Hostname: page-component-76fb5796d-vvkck Total loading time: 0 Render date: 2024-04-26T04:06:05.957Z Has data issue: false hasContentIssue false

Review: Regulation of gastrointestinal and renal transport of calcium and phosphorus in ruminants

Published online by Cambridge University Press:  06 February 2020

M. R. Wilkens*
Affiliation:
Institute of Physiology and Cell Biology, University of Veterinary Medicine Hannover, Bischofsholer Damm 15/102, 30172 Hanover, Germany
A. S. Muscher-Banse
Affiliation:
Institute of Physiology and Cell Biology, University of Veterinary Medicine Hannover, Bischofsholer Damm 15/102, 30172 Hanover, Germany

Abstract

In comparison to monogastric animals, ruminants show some peculiarities in respect to the regulation of mineral homeostasis, which can be regarded as a concerted interplay between gastrointestinal absorption, renal excretion and bone mobilisation to maintain physiological Ca and phosphate (Pi) concentrations in serum. Intestinal absorption of Ca or Pi is mediated by two general mechanisms: paracellular, passive transport dominates when luminal Ca or Pi concentrations are high and transcellular. The contribution of active transport becomes more important when dietary Ca or Pi supply is restricted or the demand increased. Both pathways are modulated directly by dietary interventions, influenced by age and regulated by endocrine factors such as 1,25-dihydroxyvitamin D3. Similar transport processes are observed in the kidney. After filtration, Ca and Pi are resorbed along the nephron. However, as urinary Ca and Pi excretion is very low in ruminants, the regulation of these renal pathways differs from that described for monogastric species, too. Furthermore, salivary secretion, as part of endogenous Pi recycling, and bone mobilisation participate in the maintenance of Ca and Pi homeostasis in ruminants. Saliva contains large amounts of Pi for buffering rumen pH and to ensure optimal conditions for the rumen microbiome. The skeleton is a major reservoir of Ca and Pi to compensate for discrepancies between demand and uptake. But alterations of the regulation of mineral homeostasis induced by other dietary factors such as a low protein diet were observed in growing ruminants. In addition, metabolic changes, for example, at the onset of lactation have pronounced effects on gastrointestinal mineral transport processes in some ruminant species. As disturbances of mineral homeostasis do not only increase the risk of the animals to develop other diseases, but are also associated with protein and energy metabolism, further research is needed to improve our knowledge of its complex regulation.

Type
Review Article
Copyright
© The Animal Consortium 2020

Implications

Disturbances of mineral homeostasis are of significant relevance not only in dairy cows but also in beef cattle and small ruminants. In addition, the contribution of excreted phosphorus to the pollution of surface waters necessitates a revision of our livestock feeding regimes. The present review gives an overview on our current knowledge of the regulation of mineral transport across gastrointestinal and renal epithelia derived from functional and structural studies in different ruminant species as affected by age, lactation, feeding regime, etc. It highlights the physiological differences between monogastric animals and ruminants as well as the importance of combining different scientific approaches to improve our understanding of the complex mechanisms crucial for the maintenance of mineral homeostasis.

Introduction

Depending on management strategies, milk fever occurs in dairy cows with an incidence of 0% to 1%, 1.4% to 4%, and 5.7% to 6% in the first, the second and the third lactation, while the prevalence of subclinical hypocalcaemia defined as serum Ca concentration <2 mM amounts to 5.7% to 25%, 29.0% to 41%, and 49% (Reinhardt et al., Reference Reinhardt, Lippolis, McCluskey, Goff and Horst2011; Venjakob et al., Reference Venjakob, Borchardt and Heuwieser2017). The physiological response to transient hypocalcaemia is an increase in bone mobilisation followed by enhanced gastrointestinal absorption (van´t Klooster, Reference van´t Klooster1976). If these mechanisms are compromised, either the extent or the duration of hypocalcaemia is exacerbated resulting in increased risks of developing different diseases in early lactation depending on the duration of hypocalcaemia (Neves et al., Reference Neves, Leno, Bach and McArt2018). Reliable data on the prevalence of peripartum hypocalcaemia in small ruminants are scarce. Like cows, goats develop hypocalcaemia usually at the onset of lactation, while Ca homeostasis of sheep is generally more severely challenged during late gestation (Oetzel, Reference Oetzel1988; Brozos et al., Reference Brozos, Mavrogianni and Fthenakis2011).

Homeostatic control of phosphate (Pi) is also challenged at the onset of lactation. Subclinical hypophosphatemia around parturition is observed in >50% of dairy cows (Macrae et al., Reference Macrae, Whitaker, Burrough, Dowell and Kelly2006), and low serum Pi concentrations in cows suffering from milk fever are associated with an increased risk of developing downer cow syndrome (Menard and Thompson, Reference Menard and Thompson2007). However, high P percentage of pre-calving diets was identified as a risk factor for hypocalcaemia in a meta-analysis (Lean et al., Reference Lean, DeGaris, McNeil and Block2006), and a prepartum ration low in P seems to have beneficial effects on Ca homeostasis, probably because of an impact on bone mobilisation and vitamin D metabolism (Cohrs et al., Reference Cohrs, Wilkens and Grunberg2018).

Restriction of P and CP intake may occur for economic reasons or because animals are kept on deficient pastures (McGrath et al., Reference McGrath, Savage, Nolan and Elliott2012; Elfers et al., Reference Elfers, Wilkens, Breves and Muscher-Banse2015). This might be especially relevant in growing or fattening animals. On the other hand, environmental pollution with Pi and N of animal origin is leading to legal incentives to reduce the P and CP content of ruminant rations to the lowest possible level that does not negatively affect health and productivity.

In contrast to monogastric species, including rats and horses, no significant changes in renal Ca and Pi excretion are observed in bovines kept on restricted alimentary Ca supply, and the adaptation of gastrointestinal absorption seems to be less pronounced (Martz et al., Reference Martz, Belo, Weiss and Belyea1999; van Doorn et al., Reference van Doorn, van der Spek, Everts, Wouterse and Beynen2004; Zhang et al., Reference Zhang, Lai, Leung, Che and Wong2008; Taylor et al., Reference Taylor, Knowlton, McGilliard, Swecker, Ferguson, Wu and Hanigan2009). In Table 1 we present data from balance studies carried out in different ruminant species to illustrate that Ca and P absorption and secretion out of and into different gastrointestinal segments as well as urinary excretion are influenced by age, lactation and type of diet. There is inconsistency in the contribution of the forestomach of ruminants to overall Ca absorption (Table 1), which can partly be explained with differences in the composition of rations as mineral homeostasis interferes with other dietary factors such as dietary cation–anion difference, Mg and CP supply (Goff, Reference Goff2008; Muscher and Huber, Reference Muscher and Huber2010; Elfers et al., Reference Elfers, Liesegang, Wilkens, Breves and Muscher-Banse2016a; Wilkens et al., Reference Wilkens, Elfers, Schmicke, Breves and Muscher-Banse2018). Therefore, studies using a more mechanistic approach are an important tool to enhance our knowledge.

Table 1 Results from balance studies done with different ruminant species: intake, urinary excretion (UEX), pre-intestinal (PRE) and intestinal (INT) net absorption (ABS), faecal excretion (FEX) in grams per day, apparent digestibility (AD) in percentage

GS = ryegrass silage; F = formic acid; FF = formic acid and formaldehyde; SM = soybean meal; GCS = ryegrass and clover silage; DCAD = dietary cation–anion difference; 25-OHD = 25-hydroxyvitamin; n.d. = not determined.

Unfortunately, most research on the physiological mechanisms to maintain mineral homeostasis has used rodents as models for mammals in general. It is, therefore, the aim of this review to summarise the most important peculiarities of Ca and Pi transport across gastrointestinal and renal epithelia found in ruminants and highlight differences in comparison to monogastric animals. Throughout the following text, specific results obtained in ruminants will be indicated, while more general aspects often refer to studies done in rats and mice.

Methods

To evaluate renal and gastrointestinal Ca or Pi absorption and secretion in vivo, several different quantitative methods applied: balance studies using intact or cannulated animals, radioisotope tracer techniques, and the administration of stable strontium that can be used for this purpose, as its absorption shows a close correlation with that of Ca. As these experiments do not give explanations for sometimes inconsistent results, ex vivo methods are necessary to reveal the underlying mechanisms more precisely: isolated perfused organs, micropuncture experiments on renal transport, the everted sac technique that allows to control the composition of the luminal and serosal buffer solution and thus the chemical gradient, and the Ussing chamber is used to investigate transport mechanism by altering both the chemical and electrical gradients across the epithelium to differentiate between passive and active, paracellular and transcellular mechanisms. These functional studies are completed by in vitro experiments – for example, the quantification of RNA and protein expression of transporters and the functional characterisation applying electrophysiological techniques on cloned transporters. Although all these methods can greatly improve our understanding of physiological processes, the artificial conditions used or the fact that transporter abundance does not always represent in vivo activity may also provide challenges in interpretation. Taken together, our knowledge will probably increase if we combine the information derived from all these different approaches.

Endocrine control of calcium and phosphate transport

The concentrations of ionised Ca (Ca2+) and Pi in blood are regulated in a narrow range by 1,25-dihydroxyvitamin D3 (1,25-(OH)2D3), parathyroid hormone (PTH), calcitonin and fibroblast growth factor 23 (FGF23). Homeostasis is maintained by the interplay of gastrointestinal absorption, renal resorption and mobilisation of these inorganic ions from bone. Within minutes, a drop in blood Ca2+ induces the release of PTH from the parathyroid gland (Kumar and Thompson, Reference Kumar and Thompson2011) that stimulates the mobilisation of Ca and Pi from the skeleton (Ben-awadh et al., Reference Ben-awadh, Delgado-Calle, Tu, Kuhlenschmidt, Allen, Plotkin and Bellido2014). In monogastric animals, PTH also increases renal Ca resorption and Pi excretion by direct, rapid mechanisms (Besarab and Swanson, Reference Besarab and Swanson1982). Furthermore, PTH enhances the expression and activity of 1α-hydroxylase (CYP27B1), an enzyme that converts 25-hydroxyvitamin D3 (25-OHD3) to the biologically most active vitamin D metabolite 1,25-(OH)2D3 (Fraser and Kodicek, Reference Fraser and Kodicek1973). Furthermore, direct effects of plasma Ca and Pi on 1,25-(OH)2D3 concentrations were shown in rats (Bushinsky et al., Reference Bushinsky, Riera, Favus and Coe1985; Bushinsky et al., Reference Bushinsky, Nalbantian-Brandt and Favus1989).

In lactating animals, PTH-related peptide (PTHrP) is secreted by the mammary gland into both milk and blood. Although it can bind the PTH receptor, it is probably not involved in vitamin D metabolism but likely to act on bone mobilisation (Hernández-Castellano et al., Reference Hernández-Castellano, Hernandez and Bruckmaier2019). In addition, it was suggested to exert effects on renal Ca handling such as prolactin (Herm et al., Reference Herm, Muscher-Banse, Breves, Schröder and Wilkens2015).

Depending on the concentrations of plasma Ca and calcitonin, 1,25-(OH)2D3 either increases or inhibits bone mobilisation (Kurbel et al., Reference Kurbel, Radic, Kotromanovic, Puseljic and Kratofil2003). Via its genomic effects on Ca transporter expression that become present after a certain time lag, 1,25-(OH)2D3 stimulates renal resorption and intestinal absorption of Ca (Dusso et al., Reference Dusso, Brown and Slatopolsky2005) and limits its own synthesis by inhibiting CYP27B1 and stimulating the expression of 24-hydroxylase, the enzyme that initiates the inactivation of both 25-OHD3 and 1,25-(OH)2D3 (Chen and DeLuca, Reference Chen and DeLuca1995; Beckman and DeLuca, Reference Beckman and DeLuca2002).

In addition, 1,25-(OH)2D3 induces the production of a bone-derived phosphatonin, FGF23 (Saji et al., Reference Saji, Shigematsu, Sakaguchi, Ohya, Orita, Maeda, Ooura, Mima and Negi2010), that interacts with PTH expression and vitamin D metabolism and thus decreases plasma concentrations of 1,25-(OH)2D3 (Schiavi and Kumar, Reference Schiavi and Kumar2004; Krajisnik et al., Reference Krajisnik, Björklund, Marsell, Ljunggren, Åkerström, Jonsson, Westin and Larsson2007). Low dietary P intake decreased plasma concentrations of FGF23 and concomitantly increased 1,25-OH2D3 while plasma PTH was low (Antoniucci et al., Reference Antoniucci, Yamashita and Portale2006).

Protein intake also interferes with vitamin D metabolism. Growth hormone acts mainly through insulin-like growth factor 1 (IGF1). Uncoupling of this somatotropic axis indicated by low IGF1 plasma concentrations was observed during dietary protein restriction in growing goats and in peripartum dairy cows (Muscher et al., Reference Muscher, Piechotta, Breves and Huber2011; Piechotta et al., Reference Piechotta, Holzhausen, Araujo, Heppelmann, Sipka, Pfarrer, Schuberth and Bollwein2014). Reduced IGF1 was associated with decreased expression of CYP27B1 and affected bone mobilisation and intestinal Ca absorption probably via diminished plasma concentrations of 1,25-(OH)2D3 (Wilkens et al., Reference Wilkens, Elfers, Schmicke, Breves and Muscher-Banse2018).

As all these aspects might interfere with the strategies applied to stabilise mineral homeostasis in dairy cows and beef cattle – for example, dietary interventions, vitamin D supplementation, oral and parental administration of Ca, low IGF1 during negative energy balance, etc. (Reist et al., Reference Reist, Erdin, von Euw, Tschuemperlin, Leuenberger, Delavaud, Chilliard, Hammon, Kuenzi and Blum2003; Wilkens et al., Reference Wilkens, Oberheide, Schröder, Azem, Steinberg and Breves2012a; Domino et al., Reference Domino, Korzec and McArt2017) – a better understanding of the exact mechanisms is urgently needed.

Sites and mechanisms of gastrointestinal calcium absorption

Paracellular calcium absorption

Gastrointestinal Ca absorption can occur via the transcellular as well as paracellular pathways (Hoenderop et al., Reference Hoenderop, Nilius and Bindels2005). Passive, paracellular absorption can take place when the chemical gradient is high enough (>6 mM on the luminal side) to overcome the electrical gradient and the barrier formed by tight junction proteins, both of which hinder the transport of cations (Bronner, Reference Bronner1987). As paracellular Ca transport is dependent on its luminal concentration, it is dominant when Ca intake is high (Bronner and Pansu, Reference Bronner and Pansu1999) – for example, when Ca is provided as a bolus or via drenching. In addition, paracellular absorption can be driven by the so-called solvent drag effect. When water is absorbed due to hydrostatic and osmotic pressure, mineral ions solubilised by water dipole–ion interactions can also pass through the paracellular pathway (Goff, Reference Goff2018). The osmotic pressure contributing, to a large extent, to the solvent drag effect depends mainly on the transepithelial Na gradient generated by Na+-K+-ATPase (Karbach, Reference Karbach1992).

Paracellular Ca transport in both directions can be found throughout the entire intestine, depending on the gradient. It is likely that the rumen multilayer epithelium is too dense to allow significant amounts of Ca to be absorbed via the interstitial fluid unless the luminal concentration of Ca is increased dramatically by additional supply. This hypothesis is also supported by the comparison of rumen Ca flux rates and mannitol flux rates that are used to estimate transepithelial movement of water (Figure 1) (Wilkens et al., Reference Wilkens, Mrochen, Breves and Schröder2011 and Reference Wilkens, Richter, Fraser, Liesegang, Breves and Schröder2012b).

Figure 1 Unidirectional flux rates (J) from mucosal to serosal (ms) and from serosal to mucosal (sm) of Ca as a function of those of mannitol (Man) in the rumen tissues of sheep (n = 20) and goats (n = 20) determined using the Ussing chamber in the absence of any electrochemical gradient. As mannitol is used as a marker for paracellular transport of water, the lack of any relationship between Ca J ms and Man J ms indicates transcellular Ca absorption. Modified from Wilkens et al. (Reference Wilkens, Mrochen, Breves and Schröder2011) and (Reference Wilkens, Richter, Fraser, Liesegang, Breves and Schröder2012b).

Although it has been demonstrated that 1,25-(OH)2D3 has an effect on the expression of several tight junction proteins (Chirayath et al., Reference Chirayath, Gajdzik, Hulla, Graf, Cross and Peterlik1998; Kutuzova and DeLuca, Reference Kutuzova and DeLuca2004), it is not clear to what extend it regulates paracellular Ca absorption. A stimulation of the expression of claudin-2 and claudin-12, tight junction proteins that increase the permeability for Ca, was found in response to long-term dietary Ca restriction in the small intestine of goats (Elfers et al., Reference Elfers, Marr, Wilkens, Breves, Langeheine, Brehm and Muscher-Banse2016b); and in CaCo-2 cells treated with 1,25-(OH)2D3, paracellular permeability was increased (Chirayath et al., Reference Chirayath, Gajdzik, Hulla, Graf, Cross and Peterlik1998).

Transcellular, pre-intestinal calcium absorption

In monogastric animals, transcellular Ca absorption mainly occurs in the duodenum and upper jejunum (Hoenderop et al., Reference Hoenderop, Nilius and Bindels2005). The cellular mechanism consists of at least three steps: Ca enters the cell via the transient receptor potential vanilloid channel type 6 (TRPV6), is bound to the cytosolic protein calbindin-D9K (CaBPD9K), translocated to the basolateral membrane and extruded mainly by the plasma membrane Ca2+-ATPase isoform 1b (PMCA1b) (Figure 2). A significant stimulation of expression by 1,25-(OH)2D3 has been shown in all three of the abovementioned structures (Hoenderop et al., Reference Hoenderop, Nilius and Bindels2005).

Figure 2 Ca transport mechanisms and transepithelial potential difference in the rumen, small intestine, the thick ascending limp of the loop of Henle (TAL), and the distal and connecting tubules (DT, CT) of the kidneys in ruminant species. PMCA, plasma membrane Ca2+-ATPase isoform 1b; TRPV6/5, transient receptor potential vanilloid channel type 6/5; CaBPD9K/CaBPD28K, calbindin-D9K/D9K; NKCC2, Na+-K+-Cl co-transporter type 2; ROMK1, renal outer medullary K+ channel type 1; NCX, Na+/Ca2+ exchanger type 1. Explanations of the mechanisms are given in the corresponding text.

Marked differences to monogastric species have been described for ruminants, concerning the localisation and vitamin D sensitivity of Ca absorption along the gastrointestinal axis, particularly with respect to the forestomach compartment. While active Ca transport (3.0 ± 1.9 nmol/cm2·h) determined in Ussing chambers in the omasum has only been investigated and demonstrated in sheep (Höller et al., Reference Höller, Breves and Dubberke1988b), greater Ca net flux rates in the rumen were reported for sheep, goats (Figure 3) and cattle (13.8 ± 1.8 nmol/cm2·h). But as TRPV6 and CaBPD9K are not expressed in ovine, caprine and bovine rumen epithelia, this pre-intestinal Ca absorption is probably not mediated by the classical mechanism described for the intestine of monogastric animals (Höller et al., Reference Höller, Breves, Kocabatmaz and Gerdes1988a; Schröder et al., Reference Schröder, Goebel, Huber and Breves2001, Reference Schröder, Wilkens, Ricken, Leonhard-Marek, Fraser and Breves2015; Wilkens et al., Reference Wilkens, Mrochen, Breves and Schröder2011, Reference Wilkens, Richter, Fraser, Liesegang, Breves and Schröder2012bReference Schröder, Wilkens, Ricken, Leonhard-Marek, Fraser and Breves). Neither long-term dietary Ca restriction of sheep and goats nor administration of supraphysiological amounts of 1,25-(OH)2D3 resulted in increased Ca net flux rates across rumen epithelia measured in Ussing chambers (Figure 3) (Wilkens et al., Reference Wilkens, Mrochen, Breves and Schröder2011 and Reference Wilkens, Richter, Fraser, Liesegang, Breves and Schröder2012b). Hyde and Fraser estimated Ca transport in vivo by an administration of stable strontium. In contrast to the abovementioned studies, they observed that rumen Ca transport doubled after treatment of sheep with 1α-OHD3 (Hyde and Fraser, Reference Hyde and Fraser2014). However, no satisfying explanation for this inconsistency was found. It might be speculated that alterations regarding passage rate and rumen motility as a response to the hypercalcaemic effect of the treatment contribute to overall Ca transport in vivo (Daniel, Reference Daniel1983). In vitro, rumen Ca net flux rates of sheep determined in Ussing chambers seem to depend on the presence of short-chain fatty acids (SCFA; 0, 40 and 100 mmol/l in the mucosal buffer: 2.41 ± 0.55, 9.59 ± 1.55 and 19.41 ± 3.37 nmol/cm2·h) and are increased by feeding 15 g of concentrate per kilogram body weight for 3 weeks in comparison to a ration consisting of hay only (5.63 ± 0.54, 17.43 ± 0.70 and 34.54 ± 2.67 nmol/cm2·h) (Uppal et al., Reference Uppal, Wolf and Martens2003a). Therefore, an apical transport mechanism based on a Ca2+/H+ exchange system was discussed (Lutz and Scharrer, Reference Lutz and Scharrer1991; Schröder et al., Reference Schröder, Wilkens, Ricken, Leonhard-Marek, Fraser and Breves2015). Whether a higher Ca intake in the concentrate fed sheep might have altered rumen Ca transport mechanisms directly cannot be clarified. On the one hand, low luminal Ca concentrations before sacrifice did not influence the flux rates in sheep and goats (Figure 3). On the other hand, a greater contribution of pre-intestinal Ca to overall absorption was reported in a meta-analysis (Schröder and Breves, Reference Schröder and Breves2006).

Figure 3 Rumen Ca net flux rates (Jnet) of female sheep and goats aged 6 to 7 months kept on adequate (con, 0.92% and 1.10%, n = 5) or restricted Ca supply (Ca–, 0.26% and 0.22%, n = 5) treated with a placebo or fed the same diets and treated with 1,25-dihydroxyvitamin D3 (1,25-(OH)2D3, n = 5) (0.5 µg/kg body weight) 12 h before sacrifice determined in Ussing chambers in the absence of any electrochemical gradient. Means ± SEM. Modified from Wilkens et al. (Reference Wilkens, Mrochen, Breves and Schröder2011) and (Reference Wilkens, Richter, Fraser, Liesegang, Breves and Schröder2012b).

As Na transport is also – although to a lesser extent – increased by higher luminal concentrations of SCFA (Uppal et al., Reference Uppal, Wolf, Khahra and Martens2003b), rumen Ca transport could be mediated by a more complex ion exchanging mechanism. Another candidate for the apical uptake of Ca could be transient receptor potential vanilloid channel type 3 (TRPV3). In patch clamp measurements, agonists of this channel were shown to stimulate currents mediated by Ca2, NH4 and Na into HEK-293 cells expressing bovine TRPV3 (Schrapers et al., Reference Schrapers, Sponder, Liebe, Liebe and Stumpff2018). An involvement of Na transport might also explain the finding that feeding a ration negative in dietary cation–anion difference (DCAD) to sheep affects the ratio of the electroneutral to the electrogenic component of rumen Ca transport from the mucosal to the serosal side (Figure 4) (Wilkens et al., Reference Wilkens, Praechter, Breves and Schröder2016). In vivo and in vitro studies have reported both a stimulating effect of essential oils, substances that are known to interfere with TRP channels, and also interactions between the absorption of Ca, NH4 and Na. A conductance for NH4 was blocked by divalent cations in bovine rumen epithelial cells. Addition of 10 µM menthol enhanced Ca net flux rates determined for ovine rumen epithelia in Ussing chambers from 8.60 ± 1.43 to 13.24 ± 0.91 nmol/cm2·h and Na-mediated short-circuit currents (Rosendahl et al., Reference Rosendahl, Braun, Schrapers, Martens and Stumpff2016). In dairy cows, oral administration of 1.2 g essential oils with menthol as the major compound increased plasma Ca from 2.46 to 2.53 mM and decreased plasma urea from 4.28 to 3.92 mM (Braun et al., Reference Braun, Schrapers, Mahlkow-Nerge, Stumpff and Rosendahl2019).

Figure 4 Correlation between electrical driving force and unidirectional Ca fluxes (J) from mucosal to serosal (ms) and from serosal to mucosal (sm) of castrated male sheep aged 8 months kept either on a ration positive in dietary cation–anion difference (DCAD) (control, n = 4) or negative in DCAD (low DCAD, n = 5). The electroneutral component of Jms represented by the intercept of the linear function revealed by regression analysis is greater (P < 0.01) in sheep kept on a diet low in DCAD (control: Jms = 7.76 (±1.23) + 7.53 (±0.77)·ξ –0.5; low DCAD: Jms = 13.32 (±4.42) + 9.95 (±2.77)·ξ –0.5). Means ± SEM. Modified from Wilkens et al. (Reference Wilkens, Praechter, Breves and Schröder2016).

As also shown in Table 1, these findings show that rumen Ca transport depends on luminal abundance of different factors and nutrients. Therefore, greater Ca flux rates determined for rumen tissue of lactating goats in comparison to dried-off animals (2.28 ± 0.35 v. 6.75 ± 1.16 nmol/cm2·h) could be either a direct effect of lactation or be caused by the different feeding regime and/or an enlargement of the luminal surface (Starke et al., Reference Starke, Reimers, Muscher-Banse, Schröder, Breves and Wilkens2016). In cows, rumen Ca transport estimated by the administration of stable strontium is stimulated by lactation and decreased when forestomach motility is reduced (Hyde et al., Reference Hyde, Wilkens and Fraser2019). As impaired motility was observed with decreased plasma Ca concentrations (Daniel, Reference Daniel1983), inefficient ruminal Ca absorption following a disturbance of Ca mobilisation from the skeleton might aggravate hypocalcaemia in peripartum cows.

Transcellular, intestinal calcium absorption

Studies provide conflicting results on the intestinal absorption of Ca. 1,25-(OH)2D3-regulated proteins, essential for transcellular Ca absorption, have been identified in the small intestine of cattle (Yamagishi et al., Reference Yamagishi, Miyazaki and Naito2006; Schröder et al., Reference Schröder, Wilkens, Ricken, Leonhard-Marek, Fraser and Breves2015), sheep (Schröder et al., Reference Schröder, Goebel, Huber and Breves2001; Wilkens et al., Reference Wilkens, Kunert-Keil, Brinkmeier and Schröder2009, Reference Wilkens, Mrochen, Breves and Schröder2011) and goats (Wilkens et al., Reference Wilkens, Richter, Fraser, Liesegang, Breves and Schröder2012b; Elfers et al., Reference Elfers, Wilkens, Breves and Muscher-Banse2015). However, Ca transport across ovine and caprine epithelia when determined in vitro in the absence of an electrochemical gradient appears to be very low compared to monogastric animals such as horses using the same methods (Figure 5) (Wilkens et al., Reference Wilkens, Marholt, Eigendorf, Muscher-Banse, Feige, Schröder, Breves and Cehak2017). In the colon, Ca net flux rates are also very low. As in the rumen, significant active Ca transport (6.55 ± 2.01 nmol/cm2·h) across the colon of sheep is only detectable in the presence of SCFA. Unfortunately, no published data are available on intestinal Ca transport determined for bovine epithelia.

Figure 5 Intestinal Ca net flux rates (Jnet) of horses of both sexes, aged 3 to 22 years (n = 10), female sheep (n = 5) and female goats (n = 5) aged 6 to 7 months kept on adequate Ca supply determined in Ussing chambers in the absence of any electrochemical gradient. Means ± SEM. Modified from Wilkens et al. (Reference Wilkens, Mrochen, Breves and Schröder2011), (Reference Wilkens, Richter, Fraser, Liesegang, Breves and Schröder2012b) and (Reference Wilkens, Marholt, Eigendorf, Muscher-Banse, Feige, Schröder, Breves and Cehak2017).

In goats kept on a low Ca diet or treated with vitamin D, duodenal Ca flux rates measured in Ussing chambers were significantly increased in some (Wilkens et al., Reference Wilkens, Richter, Fraser, Liesegang, Breves and Schröder2012b) but not in all studies (Schröder et al., Reference Schröder, Rittmann, Pfeffer and Breves1997; Sidler-Lauff et al., Reference Sidler-Lauff, Boos, Kraenzlin and Liesegang2010). Higher flux rates and a more pronounced stimulation of transcellular Ca transport by dietary Ca restriction was accompanied by an increase in RNA, and protein expression of TRPV6 could be shown for the jejunum of goats indicating that this segment is more active for overall Ca absorption (Figure 6a) (Wilkens et al., Reference Wilkens, Richter, Fraser, Liesegang, Breves and Schröder2012b; Elfers et al., Reference Elfers, Wilkens, Breves and Muscher-Banse2015). Although the efficiency of net Ca absorption from the jejunum, measured by applying the Thiry-Vella loop technique, was increased in sheep with dietary Ca restriction (Abdel-Hafeez et al., Reference Abdel-Hafeez, Manas-Almendros, Ross, Care and Marshall1982), this could not be demonstrated in protein expression studies and Ussing chamber experiments (Figure 6b) (Wilkens et al., Reference Wilkens, Mrochen, Breves and Schröder2011). In goats kept on a reduced protein diet, the intestinal absorption of Ca was diminished with a concomitant reduction of CaBPD9K and PMCA1b, probably caused by decreased 1,25-(OH)2D3 concentrations (Figure 6c) (Elfers et al., Reference Elfers, Wilkens, Breves and Muscher-Banse2015).

Figure 6 Intestinal Ca net flux rates (Jnet) of female sheep and goats aged 6 to 7 months kept on adequate (control, 0.92% and 1.10%, n = 5) or restricted Ca supply (Ca restriction, 0.26% and 0.22%, n = 5) and male goats aged 3 to 4 months kept on adequate (control, 22% CP, n = 7) or restricted n supply (n restriction, 8% CP, n = 6) determined in Ussing chambers in the absence of any electrochemical gradient. Significant differences revealed by the Student’s t test are marked with asterisks. Means ± SEM; *, P < 0.05. Modified from Elfers et al. (Reference Elfers, Wilkens, Breves and Muscher-Banse2015), Wilkens et al. (Reference Wilkens, Mrochen, Breves and Schröder2011) and (Reference Wilkens, Richter, Fraser, Liesegang, Breves and Schröder2012b).

Taken together with results from lactating and dried-off sheep and goats, it might be concluded that the responsiveness of intestinal Ca absorption to enhanced demand or restricted supply varies between different species and ages (Wilkens et al., Reference Wilkens, Breves and Schröder2014; Klinger et al., Reference Klinger, Schröder, Gemmer, Reimers, Breves, Herrmann and Wilkens2016; Starke et al., Reference Starke, Reimers, Muscher-Banse, Schröder, Breves and Wilkens2016). In lactating and non-lactating cows, balance studies demonstrated that Ca digestibility is not increased with dietary Ca restriction, although lactation itself seems to enhance gastrointestinal absorption (Table 1). However, a full adaptation to increased Ca demand during lactation seems to take at least 2 days (van´t Klooster, Reference van´t Klooster1976).

Salivary secretion of phosphorus

As rumen Pi concentrations play a pivotal role for rumen buffering, fermentation and microbial protein synthesis, large amounts of Pi are secreted with saliva and resorbed in the lower digestive tract. Rumen Pi concentration thus depends on dietary P intake and the rate of salivary Pi secretion (Breves and Schröder, Reference Breves and Schröder1991). The role of salivary Pi is also reflected by the observation that salivary Pi concentrations and expression of NaPi IIb (SLC34A2) and Pi transporter PiT1 (SLC20A1), both Na-dependent Pi transporters, increase with age, that is, with the development of the gastrointestinal tract (Huber et al., Reference Huber, Roesler, Muscher, Hansen, Widiyono, Pfeffer and Breves2003). Interestingly, significant differences in both rumen Pi concentrations (see below) and salivary Pi were found when adult sheep (11.3 ± 1.2 mM) and goats (23.1 ± 3.2 mM) were kept on the same ration, indicating species differences in respect to salivary Pi secretion (Wilkens et al., Reference Wilkens, Breves and Schröder2014).

Data on the regulation of salivary Pi secretion are inconsistent. Furthermore, data on potential molecular regulatory mechanisms of Pi transport in salivary glands are lacking. Intravenous loading with Pi resulted in an increase in Pi secretion via the parotid gland of sheep and cows, indicating plasma Pi concentration is the most important factor (Scott and Beastall, Reference Scott and Beastall1978; Riad et al., Reference Riad, Lefaivre and Barlet1987). In goats and sheep, the administration of PTH induced an increase in saliva Pi concentration in some studies (Wright et al., Reference Wright, Blair-West, Nelson and Tregear1982; Isac et al., Reference Isac, Valverde, Lupiani, Martinez de Victoria and Manas1989), while others found a decreasing effect (Mañas-Almendros et al., Reference Mañas-Almendros, Ross and Care1982). An injection of exogenous 1,25-(OH)2D3 reduced salivary Pi concentrations in sheep and cows (Mañas-Almendros et al., Reference Mañas-Almendros, Ross and Care1982; Riad et al., Reference Riad, Lefaivre and Barlet1987). A possible explanation for these contradictions could be the alteration of saliva flow rate, which is difficult to be addressed in the experimental design (Isac et al., Reference Isac, Valverde, Lupiani, Martinez de Victoria and Manas1989). The salivary flow rate is mainly regulated by the physical nature of the diet fed (Wilson and Tribe, Reference Wilson and Tribe1963). Pelleted diets induced lower daily saliva flow rates than chopped or long hay based on less chewing (Duric et al., Reference Duric, Zhao, Orskov and Chen1994).

However, rumen Pi concentrations were significantly increased in sheep (24.2 ± 1.0 v. 28.0 ± 0.9 mM) and goats (42.1 ± 2.9 v. 50.0 ± 3.8 mM) kept on a Ca-restricted ration for several weeks that led to an increase in an endogenous production of 1,25-(OH)2D3, even though plasma (1.97 ± 0.13 v. 1.87 ± 0.21 mM) and salivary concentrations of Pi (37.3 ± 3.4 v. 37.0 ± 1.2 mM) were not affected by this feeding regime in goats (Wilkens et al., Reference Wilkens, Richter, Fraser, Liesegang, Breves and Schröder2012b, Reference Wilkens, Breves and Schröder2014).

Sites and mechanisms of gastrointestinal phosphorus absorption

The absorption of Pi takes place along the whole of the intestinal tract. In principle, Pi absorption can be divided into a passive paracellular process and a saturable, active transcellular process. In vivo studies with the temporarily isolated reticulo-rumen from sheep demonstrated a positive linear relationship between rumen Pi concentrations and net Pi disappearance, indicating passive paracellular absorption of Pi. No indications of active Pi transport or saturation phenomena could be determined (Breves et al., Reference Breves, Höller, Packheiser, Gabel and Martens1988; Beardsworth et al., Reference Beardsworth, Beardsworth and Care1989). In vitro studies with rumen epithelium confirmed that no Pi net flux was found under short-circuit conditions, that is, in the absence of any electrochemical gradient, in Ussing chamber experiments (Breves et al., Reference Breves, Höller, Packheiser, Gabel and Martens1988). A passive process of Pi absorption also occurs in the omasal epithelium of sheep (Höller et al., Reference Höller, Breves and Dubberke1988b). However, balance studies clearly indicate that there is no net absorption from but a substantial secretion of Pi into the forestomach in vivo (Table 1).

In ruminants as in monogastric species, the small intestine is the major site for Pi absorption (Pfeffer et al., Reference Pfeffer, Thompson and Armstrong1970). Dietary P concentration and 1,25-(OH)2D3 are the main regulators of intestinal Pi transport in monogastric species. Paracellular Pi transport across the intercellular spaces of the small intestines has been postulated. However, no potential candidate genes which might mediate such mechanisms have been identified.

An H+-dependent Pi co-transport into duodenal brush border membrane vesicles (BBMV) from sheep and cattle was demonstrated, and this was stimulated by low dietary P (Shirazi-Beechey et al., Reference Shirazi-Beechey, Kemp, Dyer and Beechey1989, Reference Shirazi-Beechey, Beechey, Penny, Vayro, Buchan and Scott1991). In the jejunum of sheep, the saturation of Pi absorption was demonstrated with the use of a Thiry-Vella loop when the infused solution was as high as 15 mM of Pi (Care et al., Reference Care, Barlet and Abdel-Hafeez1980). Interestingly, studies on jejunal unidirectional Pi flux rates in Ussing chambers using intestinal tissue from sheep and goats demonstrated a substantial part of active Pi transport which was inhibited by arsenate by a reduction of luminal Na+ concentrations and by serosal addition of ouabain. This Na+-dependent Pi co-transport could be stimulated by dietary P depletion, while changes in vitamin D metabolism were not involved (Schröder et al., Reference Schröder, Kappner, Failing, Pfeffer and Breves1995). This provides evidence for an active Pi transport mechanism like that in non-ruminant species, with the highest absorption rates being found in the ileum of young goats (about 3 to 4 months) and adult sheep (Schröder et al., Reference Schröder, Kappner, Failing, Pfeffer and Breves1995; Elfers et al., Reference Elfers, Wilkens, Breves and Muscher-Banse2015). To confirm that active Pi transport is Na+-dependent, Pi uptake studies into isolated BBMV from goat jejunum were performed under different conditions of extravesicular Na+ and H+ concentrations (Schröder and Breves, Reference Schröder and Breves1996). The results are similar to data from monogastric species and showed that a major proportion of jejunal Pi uptake is Na+-dependent, and can be stimulated by H+, in contrast to duodenal Pi transport which is H+-dependent and Na-sensitive.

After the molecular identification of an intestinal Na+-dependent Pi transporter in mice (NaPi IIb) (Hilfiker et al., Reference Hilfiker, Hattenhauer, Traebert, Forster, Murer and Biber1998), it could be shown that caprine NaPi IIb expression corresponded to murine NaPi IIb (Huber et al., Reference Huber, Walter, Schröder, Biber, Murer and Breves2000). Both NaPi IIb mRNA and protein were absent in the duodenum of goats, while NaPi IIb was strongly expressed in the jejunum (Huber et al., Reference Huber, Walter, Schröder and Breves2002). With jejunal BBMV, it could be shown that a high linear correlation exists between transport capacity for Pi and NaPi IIb protein expression, indicating that the majority of Na-dependent Pi transport was mediated by NaPi IIb. Furthermore, the existence of an additional electrogenic Na-dependent Pi transporter, called PiT1, was shown in the small intestine of goats (Figure 7) (Elfers et al., Reference Elfers, Wilkens, Breves and Muscher-Banse2015). PiT1 belongs to the Pi transporter family that uses either Na or H+ gradients to transport Pi (Saier, Reference Saier2000). The mechanism for extrusion of Pi is still under investigation. In Holstein cows, the highest NaPi IIb RNA expression was found in the distal jejunum and ileum, while the expression in the upper intestinal segments was nearly absent (Foote et al., Reference Foote, Lambert, Brady and Muir2011).

Figure 7 Pi transport mechanisms and transepithelial potential difference in the small intestine and proximal tubule of the kidneys in ruminant species. Apical entry occurs through Na-dependent Pi transporter family (NaPi) subtypes IIa and IIc or IIb and Na-dependent phosphate transporter 1 (PiT1). Basolateral extrusion mechanism of Pi is currently unknown. Further explanations of the mechanisms are given in the corresponding text.

To characterise Pi transport in the duodenum in more detail, transepithelial Pi flux rates have been performed in Ussing chamber experiments, in the presence or absence of mucosal Na at different pH levels. From these studies, it could be concluded that at least two different Pi transport mechanisms exist in the goat intestinal tract which are regionally separate: in the duodenum, Pi uptake is mainly mediated by an H+-dependent Na-sensitive mechanism, while in the jejunum, Pi uptake is mainly Na-dependent and H+-sensitive (Schröder et al., Reference Schröder, Kappner, Failing, Pfeffer and Breves1995; Huber et al., Reference Huber, Walter, Schröder and Breves2002).

In lactating goats, Pi flux rates from the mucosal to the serosal side of the epithelium, determined in the duodenum and jejunum, were significantly smaller in comparison to dried-off animals, resulting in a reduced net absorption (Figure 8). This was accompanied by a downregulation of jejunal NaPi IIb, both on RNA and protein levels, probably as a consequence of either higher P intake or enhanced mobilisation of Pi from the skeleton (Starke et al., Reference Starke, Reimers, Muscher-Banse, Schröder, Breves and Wilkens2016). In line with other studies, NaPi IIb was not detectable in the duodenum (Huber et al., Reference Huber, Walter, Schröder and Breves2002). An ontogenetic study with goats found that Pi-binding properties changed during the development of the gastrointestinal tract of growing animals, indicating that alterations of NaPi IIb and/or PiT1 must be taking place (Huber et al., Reference Huber, Roesler, Muscher, Hansen, Widiyono, Pfeffer and Breves2003).

Figure 8 Unidirectional, duodenal (Duo) and jejunal (Jeju) Pi flux rates (J) from serosal to mucosal (sm) and from mucosal to serosal (ms) of dried-off (n = 6) and lactating goats (n = 6) determined in Ussing chambers in the absence of any electrochemical gradient. Significant differences revealed by the Student’s t test are marked with asterisks. Means ± SEM; *, P < 0.05. Modified from Starke et al. (Reference Starke, Reimers, Muscher-Banse, Schröder, Breves and Wilkens2016).

Interestingly, dietary P depletion modifies the intestinal absorption of Pi in young goats (4 to 5 months) but without the involvement of vitamin D metabolism (Schröder et al., Reference Schröder, Kappner, Failing, Pfeffer and Breves1995). Therefore, an unknown Pi-sensing mechanism is hypothesised in the small intestine of ruminant species. Even when 1,25-(OH)2D3 concentrations were altered during dietary protein reduction, modulation of the expression of NaPi IIb and PiT1 in the small intestine was not found (Elfers et al., Reference Elfers, Wilkens, Breves and Muscher-Banse2015). These results contrast with data from monogastric species where a 1,25-(OH)2D3-dependent regulation of NaPi IIb was found (Murer et al., Reference Murer, Forster and Biber2004).

In young lambs (1 week old), the efficacy of Pi absorption from the colon was almost the same as in the upper and mid-jejunum, but the velocity of Pi absorption decreased during subsequent development (Scharrer, Reference Scharrer1985). In adult sheep fitted with re-entrant cannulae, the proximal colon was perfused with an electrolyte solution free of Pi, and net Pi secretion was determined. Net absorption of Pi from the colon was shown when Pi concentrations of the electrolyte solution were between 2.5 and 6.5 mM (Höller et al., Reference Höller, Figge, Richter and Breves1988c).

Renal handling of calcium

In the kidneys of rats and hamsters, 70% of filtered Ca is resorbed paracellulary in the proximal tubules, while up to 20% is resorbed in the thick ascending limb of the loop of Henle (TAL) (Lassiter et al., Reference Lassiter, Gottschalk and Mylle1963). In the proximal tubules, where an osmotic gradient is built up due to the resorption of Na, glucose and amino acids, paracellular Ca transport is driven mainly by the solvent drag effect (Friedman and Gesek, Reference Friedman and Gesek1995). In TAL, a lumen-positive transepithelial potential difference is generated by the electroneutral uptake of Na, K and Cl via the Na+-K+-2Cl-co-transporter (NKCC) followed by the basolateral extrusion of Cl and the apical secretion of K. Tight junctions in this segment contain claudin-16 that increases cation permeability, claudin-19 that blocks anion permeability, and claudin-14 that decreases cation permeability mediated by claudin-16 (Negri, Reference Negri2015). Interestingly, we observed a downregulation of claudin-19 with dietary Ca restriction in sheep and goats, which contrasts findings in rats (Frick et al., Reference Frick, Asplin, Favus, Culbertson, Krieger and Bushinsky2013), and an upregulation of claudin-16 during lactation in goats (unpublished results).

Active, transcellular, 1,25-(OH)2D3-regulated Ca transport is found in the distal and connecting tubules. For active resorption of Ca, a transport mechanism similar to that generally accepted for the small intestine (TRPV5, calbindin-D28K and basolateral extrusion by the Na+/Ca2+ exchanger NCX1) has been described (Figure 2) (Hoenderop et al., Reference Hoenderop, Dardenne, van Abel, van der Kemp, van Os, St. Arnaud and Bindels2002). In rodents fed a diet low in Ca, there was an increase in RNA expression of TRPV5 and CaBPD28K (Hoenderop et al., Reference Hoenderop, Dardenne, van Abel, van der Kemp, van Os, St. Arnaud and Bindels2002; Ko et al., Reference Ko, Lee, Vo, Jung, Choi, Cheung, Kim, Park, Oh and Jeung2009). Furthermore, it was demonstrated in mice that lactation stimulated renal RNA expression of TRPV5 and CaBPD28K (van Cromphaut et al., Reference van Cromphaut, Rummens, Stockmans, van Herck, Dijcks, Ederveen, Carmeliet, Bouillon and Carmeliet2003). For adult sheep and goats, we found that ruminant kidney does not respond to a challenge of Ca homeostasis by altered expression of structures mediating Ca resorption. With respect to CaBPD28K, we even observed a downregulation in dietary Ca-restricted or lactating goats, instead of the stimulation that has been reported for mice (Herm et al., Reference Herm, Muscher-Banse, Breves, Schröder and Wilkens2015). Interestingly, in lactating goats, urinary Ca excretion was not increased. We speculated that enhanced resorption in TAL mediated by prolactin and/or PTHrP might have compensated for the downregulation of TRPV5, CaBPD28K and NCX1 RNA expressions in the distal parts of nephron (Herm et al., Reference Herm, Muscher-Banse, Breves, Schröder and Wilkens2015). Our findings on the structural level regarding animals kept on a low Ca diet could be explained by characteristically low renal Ca excretion in adult ruminants that cannot be further diminished when Ca homeostasis is challenged. As in cattle and lactating cows (Table 1), fractional excretion of Ca was not reduced by dietary Ca restriction in small ruminants (sheep: 0.83 ± 0.22 v. 1.06 ± 0.24%, goats: 0.71 ± 0.13 v. 1.03 ± 0.21%).

However, in young goats (3 to 4 months) kept on a Ca-reduced diet, a stimulation of CaBPD28K and NCX1 RNA expression occurred based on elevated 1,25-(OH)2D3 levels. This is in line with data from balance studies conducted with goat kids (Table 1). A concomitant decrease in dietary Ca and protein in these young animals caused a decrease in 1,25-(OH)2D3 concentrations, resulting in a downregulation of TRPV5, CaBPD28K and NCX1 protein expressions (Firmenich et al., Reference Firmenich, Elfers, Wilkens, Breves and Muscher-Banse2018).

A way to increase renal Ca excretion in ruminants is to feed a ration negative in DCAD (Table 1). This feeding regime induces a compensated acidosis that results in increased tissue responsiveness to PTH (Goff et al., Reference Goff, Liesegang and Horst2014). In addition, DCAD treatment leads to significant changes in Ca balance before parturition. Several studies in dairy cows have demonstrated that urinary pH is decreased, while renal excretion of Ca is increased up to 10-fold (0.4 ± 0.2 v. 4.1 ± 0.9 g per day) (Grünberg et al., Reference Grünberg, Donkin and Constable2011; Wilkens et al., Reference Wilkens, Oberheide, Schröder, Azem, Steinberg and Breves2012a). As TRPV5 activity is pH-dependent, increased renal Ca excretion might be caused by a direct inhibitory effect of tubular acidosis on renal resorption of Ca as shown in the kidney of dogs and for rabbit TRPV5 (Sutton et al., Reference Sutton, Wong and Dirks1979; Yeh et al., Reference Yeh, Sun, Lee, Chen and Huang2003). In preliminary experiments conducted with sheep, we observed that the expression of TRPV5, CaBPD28K and NCX1 was not significantly altered under these conditions (unpublished results). Assuming that this occurs in cows kept on a low DCAD ration, too, this might indicate that renal resorption is immediately restored when the ration is changed postpartum. An adaptation on the functional level occurs faster than the stimulation of gene expression and could contribute to the beneficial effects of a low DCAD diet in peripartum cow.

Renal handling of phosphorus

During normophosphataemia, about 98% to 99% of filtered Pi is resorbed in the kidneys of ruminant species (Widiyono et al., Reference Widiyono, Huber, Failing and Breves1998). The mean plasma threshold for renal Pi excretion in goats lies around 4.3 mM (Widiyono et al., Reference Widiyono, Huber, Failing and Breves1998). Filtered Pi is reabsorbed mainly by the proximal tubule cells. The uptake of filtered Pi at the apical side is mediated by Na+-dependent Pi transporters: electrogenic NaPi IIa (SLC34A1) and electroneutral NaPi IIc (SLC34A3) in both ruminant and non-ruminant species (Figure 7) (Biber and Murer, Reference Biber and Murer1994; Shirazi-Beechey et al., Reference Shirazi-Beechey, Penny, Dyer, Wood, Tarpey, Scott and Buchan1996; Huber et al., Reference Huber, Roesler, Muscher, Hansen, Widiyono, Pfeffer and Breves2003; Starke et al., Reference Starke, Cox, Sudekum and Huber2013). Ovine and caprine amino acid sequence, kinetic and stoichiometric parameters of renal cortex Na+-dependent Pi transport are comparable to the type IIa Na+/Pi co-transport in monogastric species (Schröder et al., Reference Schröder, Walter, Breves and Huber2000). Basolateral Pi extrusion mechanism is still unknown.

Changes in dietary P intake and consequent changes in extracellular Pi and PTH are the main regulators of renal Pi transporters in monogastric species (Biber et al., Reference Biber, Murer and Forster1998). In mature goats and sheep on a P-reduced diet, no changes in renal transport capacities (Schröder et al., Reference Schröder, Walter, Breves and Huber2000) or on NaPi IIa expression (Huber et al., Reference Huber, Roesler, Holthausen, Pfeffer and Breves2007) were determined. In contrast, dietary P restriction altered urinary Pi excretion in goat kids (Table 1). In young goats (3 to 4 months) on a high P diet, there was a decrease in renal Pi reabsorption capacity and an internalisation of NaPi IIa occurred (Huber et al., Reference Huber, Roesler, Holthausen, Pfeffer and Breves2007; Muscher et al., Reference Muscher, Hattendorf, Pfeffer, Breves and Huber2008). Strong correlations between NaPi IIa mRNA and plasma Pi as well as plasma PTH concentrations indicated that elevated Pi and high PTH concentrations were able to modulate renal Pi excretion by reducing Pi reabsorption (Muscher et al., Reference Muscher, Hattendorf, Pfeffer, Breves and Huber2008). This phenomenon is different to that in monogastric animals where NaPi IIa expression was decreased only at the protein level (Murer et al., Reference Murer, Forster, Hernando, Lambert, Traebert and Biber1999).

Besides dietary P and PTH, a reduction in dietary protein also modulates mineral homeostasis in young goats (4 to 5 months) (Muscher et al., Reference Muscher, Piechotta, Breves and Huber2011). A significant increase in NaPi IIa expression and a concomitant decrease in PTH receptor expression were observed in young goats (4 to 5 months) when dietary protein was diminished. The concentration of 1,25-(OH)2D3 was reduced while PTH levels were not affected (Starke et al., Reference Starke, Cox, Sudekum and Huber2013; Starke and Huber, Reference Starke and Huber2014; Firmenich et al., Reference Firmenich, Elfers, Wilkens, Breves and Muscher-Banse2018). The stimulation of NaPi IIa expression during a protein-reduced diet is not obvious. It was postulated that a reduction in Pi concentrations in the ultrafiltrate stimulated the expression of NaPi IIa in apical membranes. The decline in Pi in the ultrafiltrate could be caused by a drop in the glomerular filtration rate (GFR) to conserve urea because a reduction in GFR by 60% was detected in goats fed a low protein diet (Eriksson and Valtonen, Reference Eriksson and Valtonen1982; Valtonen et al., Reference Valtonen, Uusi-Rauva and Eriksson1982). Therefore, an unknown Pi-sensing mechanism(s) in the proximal tubules must exist. Interestingly, a stimulation of NaPi IIa expression was accomplished by dietary protein reduction and thereby, presumably, a reduction in Pi in the ultrafiltrate. A direct dietary Pi depletion without manipulation of GFR did not show the same effects (Schröder et al., Reference Schröder, Walter, Breves and Huber2000).

In pre-ruminant animals, the kidneys are the main excretory pathway for an excess of Pi. During the development of the rumen, changes occurred. When the threshold of plasma Pi exceeded, renal elimination of Pi is neither stimulated nor eliminated, but more Pi is secreted in the saliva to the rumen, where it is used by microorganisms. Therefore, PTH-mediated regulation of renal Pi excretion is less important in adult ruminants than in growing ruminants.

In adult ruminants, renal Pi excretion does not seem to be regulated. An intravenous infusion of PTH did not alter renal excretion of Pi in sheep, and a dietary Ca restriction for several weeks did not affect fractional excretion of Pi in small ruminants (sheep: 1.23 ± 0.23% v. 0.82 ± 0.07%, goats: 2.26 ± 0.0.80% v. 2.79 ± 0.80%) (Clark et al., Reference Clark, French, Beal, Cross and Budtz-Olsen1975; Herm et al., Reference Herm, Muscher-Banse, Breves, Schröder and Wilkens2015). This is in line with former results from sheep and cows (Braithwaite, Reference Braithwaite1975; Taylor et al., Reference Taylor, Knowlton, McGilliard, Swecker, Ferguson, Wu and Hanigan2009).

Conclusions and perspectives

The regulation of mineral homeostasis in ruminants differs not only from monogastric animals but also between and within ruminant species. Although the molecular structures that are involved in Ca and Pi transports in the intestinal tract and the kidneys have been characterised in several ruminant species, the modulation of these by different dietary interventions, by the supply of other minerals and nutrients, or as a consequence of hormonal changes in 1,25-(OH)2D3, FGF23, PTH or calcitonin are still under investigation. Ruminal Ca transport mechanisms are still not clarified. In addition, more information is required in respect to the contribution of salivary mineral secretion and bone turnover. Further research is also needed to better understand imbalances of mineral homeostasis, such as hypocalcaemia and the capacities of ruminants to adapt to marginal mineral supply when kept on P-deficient pasture. In this regard, the interplay between mineral homeostasis, availability and digestibility of nutrients and metabolic pathways regulating energy and protein metabolism should be elucidated as they are important for lactating cows as well as animals kept for meat production.

Acknowledgements

The authors wish to thank Prof. David R. Fraser for proofreading the manuscript. The presented research was partly supported by the Deutsche Forschungsgemeinschaft (Schr 342/8-1, Schr 342/8-2, Wi 3668/1-1, Mu 3585/1-1, Mu 3585/1-3) and the H. Wilhelm Schaumann Foundation.

M. R. Wilkens 0000-0003-1734-5024

A. S. Muscher-Banse 0000-0002-4149-7987

Declaration of interest

The authors declare that there are no conflicts of interest.

Ethics statement

None.

Software and data repository resources

None of the data were deposited in an official repository.

References

Abdel-Hafeez, HM, Manas-Almendros, M, Ross, R, Care, AD and Marshall, DH 1982. Effects of dietary phosphorus and calcium on the intestinal absorption of Ca in sheep. British Journal of Nutrition 47, 6977.CrossRefGoogle Scholar
Antoniucci, DM, Yamashita, T and Portale, AA 2006. Dietary phosphorus regulates serum fibroblast growth factor-23 concentrations in healthy men. The Journal of Clinical Endocrinology and Metabolism 91, 31443149.CrossRefGoogle ScholarPubMed
Beardsworth, LJ, Beardsworth, PM and Care, AD 1989. The effect of ruminal phosphate concentration on the absorption of calcium, phosphorus and magnesium from the reticulo-rumen of the sheep. British Journal of Nutrition 61, 715723.CrossRefGoogle ScholarPubMed
Beckman, MJ and DeLuca, HF 2002. Regulation of renal vitamin D receptor is an important determinant of 1α,25-dihydroxyvitamin D3 levels in vivo. Archives of Biochemistry and Biophysics 401, 4452.CrossRefGoogle ScholarPubMed
Ben-awadh, AN, Delgado-Calle, J, Tu, X, Kuhlenschmidt, K, Allen, MR, Plotkin, LI and Bellido, T 2014. Parathyroid hormone receptor signaling induces bone resorption in the adult skeleton by directly regulating the RANKL gene in osteocytes. Endocrinology 155, 27972809.CrossRefGoogle ScholarPubMed
Besarab, A and Swanson, JW 1982. Differences in effects of amino-terminal and intact parathyroid hormone on calcium, phosphate, and cAMP excretion by the isolated perfused rat kidney. Renal Physiology 5, 261271.Google ScholarPubMed
Biber, J and Murer, H 1994. A molecular view of renal Na-dependent phosphate transport. Renal Physiology and Biochemistry 17, 212215.Google ScholarPubMed
Biber, J, Murer, H and Forster, I 1998. The renal type II Na+/phosphate cotransporter. Journal of Bioenergetics and Biomembranes 30, 187194.CrossRefGoogle Scholar
Braithwaite, GD 1975. Studies on the absorption and retention of calcium and phosphorus by young and mature Ca-deficient sheep. British Journal of Nutrition 34, 311324.CrossRefGoogle Scholar
Braun, HS, Schrapers, KT, Mahlkow-Nerge, K, Stumpff, F and Rosendahl, J 2019. Dietary supplementation of essential oils in dairy cows: evidence for stimulatory effects on nutrient absorption. Animal 13, 518523.CrossRefGoogle ScholarPubMed
Breves, G, Beyerbach, M, Holler, H and Lessmann, HW 1985. Fluid exchange in the rumen of sheep in low and adequate phosphorus administration. Deutsche tierärztliche Wochenschrift 92, 4749.Google ScholarPubMed
Breves, G, Höller, H, Packheiser, P, Gabel, G and Martens, H 1988. Flux of inorganic phosphate across the sheep rumen wall in vivo and in vitro. Quarterly Journal of Experimental Physiology 73, 343351.CrossRefGoogle ScholarPubMed
Breves, G and Schröder, B 1991. Comparative aspects of gastrointestinal phosphorus metabolism. Nutrition Research Reviews 4, 125140.CrossRefGoogle ScholarPubMed
Bronner, F 1987. Intestinal calcium absorption: mechanisms and applications. Journal of Nutrition 117, 13471352.CrossRefGoogle ScholarPubMed
Bronner, F and Pansu, D 1999. Nutritional aspects of calcium absorption. The Journal of Nutrition 129, 912.CrossRefGoogle ScholarPubMed
Brozos, C, Mavrogianni, VS and Fthenakis, GC 2011. Treatment and control of peri-parturient metabolic diseases: pregnancy toxemia, hypocalcemia, hypomagnesemia. Veterinary Clinics of North America: Food Animal Practice 27, 105113.Google Scholar
Bushinsky, DA, Nalbantian-Brandt, C and Favus, MJ 1989. Elevated Ca2+ does not inhibit the 1,25(OH)2D3 response to phosphorus restriction. American Journal of Physiology 256, F285F289.Google Scholar
Bushinsky, DA, Riera, GS, Favus, MJ and Coe, FL 1985. Evidence that blood ionized calcium can regulate serum 1,25(OH)2D3 independently of parathyroid hormone and phosphorus in the rat. Journal of Clinical Investigation 76, 15991604.CrossRefGoogle ScholarPubMed
Care, AD, Barlet, J-P and Abdel-Hafeez, HM 1980. Calcium and phosphate homoeostasis in ruminants and its relationship to the aetiology and prevention of parturient paresis. doi:10.1007/978-94-011-8067-2_20.CrossRefGoogle Scholar
Chen, K-S and DeLuca, HF 1995. Cloning of the human 1α,25-dihydroxyvitamin D-3 24-hydroxylase gene promoter and identification of two vitamin D-responsive elements. Biochimica et Biophysica Acta (BBA) - Gene Structure and Expression 1263, 19.CrossRefGoogle Scholar
Chirayath, MV, Gajdzik, L, Hulla, W, Graf, J, Cross, HS and Peterlik, M 1998. Vitamin D increases tight-junction conductance and paracellular Ca 2+ -transport in Caco-2 cell cultures. American Journal of Physiology 274, G389G396.Google ScholarPubMed
Clark, RC, French, TJ, Beal, AM, Cross, RB and Budtz-Olsen, OE 1975. The acute effects of intravenous infusion of parathyroid hormone on urine, plasma and saliva in the sheep. Quarterly Journal of Experimental Physiology and Cognate Medical Sciences 60, 95106.CrossRefGoogle ScholarPubMed
Cohrs, I, Wilkens, MR and Grunberg, W 2018. Short communication: effect of dietary phosphorus deprivation in late gestation and early lactation on the calcium homeostasis of periparturient dairy cows. Journal of Dairy Science 101, 95919598.CrossRefGoogle ScholarPubMed
Daniel, RC 1983. Motility of the rumen and abomasum during hypocalcaemia. Canadian Journal of Comparative Medicine 47, 276280.Google ScholarPubMed
Domino, AR, Korzec, HC and McArt, JAA 2017. Field trial of 2 calcium supplements on early lactation health and production in multiparous Holstein cows. Journal of Dairy Science 100, 96819690.CrossRefGoogle ScholarPubMed
Duric, M, Zhao, GY, Orskov, ER and Chen, XB 1994. Indirect measurement of saliva secretion in sheep fed diets of different structures and the effect of such diets on ruminal fluid kinetics and fermentation pattern. Experimental Physiology 79, 823830.CrossRefGoogle ScholarPubMed
Dusso, AS, Brown, AJ and Slatopolsky, E 2005. Vitamin D. American Journal of Physiology - Renal Physiology 289, F8F28.CrossRefGoogle ScholarPubMed
Elfers, K, Liesegang, A, Wilkens, MR, Breves, G and Muscher-Banse, AS 2016a. Dietary nitrogen and calcium modulate bone metabolism in young goats. The Journal of Steroid Biochemistry and Molecular Biology 164, 188193.CrossRefGoogle ScholarPubMed
Elfers, K, Marr, I, Wilkens, MR, Breves, G, Langeheine, M, Brehm, R and Muscher-Banse, AS 2016b. Expression of tight junction proteins and Cadherin 17 in the small intestine of young goats offered a reduced N and/or Ca diet. PLoS ONE 11, e0154311.CrossRefGoogle ScholarPubMed
Elfers, K, Wilkens, MR, Breves, G and Muscher-Banse, AS 2015. Modulation of intestinal calcium and phosphate transport in young goats fed a nitrogen- and/or calcium-reduced diet. British Journal of Nutrition 114, 19491964.CrossRefGoogle ScholarPubMed
Eriksson, L and Valtonen, M 1982. Renal urea handling in goats fed high and low protein diets. Journal of Dairy Science 65, 385389.CrossRefGoogle ScholarPubMed
Firmenich, CS, Elfers, K, Wilkens, MR, Breves, G and Muscher-Banse, AS 2018. Modulation of renal Ca and Pi transporting proteins by dietary nitrogen and/or calcium in young goats. Journal of Animal Science 96, 32083220.Google ScholarPubMed
Foote, AP, Lambert, BD, Brady, JA and Muir, JP 2011. Phosphate transporter expression in Holstein cows. Journal of Dairy Science 94, 19131916.CrossRefGoogle ScholarPubMed
Fraser, DR and Kodicek, E 1973. Regulation of 25-hydroxycholecalciferol-1-hydroxylase activity in kidney by parathyroid hormone. Nature: New Biology 241, 163166.Google ScholarPubMed
Frick, KK, Asplin, JR, Favus, MJ, Culbertson, C, Krieger, NS and Bushinsky, DA 2013. Increased biological response to 1,25(OH)2D3 in genetic hypercalciuric stone-forming rats. American Journal of Physiology - Renal Physiology 304, F718F726.CrossRefGoogle Scholar
Friedman, PA and Gesek, FA 1995. Cellular calcium transport in renal epithelia: measurement, mechanisms, and regulation. Physiological Reviews 75, 429471.CrossRefGoogle ScholarPubMed
Goff, JP 2008. The monitoring, prevention, and treatment of milk fever and subclinical hypocalcemia in dairy cows. The Veterinary Journal 176, 5057.CrossRefGoogle ScholarPubMed
Goff, JP 2018. Invited review: mineral absorption mechanisms, mineral interactions that affect acid-base and antioxidant status, and diet considerations to improve mineral status. Journal of Dairy Science 101, 27632813.CrossRefGoogle ScholarPubMed
Goff, JP, Liesegang, A and Horst, RL 2014. Diet-induced pseudohypoparathyroidism: a hypocalcemia and milk fever risk factor. Journal of Dairy Science 97, 15201528.CrossRefGoogle ScholarPubMed
Grünberg, W, Donkin, SS and Constable, PD 2011. Periparturient effects of feeding a low dietary cation-anion difference diet on acid-base, calcium, and phosphorus homeostasis and on intravenous glucose tolerance test in high-producing dairy cows. Journal of Dairy Science 94, 727745.CrossRefGoogle ScholarPubMed
Herm, G, Muscher-Banse, AS, Breves, G, Schröder, B and Wilkens, MR 2015. Renal mechanisms of calcium homeostasis in sheep and goats. Journal of Animal Science 93, 16081621.CrossRefGoogle ScholarPubMed
Hernández-Castellano, LE, Hernandez, LL and Bruckmaier, RM 2019. Review: endocrine pathways to regulate calcium homeostasis around parturition and the prevention of hypocalcemia in periparturient dairy cows. Animal 19. doi:10.1017/S1751731119001605.Google ScholarPubMed
Hilfiker, H, Hattenhauer, O, Traebert, M, Forster, I, Murer, H and Biber, J 1998. Characterization of a murine type II sodium-phosphate cotransporter expressed in mammalian small intestine. Proceedings of the National Academy of Science of the United States of America 95, 1456414569.CrossRefGoogle ScholarPubMed
Hoenderop, JG, Dardenne, O, van Abel, M, van der Kemp, AW, van Os, CH, St. Arnaud, R and Bindels, RJ 2002. Modulation of renal Ca2+ transport protein genes by dietary Ca2+ and 1,25-dihydroxyvitamin D3 in 25-hydroxyvitamin D3-1alpha-hydroxylase knockout mice. FASEB Journal 16, 13981406.CrossRefGoogle ScholarPubMed
Hoenderop, JG, Nilius, B and Bindels, RJ 2005. Calcium absorption across epithelia. Physiological Reviews 85, 373422.CrossRefGoogle ScholarPubMed
Höller, H, Breves, G and Dubberke, M 1988b. Flux of inorganic phosphate and calcium across the isolated mucosa of the sheep omasum. Zentralblatt für Veterinärmedizin Reihe A 35, 709716.Google ScholarPubMed
Höller, H, Breves, G, Kocabatmaz, M and Gerdes, H 1988a. Flux of calcium across the sheep rumen wall in vivo and in vitro. Quarterly Journal of Experimental Physiology 73, 609618.CrossRefGoogle ScholarPubMed
Höller, H, Figge, A, Richter, J and Breves, G 1988c. Calcium and inorganic phosphate net absorption from the sheep colon and rectum perfused in vivo. Journal of Animal Physiology and Animal Nutrition 59, 915.CrossRefGoogle Scholar
Huber, K, Roesler, U, Holthausen, A, Pfeffer, E and Breves, G 2007. Influence of dietary calcium and phosphorus supply on epithelial phosphate transport in preruminant goats. Journal of Comparative Physiology B 177, 193203.CrossRefGoogle ScholarPubMed
Huber, K, Roesler, U, Muscher, A, Hansen, K, Widiyono, I, Pfeffer, E and Breves, G 2003. Ontogenesis of epithelial phosphate transport systems in goats. American Journal of Physiology - Regulatory, Integrative and Comparative Physiology 284, R413R421.CrossRefGoogle ScholarPubMed
Huber, K, Walter, C, Schröder, B, Biber, J, Murer, H and Breves, G 2000. Epithelial phosphate transporters in small ruminants. Annals of the New York Academy of Sciences 915, 9597.CrossRefGoogle ScholarPubMed
Huber, K, Walter, C, Schröder, B and Breves, G 2002. Phosphate transport in the duodenum and jejunum of goats and its adaptation by dietary phosphate and calcium. American Journal of Physiology - Regulatory, Integrative and Comparative Physiology 283, R296R302.CrossRefGoogle ScholarPubMed
Hyde, ML and Fraser, DR 2014. In vivo measurement of the absorption of strontium in the rumen and small intestine of sheep as an index of calcium absorption capacity. British Journal of Nutrition 112, 718724.CrossRefGoogle ScholarPubMed
Hyde, ML, Wilkens, MR and Fraser, DR 2019. In vivo measurement of strontium absorption from the rumen of dairy cows as an index of calcium absorption capacity. Journal of Dairy Science 102, 56995705.CrossRefGoogle ScholarPubMed
Isac, MD, Valverde, A, Lupiani, MJ, Martinez de Victoria, E and Manas, M 1989. The influence of PTH, calcium gluconate and EDTA on the parotid saliva in the goat. Acta Physiologica Hungarica 74, 295304.Google ScholarPubMed
Karbach, U 1992. Paracellular calcium transport across the small intestine. Journal of Nutrition 122, 672677.CrossRefGoogle ScholarPubMed
Khorasani, GR and Armstrong, DG 1990. Effect of sodium and potassium level on the absorption of magnesium and other macrominerals in sheep. Livestock Production Science 24, 223235.CrossRefGoogle Scholar
Khorasani, GR and Armstrong, DG 1992. Calcium, phosphorus, and magnesium absorption and secretion in the bovine digestive tract as influenced by dietary concentration of these elements. Livestock Production Science 31, 271286.CrossRefGoogle Scholar
Khorasani, GR, Janzen, RA, McGill, WB and Kennelly, JJ 1997. Site and extent of mineral absorption in lactating cows fed whole-crop cereal grain silage of alfalfa silage. Journal of Animal Science 75, 239248.CrossRefGoogle ScholarPubMed
Klinger, S, Schröder, B, Gemmer, A, Reimers, J, Breves, G, Herrmann, J and Wilkens, MR 2016. Gastrointestinal transport of calcium and glucose in lactating ewes. Physiological Reports 4, e12817. doi:10.14814/phy2.12817.CrossRefGoogle ScholarPubMed
Ko, SH, Lee, GS, Vo, TT, Jung, EM, Choi, KC, Cheung, KW, Kim, JW, Park, JG, Oh, GT and Jeung, EB 2009. Dietary calcium and 1,25-dihydroxyvitamin D3 regulate transcription of calcium transporter genes in calbindin-D9k knockout mice. Journal of Reproduction and Development 55, 137142.CrossRefGoogle ScholarPubMed
Krajisnik, T, Björklund, P, Marsell, R, Ljunggren, Ö, Åkerström, G, Jonsson, KB, Westin, G and Larsson, TE 2007. Fibroblast growth factor-23 regulates parathyroid hormone and 1α-hydroxylase expression in cultured bovine parathyroid cells. Journal of Endocrinology 195, 125131.CrossRefGoogle ScholarPubMed
Kumar, R and Thompson, JR 2011. The regulation of parathyroid hormone secretion and synthesis. Journal of the American Society of Nephrology 22, 216224.CrossRefGoogle ScholarPubMed
Kurbel, S, Radic, R, Kotromanovic, Z, Puseljic, Z and Kratofil, B 2003. A calcium homeostasis model: orchestration of fast acting PTH and calcitonin with slow calcitriol. Medical Hypotheses 61, 346350.CrossRefGoogle ScholarPubMed
Kutuzova, GD and DeLuca, HF 2004. Gene expression profiles in rat intestine identify pathways for 1,25-dihydroxyvitamin D3 stimulated calcium absorption and clarify its immunomodulatory properties. Archives of Biochemistry and Biophysics 432, 152166.CrossRefGoogle ScholarPubMed
Lassiter, WE, Gottschalk, CW and Mylle, M 1963. Micropuncture study of renal tubular reabsorption of calcium in normal rodents. American Journal of Physiology-Legacy Content 204, 771775.CrossRefGoogle Scholar
Lean, IJ, DeGaris, PJ, McNeil, DM and Block, E 2006. Hypocalcemia in dairy cows: meta-analysis and dietary cation anion difference theory revisited. Journal of Dairy Science 89, 669684.CrossRefGoogle ScholarPubMed
Lutz, T and Scharrer, E 1991. Effect of short chain fatty acids on calcium absorption by the rat colon. Experimental Physiology 76, 615618.CrossRefGoogle ScholarPubMed
Macrae, AI, Whitaker, DA, Burrough, E, Dowell, A and Kelly, JM 2006. Use of metabolic profiles for the assessment of dietary adequacy in UK dairy herds. Veterinary Record 159, 655661.CrossRefGoogle ScholarPubMed
Mañas-Almendros, M, Ross, R and Care, AD 1982. Factors affecting the secretion of phosphate in parotid saliva in the sheep and goat. Quarterly Journal of Experimental Physiology 67, 269280.CrossRefGoogle ScholarPubMed
Martz, FA, Belo, AT, Weiss, MF and Belyea, RL 1999. True absorption of calcium and phosphorus from corn silage fed to nonlactating, pregnant dairy cows. Journal of Dairy Science 82, 618622.CrossRefGoogle ScholarPubMed
McGrath, JJ, Savage, DB, Nolan, JV and Elliott, R 2012. Phosphorus and calcium retention in steers fed a roughage diet is influenced by dietary 25OH-vitamin D. Animal Production Science 52, 636640.CrossRefGoogle Scholar
Menard, L and Thompson, A 2007. Milk fever and alert downer cows: does hypophosphatemia affect the treatment response? Canadian Veterinary Journal 48, 487491.Google ScholarPubMed
Murer, H, Forster, I and Biber, J 2004. The sodium phosphate cotransporter family SLC34. Pflügers Archiv 447, 763767.CrossRefGoogle ScholarPubMed
Murer, H, Forster, I, Hernando, N, Lambert, G, Traebert, M and Biber, J 1999. Posttranscriptional regulation of the proximal tubule NaPi-II transporter in response to PTH and dietary P(i). American Journal of Physiology 277, F676F684.Google Scholar
Müschen, H, Petri, A, Breves, G, Pfeffer, E 1988. Response of lactating goats to low phosphorus intake : [part] 1, milk yield and faecal excretion of P and Ca. The Journal of Agricultural Science 111, 255263.CrossRefGoogle Scholar
Muscher, A, Hattendorf, J, Pfeffer, E, Breves, G and Huber, K 2008. Hormonal regulation of phosphate homeostasis in goats during transition to rumination. Journal of Comparative Physiology B 178, 585596.CrossRefGoogle ScholarPubMed
Muscher, A and Huber, K 2010. Effects of a reduced nitrogen diet on calcitriol levels and calcium metabolism in growing goats. The Journal of Steroid Biochemistry and Molecular Biology 121, 304307.CrossRefGoogle ScholarPubMed
Muscher, AS, Piechotta, M, Breves, G and Huber, K 2011. Modulation of electrolyte homeostasis by dietary nitrogen intake in growing goats. British Journal of Nutrition 105, 16191626.CrossRefGoogle ScholarPubMed
Negri, AL 2015. Role of claudins in renal calcium handling. Nefrología 35, 347352.CrossRefGoogle ScholarPubMed
Neves, RC, Leno, BM, Bach, KD and McArt, JAA 2018. Epidemiology of subclinical hypocalcemia in early-lactation Holstein dairy cows: the temporal associations of plasma calcium concentration in the first 4 days in milk with disease and milk production. Journal of Dairy Science 101, 93219331.CrossRefGoogle ScholarPubMed
Oehlschlaeger, V, Wilkens, M, Schröder, B, Daenicke, S and Breves, G 2014. Effects of 25-hydroxyvitamin D3 on localisation and extent of gastrointestinal calcium absorption in dairy cattle. Animal Production Science 54, 13941398.CrossRefGoogle Scholar
Oetzel, GR 1988. Parturient paresis and hypocalcemia in ruminant livestock. Veterinary Clinics of North America: Food Animal Practice 4, 351364.Google ScholarPubMed
Pfeffer, E, Rodehutscord, M, Breves, G 1995. Effects of reducing dietary calcium and/or phosphorus on performance and body composition in male kids. Journal of Animal Physiology and Animal Nutrition 74, 243252.CrossRefGoogle Scholar
Pfeffer, E, Thompson, A and Armstrong, DG 1970. Studies on intestinal digestion in the sheep. 3. Net movement of certain inorganic elements in the digestive tract on rations containing different proportions of hay and rolled barley. British Journal of Nutrition 24, 197204.CrossRefGoogle ScholarPubMed
Piechotta, M, Holzhausen, L, Araujo, MG, Heppelmann, M, Sipka, A, Pfarrer, C, Schuberth, HJ and Bollwein, H 2014. Antepartal insulin-like growth factor concentrations indicating differences in the metabolic adaptive capacity of dairy cows. Journal of Veterinary Science 15, 343352.CrossRefGoogle ScholarPubMed
Reinhardt, TA, Lippolis, JD, McCluskey, BJ, Goff, JP and Horst, RL 2011. Prevalence of subclinical hypocalcemia in dairy herds. Veterinary Journal 188, 122124.CrossRefGoogle ScholarPubMed
Reist, M, Erdin, D, von Euw, D, Tschuemperlin, K, Leuenberger, H, Delavaud, C, Chilliard, Y, Hammon, HM, Kuenzi, N and Blum, JW 2003. Concentrate feeding strategy in lactating dairy cows: metabolic and endocrine changes with emphasis on leptin. Journal of Dairy Science 86, 16901706CrossRefGoogle Scholar
Riad, F, Lefaivre, J and Barlet, JP 1987. 1,25-Dihydroxycholecalciferol regulates salivary phosphate secretion in cattle. Journal of Endocrinology 112, 427430.CrossRefGoogle ScholarPubMed
Rosendahl, J, Braun, HS, Schrapers, KT, Martens, H and Stumpff, F 2016. Evidence for the functional involvement of members of the TRP channel family in the uptake of Na+ and NH4+ by the ruminal epithelium. Pflügers Archiv 468, 13331352.CrossRefGoogle ScholarPubMed
Saier, MH, Jr 2000. A functional-phylogenetic classification system for transmembrane solute transporters. Microbiology and Molecular Biology Reviews 64, 354411.CrossRefGoogle ScholarPubMed
Saji, F, Shigematsu, T, Sakaguchi, T, Ohya, M, Orita, H, Maeda, Y, Ooura, M, Mima, T and Negi, S 2010. Fibroblast growth factor 23 production in bone is directly regulated by 1α,25-dihydroxyvitamin D, but not PTH. American Journal of Physiology-Renal Physiology 299, F1212F1217.CrossRefGoogle Scholar
Scharrer, E 1985. Phospahte absorption at different intestinal sites in the devloping lamb. Quarterly Journal of Experimental Physiology 70, 615621.CrossRefGoogle Scholar
Schiavi, SC and Kumar, R 2004. The phosphatonin pathway: new insights in phosphate homeostasis. Kidney International 65, 114.CrossRefGoogle ScholarPubMed
Schrapers, KT, Sponder, G, Liebe, F, Liebe, H and Stumpff, F 2018. The bovine TRPV3 as a pathway for the uptake of Na+, Ca2+, and NH4+. PLoS ONE 13, e0193519.CrossRefGoogle ScholarPubMed
Schröder, B and Breves, G 1996. Mechanisms of phosphate uptake into brush-border membrane vesicles from goat jejunum. Journal of Comparative Physiology B 166, 230240.CrossRefGoogle ScholarPubMed
Schröder, B and Breves, G 2006. Mechanisms and regulation of calcium absorption from the gastrointestinal tract in pigs and ruminants: comparative aspects with special emphasis on hypocalcemia in dairy cows. Animal Health Research Reviews 7, 3141.CrossRefGoogle ScholarPubMed
Schröder, B, Goebel, W, Huber, K and Breves, G 2001. No effect of vitamin D3 treatment on active calcium absorption across ruminal epithelium of sheep. Journal of Veterinary Medicine Series A Physiology Pathology Clinical Medicine 48, 353363.CrossRefGoogle ScholarPubMed
Schröder, B, Kappner, H, Failing, K, Pfeffer, E and Breves, G 1995. Mechanisms of intestinal phosphate transport in small ruminants. British Journal of Nutrition 74, 635648.CrossRefGoogle ScholarPubMed
Schröder, B, Rittmann, I, Pfeffer, E and Breves, G 1997. In vitro studies on calcium absorption from the gastrointestinal tract in small ruminants. Journal of Comparative Physiology B 167, 4351.Google ScholarPubMed
Schröder, B, Walter, C, Breves, G and Huber, K 2000. Comparative studies on Na-dependent Pi transport in ovine, caprine and porcine renal cortex. Journal of Comparative Physiology B 170, 387393.Google ScholarPubMed
Schröder, B, Wilkens, MR, Ricken, GE, Leonhard-Marek, S, Fraser, DR and Breves, G 2015. Calcium transport in bovine rumen epithelium as affected by luminal Ca concentrations and Ca sources. Physiological Reports 3, e12615. doi:10.14814/phy2.12615.CrossRefGoogle ScholarPubMed
Scott, D and Beastall, G 1978. The effects of intravenous phosphate loading on salivary phosphate secretion and plasma parathyroid hormone levels in the sheep. Quarterly Journal of Experimental Physiology and Cognate Medical Sciences 63, 147156.CrossRefGoogle ScholarPubMed
Shirazi-Beechey, SP, Beechey, RB, Penny, J, Vayro, S, Buchan, W and Scott, D 1991. Mechanisms of phosphate transport in sheep intestine and parotid gland: response to variation in dietary phosphate supply. Experimental Physiology 76, 231241.CrossRefGoogle ScholarPubMed
Shirazi-Beechey, SP, Kemp, RB, Dyer, J and Beechey, RB 1989. Changes in the functions of the intestinal brush border membrane during the development of the ruminant habit in lambs. Comparative Biochemistry and Physiology Part B 94, 801806.CrossRefGoogle ScholarPubMed
Shirazi-Beechey, SP, Penny, JI, Dyer, J, Wood, IS, Tarpey, PS, Scott, D and Buchan, W 1996. Epithelial phosphate transport in ruminants, mechanisms and regulation. Kidney International 49, 992996.CrossRefGoogle ScholarPubMed
Sidler-Lauff, K, Boos, A, Kraenzlin, M and Liesegang, A 2010. Influence of different calcium supplies and a single vitamin D injection on vitamin D receptor- and calbindin D9k-immunoreactivities in the gastrointestinal tract of goat kids. Journal of Animal Science 88, 35983610.CrossRefGoogle Scholar
Starke, S, Cox, C, Sudekum, KH and Huber, K 2013. Adaptation of electrolyte handling to low crude protein intake in growing goats and consequences for in vivo electrolyte excretion. Small Ruminant Research 114, 9096.CrossRefGoogle Scholar
Starke, S and Huber, K 2014. Adaptive responses of calcium and phosphate homeostasis in goats to low nitrogen intake: renal aspects. Journal of Animal Physiology and Animal Nutrition 98, 853859.CrossRefGoogle ScholarPubMed
Starke, S, Reimers, J, Muscher-Banse, AS, Schröder, B, Breves, G and Wilkens, MR 2016. Gastrointestinal transport of calcium and phosphate in lactating goats. Livestock Science 189, 2331.CrossRefGoogle Scholar
Sutton, RA, Wong, NL and Dirks, JH 1979. Effects of metabolic acidosis and alkalosis on sodium and calcium transport in the dog kidney. Kidney International 15, 520533.CrossRefGoogle ScholarPubMed
Taylor, MS, Knowlton, KF, McGilliard, ML, Swecker, WS, Ferguson, JD, Wu, Z and Hanigan, MD 2009. Dietary calcium has little effect on mineral balance and bone mineral metabolism through twenty weeks of lactation in Holstein cows. Journal of Dairy Science 92, 223237.CrossRefGoogle ScholarPubMed
Uppal, SK, Wolf, K, Khahra, SS and Martens, H 2003b. Modulation of Na+ transport across isolated rumen epithelium by short-chain fatty acids in hay- and concentrate-fed sheep. Journal of Animal Physiology and Animal Nutrition 87, 380388.CrossRefGoogle ScholarPubMed
Uppal, SK, Wolf, K and Martens, H 2003a. The effect of short chain fatty acids on calcium flux rates across isolated rumen epithelium of hay-fed and concentrate-fed sheep. Journal of Animal Physiology and Animal Nutrition 87, 1220.CrossRefGoogle ScholarPubMed
Valtonen, MH, Uusi-Rauva, A and Eriksson, L 1982. The effect of protein deprivation on the validity of creatinine and urea in evaluation of renal function. An experimental study in the goat. Scandinavian Journal of Clinical and Laboratory Investigation 42, 507512.CrossRefGoogle ScholarPubMed
van Cromphaut, SJ, Rummens, K, Stockmans, I, van Herck, E, Dijcks, FA, Ederveen, AG, Carmeliet, P, Bouillon, R and Carmeliet, G 2003. Intestinal calcium transporter genes are upregulated by estrogens and the reproductive cycle through vitamin D receptor-independent mechanisms. Journal of Bone and Mineral Research 18, 17251736.CrossRefGoogle ScholarPubMed
van Doorn, DA, van der Spek, ME, Everts, H, Wouterse, H and Beynen, AC 2004. The influence of calcium intake on phosphorus digestibility in mature ponies. Journal of Animal Physiology and Animal Nutrition 88, 412418.CrossRefGoogle ScholarPubMed
van´t Klooster, AT 1976. Adaptation of calcium absorption from the small intestine of dairy cows to changes in the dietary calcium intake and at the onset of lactation. Zeitschrift für Tierphysiologie Tierernährung und Futtermittelkunde 37, 169182.CrossRefGoogle Scholar
Venjakob, PL, Borchardt, S and Heuwieser, W 2017. Hypocalcemia-Cow-level prevalence and preventive strategies in German dairy herds. Journal of Dairy Science 100, 92589266.CrossRefGoogle ScholarPubMed
Widiyono, I, Huber, K, Failing, K and Breves, G 1998. Renal phosphate excretion in goats. Zentralblatt für Veterinärmedizin Reihe A 45, 145153.CrossRefGoogle ScholarPubMed
Wilkens, MR, Breves, G and Schröder, B 2014. A goat is not a sheep: physiological similarities and differences observed in two ruminant species facing a challenge of calcium homeostatic mechanisms. Animal Production Science 54, 15071511.CrossRefGoogle Scholar
Wilkens, MR, Elfers, K, Schmicke, M, Breves, G and Muscher-Banse, AS 2018. Dietary nitrogen and calcium modulate CYP27B1 expression in young goats. Domestic Animal Endocrinology 64, 7076.CrossRefGoogle ScholarPubMed
Wilkens, MR, Kunert-Keil, C, Brinkmeier, H and Schröder, B 2009. Expression of calcium channel TRPV6 in ovine epithelial tissue. Veterinary Journal 182, 294300.CrossRefGoogle ScholarPubMed
Wilkens, MR, Marholt, L, Eigendorf, N, Muscher-Banse, AS, Feige, K, Schröder, B, Breves, G and Cehak, A 2017. Trans- and paracellular calcium transport along the small and large intestine in horses. Comparative Biochemistry and Physiology Part A: Molecular and Integrative Physiology 204, 157163.CrossRefGoogle ScholarPubMed
Wilkens, MR, Mrochen, N, Breves, G and Schröder, B 2011. Gastrointestinal calcium absorption in sheep is mostly insensitive to an alimentary induced challenge of calcium homeostasis. Comparative Biochemistry and Physiology – Part B Biochemical and Molecular Biology 158, 199207.CrossRefGoogle Scholar
Wilkens, MR, Oberheide, I, Schröder, B, Azem, E, Steinberg, W and Breves, G 2012a. Influence of the combination of 25-hydroxyvitamin D3 and a diet negative in cation-anion difference on peripartal calcium homeostasis of dairy cows. Journal of Dairy Science 95, 151164.CrossRefGoogle Scholar
Wilkens, MR, Praechter, C, Breves, G and Schröder, B 2016. Stimulating effects of a diet negative in dietary cation-anion difference on calcium absorption from the rumen in sheep. Journal of Animal Physiology and Animal Nutrition 100, 156166.CrossRefGoogle Scholar
Wilkens, MR, Richter, J, Fraser, DR, Liesegang, A, Breves, G and Schröder, B 2012b. In contrast to sheep, goats adapt to dietary calcium restriction by increasing intestinal absorption of calcium. Comparative Biochemistry and Physiology Part A: Molecular and Integrative Physiology 163, 396406.CrossRefGoogle Scholar
Wilson, AD and Tribe, DE 1963. Effect of diet on secretion of parotid saliva by sheep. 1. Daily secretion of saliva by caged sheep. Australian Journal of Agricultural Research 14, 670679.CrossRefGoogle Scholar
Wright, RD, Blair-West, JR, Nelson, JF and Tregear, GW 1982. Local action of parathyroid hormone (1-34) and (1-84) at the parotid gland of sheep. Journal of Endocrinology 94, 3741.CrossRefGoogle ScholarPubMed
Yamagishi, N, Miyazaki, M and Naito, Y 2006. The expression of genes for transepithelial calcium-transporting proteins in the bovine duodenum. Veterinary Journal 171, 363366.CrossRefGoogle ScholarPubMed
Yeh, BI, Sun, TJ, Lee, JZ, Chen, HH and Huang, CL 2003. Mechanism and molecular determinant for regulation of rabbit transient receptor potential type 5 (TRPV5) channel by extracellular pH. Journal of Biological Chemistry 278, 5104451052.CrossRefGoogle ScholarPubMed
Zhang, Y, Lai, WP, Leung, PC, Che, CT and Wong, MS 2008. Improvement of Ca balance by Fructus Ligustri Lucidi extract in aged female rats. Osteoporosis International 19, 235242.CrossRefGoogle ScholarPubMed
Figure 0

Table 1 Results from balance studies done with different ruminant species: intake, urinary excretion (UEX), pre-intestinal (PRE) and intestinal (INT) net absorption (ABS), faecal excretion (FEX) in grams per day, apparent digestibility (AD) in percentage

Figure 1

Figure 1 Unidirectional flux rates (J) from mucosal to serosal (ms) and from serosal to mucosal (sm) of Ca as a function of those of mannitol (Man) in the rumen tissues of sheep (n = 20) and goats (n = 20) determined using the Ussing chamber in the absence of any electrochemical gradient. As mannitol is used as a marker for paracellular transport of water, the lack of any relationship between Ca J ms and Man J ms indicates transcellular Ca absorption. Modified from Wilkens et al. (2011) and (2012b).

Figure 2

Figure 2 Ca transport mechanisms and transepithelial potential difference in the rumen, small intestine, the thick ascending limp of the loop of Henle (TAL), and the distal and connecting tubules (DT, CT) of the kidneys in ruminant species. PMCA, plasma membrane Ca2+-ATPase isoform 1b; TRPV6/5, transient receptor potential vanilloid channel type 6/5; CaBPD9K/CaBPD28K, calbindin-D9K/D9K; NKCC2, Na+-K+-Cl co-transporter type 2; ROMK1, renal outer medullary K+ channel type 1; NCX, Na+/Ca2+ exchanger type 1. Explanations of the mechanisms are given in the corresponding text.

Figure 3

Figure 3 Rumen Ca net flux rates (Jnet) of female sheep and goats aged 6 to 7 months kept on adequate (con, 0.92% and 1.10%, n = 5) or restricted Ca supply (Ca–, 0.26% and 0.22%, n = 5) treated with a placebo or fed the same diets and treated with 1,25-dihydroxyvitamin D3 (1,25-(OH)2D3, n = 5) (0.5 µg/kg body weight) 12 h before sacrifice determined in Ussing chambers in the absence of any electrochemical gradient. Means ± SEM. Modified from Wilkens et al. (2011) and (2012b).

Figure 4

Figure 4 Correlation between electrical driving force and unidirectional Ca fluxes (J) from mucosal to serosal (ms) and from serosal to mucosal (sm) of castrated male sheep aged 8 months kept either on a ration positive in dietary cation–anion difference (DCAD) (control, n = 4) or negative in DCAD (low DCAD, n = 5). The electroneutral component of Jms represented by the intercept of the linear function revealed by regression analysis is greater (P < 0.01) in sheep kept on a diet low in DCAD (control: Jms = 7.76 (±1.23) + 7.53 (±0.77)·ξ–0.5; low DCAD: Jms = 13.32 (±4.42) + 9.95 (±2.77)·ξ–0.5). Means ± SEM. Modified from Wilkens et al. (2016).

Figure 5

Figure 5 Intestinal Ca net flux rates (Jnet) of horses of both sexes, aged 3 to 22 years (n = 10), female sheep (n = 5) and female goats (n = 5) aged 6 to 7 months kept on adequate Ca supply determined in Ussing chambers in the absence of any electrochemical gradient. Means ± SEM. Modified from Wilkens et al. (2011), (2012b) and (2017).

Figure 6

Figure 6 Intestinal Ca net flux rates (Jnet) of female sheep and goats aged 6 to 7 months kept on adequate (control, 0.92% and 1.10%, n = 5) or restricted Ca supply (Ca restriction, 0.26% and 0.22%, n = 5) and male goats aged 3 to 4 months kept on adequate (control, 22% CP, n = 7) or restricted n supply (n restriction, 8% CP, n = 6) determined in Ussing chambers in the absence of any electrochemical gradient. Significant differences revealed by the Student’s t test are marked with asterisks. Means ± SEM; *, P < 0.05. Modified from Elfers et al. (2015), Wilkens et al. (2011) and (2012b).

Figure 7

Figure 7 Pi transport mechanisms and transepithelial potential difference in the small intestine and proximal tubule of the kidneys in ruminant species. Apical entry occurs through Na-dependent Pi transporter family (NaPi) subtypes IIa and IIc or IIb and Na-dependent phosphate transporter 1 (PiT1). Basolateral extrusion mechanism of Pi is currently unknown. Further explanations of the mechanisms are given in the corresponding text.

Figure 8

Figure 8 Unidirectional, duodenal (Duo) and jejunal (Jeju) Pi flux rates (J) from serosal to mucosal (sm) and from mucosal to serosal (ms) of dried-off (n = 6) and lactating goats (n = 6) determined in Ussing chambers in the absence of any electrochemical gradient. Significant differences revealed by the Student’s t test are marked with asterisks. Means ± SEM; *, P < 0.05. Modified from Starke et al. (2016).