch.
Di Stefano, A., Wittner, M., Nitschke, R., Braitsch, R., Greger, R., Bailly, C., et al.
396(4): 315–324. doi:10.1007/BF01063937. PMID:6844136.
1990. Effects of parathyroid hormone and calcitonin on Na+, Cl−, K+, Mg2+ and
Ca2+ transport in cortical and medullary thick ascending limbs of mouse Greger, R., and Velazquez, H. 1987. The cortical thick ascending limb and early
kidney. Pflugers Arch. 417(2): 161–167. doi:10.1007/BF00370694. PMID:2084613. distal convoluted tubule in the urinary concentrating mechanism. Kidney
Int. 31(2): 590–596. doi:10.1038/ki.1987.39. PMID:3550228.
Di Stefano, A., Roinel, N., de Rouffignac, C., and Wittner, M. 1993. Transepithe-
lial Ca2+ and Mg2+ transport in the cortical thick ascending limb of Henle's Grinstead, W.C., Pak, C.Y., and Krejs, G.J. 1984. Effect of 1,25-dihydroxyvitamin
loop of the mouse is a voltage-dependent process. Ren. Physiol. Biochem. D3 on calcium absorption in the colon of healthy humans. Am. J. Physiol.
16(4): 157–166. doi:10.1159/000173762. PMID:7689239. 247(2): G189–G192. PMID:6547811.
Diamond, J.M., and Bossert, W.H. 1967. Standing-gradient osmotic flow. A mech- Guise, T.A., and Mundy, G.R. 1995. Clinical review 69: Evaluation of hypocalce-
anism for coupling of water and solute transport in epithelia. J. Gen. Physiol. mia in children and adults. J. Clin. Endocrinol. Metab. 80(5): 1473–1478. doi:
50(8): 2061–2083. doi:10.1085/jgp.50.8.2061. PMID:6066064. 10.1210/jcem.80.5.7744987. PMID:7744987.
DiBona, D.R., and Mills, J.W. 1979. Distribution of Na+-pump sites in transporting Gunzel, D., and Fromm, M. 2012. Claudins and other tight junction proteins.
epithelia. Fed. Proc. 38(2): 134–143. PMID:216589. Compr. Physiol. 2(3): 1819–1852. doi:10.1002/cphy.c110045. PMID:23723025.
Dimke, H., Hoenderop, J.G., and Bindels, R.J. 2010. Hereditary tubular transport Gunzel, D., and Yu, A.S. 2008. Function and regulation of claudins in the thick
disorders: implications for renal handling of Ca2+ and Mg2+. Clin. Sci. (Lond.), ascending limb of Henle. Pflugers Arch. 458(1): 77–88. doi:10.1007/s00424-008-
118(1): 1–18. doi:10.1042/CS20090086. PMID:19780717. 0589-z. PMID:18795318.
Dimke, H., Hoenderop, J.G., and Bindels, R.J. 2011. Molecular basis of epithelial Gunzel, D., and Yu, A.S. 2013. Claudins and the modulation of tight junction
Ca2+ and Mg2+ transport: insights from the TRP channel family. J. Physiol. permeability. Physiol. Rev. 93(2): 525–569. doi:10.1152/physrev.00019.2012.
589(Pt. 7): 1535–1542. doi:10.1113/jphysiol.2010.199869. PMID:21041532. PMID:23589827.
Dimke, H., Desai, P., Borovac, J., Lau, A., Pan, W., and Alexander, R.T. 2013. Gunzel, D., Amasheh, S., Pfaffenbach, S., Richter, J.F., Kausalya, P.J.,
Activation of the Ca2+-Sensing Receptor Increases Renal Claudin-14 Expres- Hunziker, W., and Fromm, M. 2009a. Claudin-16 affects transcellular Cl- se-
sion and Urinary Ca2+ Excretion. Am. J. Physiol. Renal Physiol. 304(6): F761– cretion in MDCK cells. J. Physiol. 587(Pt. 15): 3777–3793. doi:10.1113/jphysiol.
F769. doi:10.1152/ajprenal.00263.2012. PMID:23283989. 2009.173401. PMID:19528248.
Duarte, C.G., and Watson, J.F. 1967. Calcium reabsorption in proximal tubule of Gunzel, D., Stuiver, M., Kausalya, P.J., Haisch, L., Krug, S.M., Rosenthal, R., et al.
the dog nephron. Am. J. Physiol. 212(6): 1355–1360. PMID:4952126. 2009b. Claudin-10 exists in six alternatively spliced isoforms that exhibit
Duflos, C., Bellaton, C., Pansu, D., and Bronner, F. 1995. Calcium solubility, distinct localization and function. J. Cell Sci. 122(Pt. 10): 1507–1517. doi:10.
intestinal sojourn time and paracellular permeability codetermine passive 1242/jcs.040113. PMID:19383724.
calcium absorption in rats. J. Nutr. 125(9): 2348–2355. PMID:7666252. Hoenderop, J.G., Nilius, B., and Bindels, R.J. 2005. Calcium absorption across
Edwards, B.R., Baer, P.G., Sutton, R.A., and Dirks, J.H. 1973. Micropuncture study epithelia. Physiol. Rev. 85(1): 373–422. doi:10.1152/physrev.00003.2004. PMID:
of diuretic effects on sodium and calcium reabsorption in the dog nephron. 15618484.
J. Clin. Invest. 52(10): 2418–2427. doi:10.1172/JCI107432. PMID:4729040. Hou, J., Paul, D.L., and Goodenough, D.A. 2005. Paracellin-1 and the modulation

Published by NRC Research Press


Alexander et al. 479

of ion selectivity of tight junctions. J. Cell Sci. 118(Pt. 21): 5109–5118. doi:10. et al. 2012. Claudin-17 forms tight junction channels with distinct anion
1242/jcs.02631. PMID:16234325. selectivity. Cell. Mol. Life Sci. 69(16): 2765–2778. doi:10.1007/s00018-012-
Hou, J., Shan, Q., Wang, T., Gomes, A.S., Yan, Q., Paul, D.L., et al. 2007. Transgenic 0949-x. PMID:22402829.
RNAi depletion of claudin-16 and the renal handling of magnesium. J. Biol. Laforenza, U. 2012. Water channel proteins in the gastrointestinal tract.
Chem. 282(23): 17114–17122. doi:10.1074/jbc.M700632200. PMID:17442678. Mol. Aspects Med. 33(5–6): 642–650. doi:10.1016/j.mam.2012.03.001. PMID:
Hou, J., Renigunta, A., Konrad, M., Gomes, A.S., Schneeberger, E.E., Paul, D.L., 22465691.
et al. 2008. Claudin-16 and claudin-19 interact and form a cation-selective Larsen, E.H., and Mobjerg, N. 2006. Na+ recirculation and isosmotic transport.
tight junction complex. J. Clin. Invest. 118(2): 619–628. doi:10.1172/JCI33970. J. Membr. Biol. 212(1): 1–15. doi:10.1007/s00232-006-0864-x. PMID:17206513.
PMID:18188451. Larsen, E.H., Nedergaard, S., and Ussing, H.H. 2000. Role of lateral intercellular
Hu, M.S., Kayne, L.H., Willsey, P.A., Koteva, A.B., Jamgotchian, N., and Lee, D.B. space and sodium recirculation for isotonic transport in leaky epithelia. Rev.
1993. Bile salts and ileal calcium transport in rats: a neglected factor in Physiol. Biochem. Pharmacol. 141: 153–212. doi:10.1007/BFb0119579. PMID:
intestinal calcium absorption. Am. J. Physiol. 264(2): G319–G324. PMID: 10916425.
8447415. Larsen, E.H., Mobjerg, N., and Nielsen, R. 2007. Application of the Na+ recircula-
Hylander, E., Ladefoged, K., and Jarnum, S. 1980. The importance of the colon in tion theory to ion coupled water transport in low- and high resistance osmo-
calcium absorption following small-intestinal resection. Scand. J. Gastroen- regulatory epithelia. Comp. Biochem. Physiol. A Mol. Integr. Physiol. 148(1):
Biochem. Cell Biol. Downloaded from www.nrcresearchpress.com by San Francisco (UCSF) on 12/02/14

terol. 15(1): 55–60. doi:10.3109/00365528009181432. PMID:7367822. 101–116. doi:10.1016/j.cbpa.2006.12.039. PMID:17303459.


Hylander, E., Ladefoged, K., and Jarnum, S. 1990. Calcium absorption after intes- Lassiter, W.E., Gottschalk, C.W., and Mylle, M. 1963. Micropuncture study of
tinal resection. The importance of a preserved colon. Scand. J. Gastroenterol. renal tubular reabsorption of calcium in normal rodents. Am. J. Physiol.
25(7): 705–710. doi:10.3109/00365529008997596. PMID:2396084. 204(5): 771–775.
Ikari, A., Hirai, N., Shiroma, M., Harada, H., Sakai, H., Hayashi, H., et al. 2004. Lelievre-Pegorier, M., Merlet-Benichou, C., Roinel, N., and de Rouffignac, C. 1983.
Association of paracellin-1 with ZO-1 augments the reabsorption of divalent Developmental pattern of water and electrolyte transport in rat superficial
cations in renal epithelial cells. J. Biol. Chem. 279(52): 54826–54832. doi:10. nephrons. Am. J. Physiol. 245(1): F15–F21. PMID:6869534.
1074/jbc.M406331200. PMID:15496416. Levy, F.L., Adams-Huet, B., and Pak, C.Y. 1995. Ambulatory evaluation of neph-
Ikari, A., Matsumoto, S., Harada, H., Takagi, K., Hayashi, H., Suzuki, Y., et al. rolithiasis: an update of a 1980 protocol. Am. J. Med. 98(1): 50–59. doi:10.1016/
2006. Phosphorylation of paracellin-1 at Ser217 by protein kinase A is essen- S0002-9343(99)80080-1. PMID:7825619.
tial for localization in tight junctions. J. Cell Sci. 119(Pt. 9): 1781–1789. doi:10. Lorenz, J.N., Schultheis, P.J., Traynor, T., Shull, G.E., and Schnermann, J. 1999.
1242/jcs.02901. PMID:16608877. Micropuncture analysis of single-nephron function in NHE3-deficient mice.
Ikari, A., Okude, C., Sawada, H., Sasaki, Y., Yamazaki, Y., Sugatani, J., et al. 2008. Am. J. Physiol. 277(3): F447–F453. PMID:10484528.
Activation of a polyvalent cation-sensing receptor decreases magnesium
Loupy, A., Ramakrishnan, S.K., Wootla, B., Chambrey, R., de la Faille, R.,
transport via claudin-16. Biochim. Biophys. Acta, 1778(1): 283–290. doi:10.1016/
Bourgeois, S., et al. 2012. PTH-independent regulation of blood calcium con-
j.bbamem.2007.10.002. PMID:17976367.
centration by the calcium-sensing receptor. J. Clin. Invest. 122(9): 3355–3367.
Inagaki, E., Natori, Y., Ohgishi, Y., Hayashi, H., and Suzuki, Y. 2005. Segmental
doi:10.1172/JCI57407. PMID:22886306.
difference of mucosal damage along the length of a mouse small intestine in
Lu, Z., Ding, L., Lu, Q., and Chen, Y.H. 2013. Claudins in intestines: Distribution
an Ussing chamber. J. Nutr. Sci. Vitaminol. 51(6): 406–412. doi:10.3177/jnsv.
and functional significance in health and diseases. Tissue Barriers, 1(3):
51.406. PMID:16521699.
e24978. doi:10.4161/tisb.24978. PMID:24478939.
For personal use only.

Inagaki-Tachibana, E., Hayashi, H., and Suzuki, Y. 2008a. The electrophysiolog-


Mandon, B., Siga, E., Roinel, N., and de Rouffignac, C. 1993. Ca2+, Mg2+ and K+
ical properties of the mucosal barrier in the injured mouse jejunum in Ussing
transport in the cortical and medullary thick ascending limb of the rat
chambers. J. Nutr. Sci. Vitaminol. 54(3): 269–271. doi:10.3177/jnsv.54.269.
PMID:18635917. nephron: influence of transepithelial voltage. Pflugers Arch. 424(5–6): 558–
560. doi:10.1007/BF00374924. PMID:8255743.
Inagaki-Tachibana, E., Natori, Y., Hayashi, H., and Suzuki, Y. 2008b. In vitro
diffusion barriers of the mouse jejunum in Ussing chambers. J. Nutr. Sci. Marcus, C.S., and Lengemann, F.W. 1962a. Absorption of Ca45 and Sr85 from solid
Vitaminol. 54(1): 30–38. doi:10.3177/jnsv.54.30. PMID:18388405. and liquid food at various levels of the alimentary tract of the rat. J. Nutr. 77:
155–160. PMID:14469699.
Jongwattanapisan, P., Suntornsaratoon, P., Wongdee, K., Dorkkam, N.,
Krishnamra, N., and Charoenphandhu, N. 2012. Impaired body calcium Marcus, C.S., and Lengemann, F.W. 1962b. Use of radioyttrium to study food
metabolism with low bone density and compensatory colonic calcium ab- movement in the small intestine of the rat. J. Nutr. 76: 179–182. PMID:
sorption in cecectomized rats. Am. J. Physiol. Endocrinol. Metab. 302(7): 14469700.
E852–E863. doi:10.1152/ajpendo.00503.2011. PMID:22275757. Markov, A.G., Veshnyakova, A., Fromm, M., Amasheh, M., and Amasheh, S. 2010.
Karbach, U. 1992. Paracellular calcium transport across the small intestine. Segmental expression of claudin proteins correlates with tight junction bar-
J. Nutr. 122(3 Suppl.): 672–677. PMID:1542029. rier properties in rat intestine. J. Comp. Physiol. B, 180(4): 591–598. doi:10.
Karbach, U., and Feldmeier, H. 1993. The cecum is the site with the highest 1007/s00360-009-0440-7. PMID:20049600.
calcium absorption in rat intestine. Dig. Dis. Sci. 38(10): 1815–1824. doi:10. McCormick, C.C. 2002. Passive diffusion does not play a major role in the ab-
1007/BF01296104. PMID:8404402. sorption of dietary calcium in normal adults. J. Nutr. 132(11): 3428–3430.
Karbach, U., and Rummel, W. 1987. Calcium transport across the colon ascen- PMID:12421863.
dens and the influence of 1,25-dihydroxyvitamin D3 and dexamethasone. Michelis, M.F., Drash, A.L., Linarelli, L.G., De Rubertis, F.R., and Davis, B.B. 1972.
Eur. J. Clin. Invest. 17(4): 368–374. doi:10.1111/j.1365-2362.1987.tb02202.x. Decreased bicarbonate threshold and renal magnesium wasting in a sibship
PMID:3117572. with distal renal tubular acidosis (Evaluation of the pathophysiological role
Karbach, U., Bridges, R.J., and Rummel, W. 1986. The role of the paracellular of parathyroid hormone). Metabolism, 21(10): 905–920. doi:10.1016/0026-
pathway in the net transport of calcium across the colonic mucosa. Naunyn 0495(72)90025-X. PMID:5071957.
Schmiedebergs Arch. Pharmacol. 334(4): 525–530. doi:10.1007/BF00569396. Morel, F., Roinel, N., and Le Grimellec, C. 1969. Electron probe analysis of tubular
PMID:3102979. fluid composition. Nephron, 6(3): 350–364. doi:10.1159/000179738. PMID:
Kausalya, P.J., Amasheh, S., Gunzel, D., Wurps, H., Muller, D., Fromm, M., and 4892997.
Hunziker, W. 2006. Disease-associated mutations affect intracellular traffic Motoyama, H.I., and Friedman, P.A. 2002. Calcium-sensing receptor regulation
and paracellular Mg2+ transport function of Claudin-16. J. Clin. Invest. 116(4): of PTH-dependent calcium absorption by mouse cortical ascending limbs.
878–891. doi:10.1172/JCI26323. PMID:16528408. Am. J. Physiol. Renal Physiol. 283(3): F399–F406. doi:10.1152/ajprenal.00346.
Kellett, G.L. 2011. Alternative perspective on intestinal calcium absorption: pro- 2001. PMID:12167589.
posed complementary actions of Ca(v)1.3 and TRPV6. Nutr. Rev. 69(7): 347– Murayama, Y., Morel, F., and Le Grimellec, C. 1972. Phosphate, calcium and
370. doi:10.1111/j.1753-4887.2011.00395.x. PMID:21729089. magnesium transfers in proximal tubules and loops of Henle, as measured by
Kiuchi-Saishin, Y., Gotoh, S., Furuse, M., Takasuga, A., Tano, Y., and Tsukita, S. single nephron microperfusion experiments in the rat. Pflugers Arch. 333(1):
2002. Differential expression patterns of claudins, tight junction membrane 1–16. doi:10.1007/BF00586037. PMID:5064074.
proteins, in mouse nephron segments. J. Am. Soc. Nephrol. 13(4): 875–886. Muto, S., Hata, M., Taniguchi, J., Tsuruoka, S., Moriwaki, K., Saitou, M., et al.
PMID:11912246. 2010. Claudin-2-deficient mice are defective in the leaky and cation-selective
Konrad, M., Schaller, A., Seelow, D., Pandey, A.V., Waldegger, S., Lesslauer, A., paracellular permeability properties of renal proximal tubules. Proc. Natl.
et al. 2006. Mutations in the tight-junction gene claudin 19 (CLDN19) are Acad. Sci. U.S.A. 107(17): 8011–8016. doi:10.1073/pnas.0912901107. PMID:
associated with renal magnesium wasting, renal failure, and severe ocular 20385797.
involvement. Am. J. Hum. Genet. 79(5): 949–957. doi:10.1086/508617. PMID: Nedergaard, S., Larsen, E.H., and Ussing, H.H. 1999. Sodium recirculation and
17033971. isotonic transport in toad small intestine. J. Membr. Biol. 168(3): 241–251.
Kraidith, K., Jantarajit, W., Teerapornpuntakit, J., Nakkrasae, L.I., Krishnamra, N., doi:10.1007/s002329900513. PMID:10191358.
and Charoenphandhu, N. 2009. Direct stimulation of the transcellular and Nellans, H.N. 1990. Intestinal calcium absorption. Interplay of paracellular and
paracellular calcium transport in the rat cecum by prolactin. Pflugers Arch. cellular pathways. Miner. Electrolyte Metab. 16(2–3): 101–108. PMID:2250615.
458(5): 993–1005. doi:10.1007/s00424-009-0679-6. PMID:19449156. Nellans, H.N., and Goldsmith, R.S. 1981. Transepithelial calcium transport by rat
Krawitt, E.L., and Schedl, H.P. 1968. In vivo calcium transport by rat small intes- cecum: high-efficiency absorptive site. Am. J. Physiol. 240(6): G424–G431.
tine. Am. J. Physiol. 214(2): 232–236. PMID:5635865. PMID:7246743.
Krug, S.M., Gunzel, D., Conrad, M.P., Rosenthal, R., Fromm, A., Amasheh, S., Nellans, H.N., and Kimberg, D.V. 1979. Anomalous calcium secretion in rat il-

Published by NRC Research Press


480 Biochem. Cell Biol. Vol. 92, 2014

eum: role of paracellular pathway. Am. J. Physiol. 236(4): E473–E481. PMID: 2011. Loss of claudin-15, but not claudin-2, causes Na+ deficiency and glucose
434203. malabsorption in mouse small intestine. Gastroenterology, 140(3): 913–923.
Ng, R.C., Rouse, D., and Suki, W.N. 1984. Calcium transport in the rabbit super- doi:10.1053/j.gastro.2010.08.006. PMID:20727355.
ficial proximal convoluted tubule. J. Clin. Invest. 74(3): 834–842. doi:10.1172/ Thorleifsson, G., Holm, H., Edvardsson, V., Walters, G.B., Styrkarsdottir, U.,
JCI111500. PMID:6236233. Gudbjartsson, D.F., et al. 2009. Sequence variants in the CLDN14 gene asso-
Nordin, B.E., Horsman, A., Marshall, D.H., Simpson, M., and Waterhouse, G.M. ciate with kidney stones and bone mineral density. Nat. Genet. 41(8): 926–
1979. Calcium requirement and calcium therapy. Clin. Orthop. Relat. Res. 930. doi:10.1038/ng.404. PMID:19561606.
(140): 216–239. doi:10.1097/00003086-197905000-00040. PMID:477077. Toka, H.R., Al-Romaih, K., Koshy, J.M., DiBartolo, S., III, Kos, C.H., Quinn, S.J.,
Pak, C.Y., Kaplan, R., Bone, H., Townsend, J., and Waters, O. 1975. A simple test et al. 2012. Deficiency of the calcium-sensing receptor in the kidney causes
for the diagnosis of absorptive, resorptive and renal hypercalciurias. N. Engl. parathyroid hormone-independent hypocalciuria. J. Am. Soc. Nephrol. 23(11):
J. Med. 292(10): 497–500. doi:10.1056/NEJM197503062921002. PMID:163960. 1879–1890. doi:10.1681/ASN.2012030323. PMID:22997254.
Pan, W., Borovac, J., Spicer, Z., Hoenderop, J.G., Bindels, R.J., Shull, G.E., et al. Tudpor, K., Teerapornpuntakit, J., Jantarajit, W., Krishnamra, N., and
2012. The Epithelial Sodium-Proton Exchanger, Nhe3, Is Necessary for Renal
Charoenphandhu, N. 2008. 1,25-dihydroxyvitamin D3 rapidly stimulates the
and Intestinal Calcium (Re)Absorption. Am. J. Physiol. Renal Physiol. 302(8):
solvent drag-induced paracellular calcium transport in the duodenum of
F943–F956. doi:10.1152/ajprenal.00504.2010. PMID:21937605.
female rats. J. Physiol. Sci. 58(5): 297–307. doi:10.2170/physiolsci.RP002308.
Biochem. Cell Biol. Downloaded from www.nrcresearchpress.com by San Francisco (UCSF) on 12/02/14

Pansu, D., Bellaton, C., Roche, C., and Bronner, F. 1983. Duodenal and ileal
PMID:18838052.
calcium absorption in the rat and effects of vitamin D. Am. J. Physiol. 244(6):
G695–G700. PMID:6602556. Van Itallie, C.M., Fanning, A.S., and Anderson, J.M. 2003. Reversal of charge
Pappenheimer, J.R., and Reiss, K.Z. 1987. Contribution of solvent drag through selectivity in cation or anion-selective epithelial lines by expression of differ-
intercellular junctions to absorption of nutrients by the small intestine ent claudins. Am. J. Physiol. Renal Physiol. 285(6): F1078–F1084. doi:10.1152/
of the rat. J. Membr. Biol. 100(2): 123–136. doi:10.1007/BF02209145. PMID: ajprenal.00116.2003. PMID:13129853.
3430569. Van Itallie, C.M., Rogan, S., Yu, A., Vidal, L.S., Holmes, J., and Anderson, J.M.
Praga, M., Vara, J., Gonzalez-Parra, E., Andres, A., Alamo, C., Araque, A., et al. 2006. Two splice variants of claudin-10 in the kidney create paracellular pores
1995. Familial hypomagnesemia with hypercalciuria and nephrocalcinosis. with different ion selectivities. Am. J. Physiol. Renal Physiol. 291(6): F1288–
Kidney Int. 47(5): 1419–1425. doi:10.1038/ki.1995.199. PMID:7637271. F1299. doi:10.1152/ajprenal.00138.2006. PMID:16804102.
Quamme, G.A. 1981. Effect of furosemide on calcium and magnesium transport Vargas-Poussou, R., Huang, C., Hulin, P., Houillier, P., Jeunemaitre, X.,
in the rat nephron. Am. J. Physiol. 241(4): F340–F347. PMID:7315959. Paillard, M., et al. 2002. Functional characterization of a calcium-sensing
Rievaj, J., Pan, W., Cordat, E., and Alexander, R.T. 2013. The Na+/H+ exchanger receptor mutation in severe autosomal dominant hypocalcemia with a
isoform 3 is required for active paracellular and transcellular Ca2+ transport Bartter-like syndrome. J. Am. Soc. Nephrol. 13(9): 2259–2266. doi:10.1097/01.
across murine cecum. Am. J. Physiol. Gastrointest. Liver Physiol. 305(4): ASN.0000025781.16723.68. PMID:12191970.
G303–G313. doi:10.1152/ajpgi.00490.2012. PMID:23764894. Wada, M., Tamura, A., Takahashi, N., and Tsukita, S. 2013. Loss of claudins 2 and
Rocha, A.S., and Kokko, J.P. 1973. Sodium chloride and water transport in the 15 from mice causes defects in paracellular Na+ flow and nutrient transport
medullary thick ascending limb of Henle. Evidence for active chloride trans- in gut and leads to death from malnutrition. Gastroenterology, 144(2): 369–
port. J. Clin. Invest. 52(3): 612–623. doi:10.1172/JCI107223. PMID:4685086. 380. doi:10.1053/j.gastro.2012.10.035. PMID:23089202.
Rosenthal, R., Milatz, S., Krug, S.M., Oelrich, B., Schulzke, J.D., Amasheh, S., et al. Walling, M.W., and Kimberg, D.V. 1973. Active secretion of calcium by adult rat
2010. Claudin-2, a component of the tight junction, forms a paracellular ileum and jejunum in vitro. Am. J. Physiol. 225(2): 415–422. PMID:4722404.
For personal use only.

water channel. J. Cell Sci. 123(Pt. 11): 1913–1921. doi:10.1242/jcs.060665. PMID: Walling, M.W., and Kimberg, D.V. 1975. Active secretion of calcium, sodium and
20460438. chloride by adult rat duodenum in vitro. Biochim. Biophys. Acta, 382(2):
Sas, D., Hu, M., Moe, O.W., and Baum, M. 2008. Effect of claudins 6 and 9 on 213–217. doi:10.1016/0005-2736(75)90179-0. PMID:1120156.
paracellular permeability in MDCK II cells. Am. J. Physiol. Regul. Integr. Wang, K.S., Ma, T., Filiz, F., Verkman, A.S., and Bastidas, J.A. 2000. Colon water
Comp. Physiol. 295(5): R1713–R1719. doi:10.1152/ajpregu.90596.2008. PMID: transport in transgenic mice lacking aquaporin-4 water channels. Am. J.
18784328. Physiol. Gastrointest. Liver Physiol. 279(2): G463–G470. PMID:10915657.
Schultheis, P.J., Clarke, L.L., Meneton, P., Miller, M.L., Soleimani, M.,
Wasserman, R.H. 2004. Vitamin D and the dual processes of intestinal calcium
Gawenis, L.R., et al. 1998. Renal and intestinal absorptive defects in mice
absorption. J. Nutr. 134(11): 3137–3139. PMID:15514288.
lacking the NHE3 Na+/H+ exchanger. Nat. Genet. 19(3): 282–285. doi:10.1038/
969. PMID:9662405. Wilcox, E.R., Burton, Q.L., Naz, S., Riazuddin, S., Smith, T.N., Ploplis, B., et al.
Seldin, D.W. 1999. Renal handling of calcium. Nephron, 81(Suppl. 1): 2–7. doi:10. 2001. Mutations in the gene encoding tight junction claudin-14 cause auto-
1159/000046292. PMID:9873208. somal recessive deafness DFNB29. Cell, 104(1): 165–172. doi:10.1016/S0092-
Sernka, T.J., and Borle, A.B. 1969. Calcium in the intestinal contents of rats on 8674(01)00200-8. PMID:11163249.
different calcium diets. Proc. Soc. Exp. Biol. Med. 131(4): 1420–1423. doi:10. Will, C., Breiderhoff, T., Thumfart, J., Stuiver, M., Kopplin, K., Sommer, K., et al.
3181/00379727-131-34121. PMID:5812009. 2010. Targeted deletion of murine Cldn16 identifies extra- and intrarenal
Simon, D.B., Lu, Y., Choate, K.A., Velazquez, H., Al-Sabban, E., Praga, M., et al. compensatory mechanisms of Ca2+ and Mg2+ wasting. Am. J. Physiol. Renal
1999. Paracellin-1, a renal tight junction protein required for paracellular Physiol. 298(5): F1152–F1161. doi:10.1152/ajprenal.00499.2009. PMID:20147368.
Mg2+ resorption. Science, 285(5424): 103–106. doi:10.1126/science.285.5424. Wittner, M., Mandon, B., Roinel, N., de Rouffignac, C., and Di Stefano, A. 1993.
103. PMID:10390358. Hormonal stimulation of Ca2+ and Mg2+ transport in the cortical thick as-
Stamatelou, K.K., Francis, M.E., Jones, C.A., Nyberg, L.M., and Curhan, G.C. 2003. cending limb of Henle's loop of the mouse: evidence for a change in the
Time trends in reported prevalence of kidney stones in the United States: paracellular pathway permeability. Pflugers Arch. 423(5–6): 387–396. doi:10.
1976–1994. Kidney Int. 63(5): 1817–1823. doi:10.1046/j.1523-1755.2003.00917.x. 1007/BF00374932. PMID:8351195.
PMID:12675858. Yu, A.S., Cheng, M.H., Angelow, S., Gunzel, D., Kanzawa, S.A., Schneeberger, E.E.,
Suki, W.N. 1979. Calcium transport in the nephron. Am. J. Physiol. 237(1): F1–F6. et al. 2009. Molecular basis for cation selectivity in claudin-2-based paracel-
PMID:380361. lular pores: identification of an electrostatic interaction site. J. Gen. Physiol.
Suki, W.N., Rouse, D., Ng, R.C., and Kokko, J.P. 1980. Calcium transport in the 133(1): 111–127. doi:10.1085/jgp.200810154. PMID:19114638.
thick ascending limb of Henle. Heterogeneity of function in the medullary Yu, A.S., Cheng, M.H., and Coalson, R.D. 2010. Calcium inhibits paracellular
and cortical segments. J. Clin. Invest. 66(5): 1004–1009. doi:10.1172/JCI109928. sodium conductance through claudin-2 by competitive binding. J. Biol.
PMID:7430341. Chem. 285(47): 37060–37069. doi:10.1074/jbc.M110.146621. PMID:20807759.
Sutton, R.A., and Dirks, J.H. 1975. The renal excretion of calcium: a review of Zhang, W., Na, T., Wu, G., Jing, H., and Peng, J.B. 2010. Down-regulation of
micropuncture data. Can. J. Physiol. Pharmacol. 53(6): 979–988. doi:10.1139/ intestinal apical calcium entry channel TRPV6 by ubiquitin E3 ligase
y75-136. PMID:769925. Nedd4-2. J. Biol. Chem. 285(47): 36586–36596. doi:10.1074/jbc.M110.175968.
Tamura, A., Hayashi, H., Imasato, M., Yamazaki, Y., Hagiwara, A., Wada, M., et al. PMID:20843805.

Published by NRC Research Press

You might also like