Skip to main content
  • Research article
  • Open access
  • Published:

Biomarkers of selenium status in dogs

Abstract

Background

Inadequate dietary selenium (Se) intake in humans and animals can lead to long term health problems, such as cancer. In view of the ownerā€™s desire for healthy longevity of companion animals, the impact of dietary Se provision on long term health effects warrants investigation. Little is currently known regards biomarkers, and rate of change of such biomarkers in relation to dietary selenium intake in dogs. In this study, selected biomarkers were assessed for their suitability to detect changes in dietary Se in adult dogs within eight weeks.

Results

Twenty-four dogs were fed a semi-purified diet with an adequate amount of Se (46.1Ā Ī¼g/MJ) over an 8Ā week period. They were then divided into two groups. The first group remained on the adequate Se diet, the second were offered a semi-purified diet with a low Se concentration (6.5Ā Ī¼g/MJ; 31Ā % of the FEDIAF minimum) for 8Ā weeks. Weekly urine and blood was collected and hair growth measurements were performed. The urinary Se to creatinine ratio and serum Se concentration were significantly lower in dogs consuming the low Se diet from week 1 onwards, by 84Ā % (adequate 25.3, low 4.1) and 7Ā % (adequate 257Ā Ī¼g/L, low 238Ā Ī¼g/L) respectively. Serum and whole blood glutathione peroxidase were also significantly lower in dogs consuming the low Se diet from weeks 6 and 8 respectively. None of the other biomarkers (mRNA expression and serum copper, creatine kinase, triiodothyronine:thyroxine ratio and hair growth) responded significantly to the low Se diet over the 8Ā week period.

Conclusions

This study demonstrated that urinary Se to creatinine ratio, serum Se and serum and whole blood glutathione peroxidase can be used as biomarkers of selenium status in dogs. Urinary Se to creatinine ratio and serum Se concentrations responded faster to decreased dietary Se than the other parameters. This makes these biomarkers candidates for early screening of long term effects of dietary Se provision on canine health.

Background

Selenium (Se) is an essential trace element in dogs [1]. It is involved in many aspects of canine physiology, such as anti-oxidant protection [2], thyroid hormone metabolism [3], and immune function [4]. Although no naturally occurring clinical cases of Se deficiency or toxicity have been reported in dogs [1], an inadequate Se status is associated with long term health effects, such as calcium oxalate calculi [5, 6] and cancer [7, 8] formation.

Considering the importance dog owners place on the healthy longevity of their pets [9, 10], the health effects of dietary Se intake should be studied in more detail. Measuring only the dietary Se content is not sufficient to study effects of Se on metabolism as the bioactivity of Se, defined as the amount of dietary Se that can be incorporated into selenoproteins such as glutathione peroxidase (GPx), is affected by many factors [11]. Long term studies are costly and difficult to perform, and therefore biomarkers are needed which can identify Se-induced changes in metabolism at an early stage.

There is currently no literature on specific and sensitive biomarkers of dietary Se intake in dogs, which confounds the accurate assessment of Se status. The main tissues and body fluids for minimally-invasive measurement of Se concentration are whole blood, plasma, serum, erythrocytes, urine, hair, and nails [12]. In this study biomarkers were selected to assess their sensitivity to a manipulation of dietary Se concentration in adult dogs with an adequate Se status.

Glutathione peroxidase (GPx) is the biomarker most often measured to estimate Se bioactivity [13ā€“15]. GPx is a selenoprotein that acts as an anti-oxidant [16] and is currently used as a proxy of selenium status, although has never been verified as the gold standard. GPx measurements in chicks receiving different concentrations of Se supplement [17] was used to inform the existing European minimum recommendation for the Se concentration in dog foods, which is 17.9Ā Ī¼g/MJ [18].

Secondly, Levander et al. [19] reported that urinary Se excretion in humans consuming a low Se diet (intake of approximately 35Ā Ī¼g/day) stabilised after 12Ā days. In cats, urinary Se concentrations increased dose-dependently within 2Ā days of Se supplementation with either sodium selenite or Se yeast at 1.5 and 2.0Ā mg Se/kg dry matter (DM) compared to 0.45 and 1.0Ā mg Se/kg DM [20].

The mRNA expression of one of the isoforms of GPx (GPx1) from liver tissue has also been reported to be a biomarker of Se status in rats [21, 22], together with selenoprotein H (SelH), selenoprotein W (SepW1), thioredoxin reductase 1 (TrxRd1), thioredoxin reductase 2 (TrxRd2), iodothyronine deiodinase 1 (DIO1), selenoprotein K (SelK), selenoprotein T (SelT), and 15Ā kDa selenoprotein (Sep15) [21, 22]. Although selenoprotein P is the most abundant selenoprotein [23] and is important in the transport of Se throughout the body [24], its mRNA has been shown to not be significantly down-regulated in Se deficiency [25].

Data from rats indicates that mRNA from whole blood can be used to measure mRNA biomarkers for Se status [26]. RNA isolated from whole blood for the determination of GPx1 mRNA was expressed at levels comparable to the levels found in liver, kidney and heart mRNA [26]. The expression of GPx1 mRNA in rats on a Se deficient torula yeast diet (0.007Ā Ī¼g Se/g diet) was only 14Ā % of the GPx1 mRNA expression in rats on a diet supplemented with 0.2Ā Ī¼g sodium selenite/g diet, which was comparable to the fall in liver mRNA [26].

Additionally, hair growth can be included as a biomarker because it has been reported to be reduced in beagles consuming a dietary Se concentration of 0.09Ā mg/kg DM (= approximately 5.5Ā Ī¼g/MJ) after 11Ā weeks [27]. Other indirect measures of Se status, such as serum creatine kinase (CK) and the thyroid hormones triiodothyronine (T3) and thyroxine (T4) were also included. Se is involved in the conversion of T4 into the active form T3 [28] and the T3:T4 ratio has been reported to decrease in puppies [29] and kittens [30] fed low compared to adequate Se concentrations. Finally, serum CK was increased in piglets fed either a diet containing no Se, although vitamin E levels were also manipulated in these studies [31].

Methods

Study design

A longitudinal, controlled and blinded study was performed using 24 adult Labrador retrievers. The dogs were assigned into 2 groups of 12 dogs, with age, gender and hair colour as blocking criteria. To ensure an adequate and stable selenium status in both groups at the start of the experimental period, the dogs received a Se adequate diet for 8Ā weeks before the experimental period. During this adaptation period, they were sampled three times (at weeks 2, 5 and 8) for baseline levels of blood and urine parameters and weekly hair growth measurements were performed. After the adaptation period, one group continued to receive the adequate Se diet and the other group received a diet with a low Se concentration for further 8Ā weeks. Weekly blood and urine samples were taken. In addition, weekly hair growth was measured. This study was approved by the WALTHAMĀ® Centre for Pet Nutrition Animal Welfare and Ethical Review Body and was conducted under UK Home Office Project Licence authorisation.

Dogs

Twenty-four adult Labrador retrievers (14 female and 10 male) were selected for this study. All Labradors used in this study were bred at WALTHAM or sourced from Home Office approved breeders for research purposes. The average age, body weight (BW) and energy intake in kJ/kg metabolic body weight (BW0.75) per dog group are shown in TableĀ 1. All dogs had an ideal body condition score (BCS) of D on the S.H.A.P.Eā„¢ BCS-scale [32]. Body weight and BCS was recorded before the start of the adaptation period and thereafter every week. The dogs were fed individually, once a day at 8:30Ā h, to energy requirements in order to maintain BW and BCS. They had access to fresh drinking water at all times. Food intake was also recorded throughout the study. Dogs were housed indoors in triplets with others of the same diet group with continuous access to an outside pen and access to an outside paddock during the day.

Table 1 Gendera, age, body weight and energy intake of the study dogs per dog group

Diets

Two semi-purified diets (Ssniff spezialdiƤten GmbH, Soest, Germany) were used, one Se-adequate and one containing a low amount of Se (TableĀ 2). Diets were formulated to be nutritionally complete for dogs with energy requirements of at least 397Ā kJ (95Ā kcal)/kg BW0.75. Both diets had the same nutritional formulation, except for Se concentration. The Se adequate diet contained 46.1Ā Ī¼g Se/MJ and the low Se diet 6.5Ā Ī¼g Se/MJ, which is 31Ā % of the FEDIAF recommended minimum for dogs with an energy intake of 95Ā kcal/kg BW0.75 [18]. Se concentration from the ingredients was approximately 0.75Ā Ī¼g/MJ. Supplementary Se was in the form of sodium selenite.

Table 2 Analyseda chemical composition of two semi-purified diets with an adequate or low selenium concentration

Blood samples

At weeks 2, 5 and 8 of the pre-feed period and weekly during the experimental period, blood samples (10.5Ā ml) were taken from each dog by jugular venipuncture using a 21 gauge needle. Blood was collected into a 2.5Ā ml PAXgene tube, 2 MicrovetteĀ® 500Ā Ī¼l lithium-heparin tubes, one 300Ā Ī¼l fluoride-heparin tube, one 200Ā Ī¼l Tri-Kalium-EDTA tube and two 4Ā ml VacuetteĀ® z serum clot activator tubes (one of them filled with only 2.5Ā ml of blood). The PAXgene tubes were incubated at room temperature for 2Ā h and stored at āˆ’20Ā Ā°C for mRNA expression analysis. One of the lithium-heparin tubes was immediately stored at āˆ’80Ā Ā°C and used for whole blood glutathione peroxidase (GPx) analysis at the end of the study. The other heparin tubes were centrifuged (accuSpinā„¢ Micro R, Fisher Scientificā„¢, Pittsburgh, PA, USA) immediately after collection at 1680ā€‰Ć—ā€‰g and 4Ā Ā°C for 10Ā min. Lithium-heparin plasma was analysed for general biochemistry parameters and fluoride-heparin plasma for glucose (Spectrophotometry, Olympus AU400). EDTA tubes were placed on a roller at room temperature until analysis for haematology parameters (OrphĆ©e Mythic 18 Vet analyser) within 4Ā h of collection. Serum tubes were incubated for 30Ā min on ice and then centrifuged (Sigma 6Ā K15, rotor 11150, cups 13550, Sigma GmbH, Osterode am Harz, Germany) for 10Ā min. at 2000ā€‰Ć—ā€‰g and 4Ā Ā°C. Serum samples were divided into Eppendorfs and stored at āˆ’80Ā Ā°C for analysis of GPx, Se, triiodothyronine (T3), thyroxine (T4), copper (Cu) and creatine kinase (CK) at the end of the study.

Urine samples

Free catch urine was collected weekly after feeding (between 8:30 and 16:00Ā h), using a Uripet urine collection device (Rocket Medical plc. Watford, England). From all urine samples, 1Ā ml was stored in an Eppendorf at āˆ’80Ā Ā°C and analysed for creatinine (CT) within one month after sampling (IDEXX laboratories, UK). The rest of the samples were stored at āˆ’20Ā Ā°C and analysed for total Se content (ICP-MS, PerkinElmer DRC-e, Waltham, USA).

Hair growth measurements

Hair growth was measured according to a method adapted from Yu et al. [27]. Due to practical issues only chocolate and black Labradors were used for the hair growth measurements (5 black, 1 chocolate for the adequate Se diet and 6 black, 1 chocolate for the low Se diet). At the start of the pre-feed and experimental period, an area in the groin of approximately 5 Ɨ 5Ā cm was shaved with a 40Ā mm blade (Andis Super AGR+ cordless clipper; Andis Company Corporate, Sturtevant, Wisconsin, USA). Directly after shaving, the area was marked with a Duramark permanent marker. The shaved area was covered with a glass slide with a ruler attached to it, and a picture was taken using a Canon EOS 1000D digital camera with EF50mm f/2.5 compact-macro lens (Canon Inc., Melville, New York, USA) and Hoya PRO1 Digital Polarising Filter (HOYA Optics, Milpitas, California, USA). The glass slide was also used to flatten the hairs, when it had grown back in the following weeks to accurately measure them. Pictures were taken weekly during the pre-feed and experimental periods and were analysed with ImageJ analysing software [33]. Hair growth (mm/week) was measured as the difference of the average length of the hairs within the marked box between two consecutive pictures.

Chemical analyses

Biochemistry and glucose analyses were carried out using spectrophotometry (Olympus AU400, Olympus Inc.) with Beckman Coulter reagents (Beckman Coulter Biomedical) within 20 mins of sampling. Haematology parameters were analysed using a Mythic 18 Vet analyser (OrphƩe S.A.). Serum and urine samples were prepared for total Se analyses with closed vessel microwave destruction as described in van Zelst et al. [34]. Se was analysed using inductively coupled plasma-MS (ICP-MS, Elan DRC-e, PerkinElmer), as described by Lavu et al. [35]. Urine CT was determined using a creatinine kit based on the Jaffe reaction (OSR6178, Beckman Coulter Biomedical, IDEXX Laboratories, London, UK).

Serum Cu and CK were analysed with the Randox copper and CK NAC-activated kit (Randox laboratories, London, UK), respectively, as per the manufacturerā€™s instructions. Thyroid hormones were analysed with canine T3 and T4 enzyme-linked immunosorbent assay (ELISA) kits (Cusabio, Wuhan, China). Whole blood GPx was analysed using the Ransel kit (Randox laboratories) and an Olympus AU400 spectrophotometer as per the manufacturerā€™s instructions. A 4-point calibration curve was used as control, with a 1:11, 1:21, 1:41 and 1:61 dilution. Whole blood and serum GPx analysis of the samples had an average coefficient of variation of 1.9 and 2.0Ā %, respectively. The 4-point calibration curve showed a recovery of 74.5, 87.1, 98.6, and 100.6Ā %, respectively.

mRNA expression was performed for the three baseline samples and the samples of weeks 2, 4, 6, and 8 of the experimental period. Total RNA, for mRNA expression, was isolated from the PAXgene tubes using the PAXgene blood RNA kit (Qiagen, Manchester, UK) as per the manufacturerā€™s instructions. The RNA was eluted in 2 Ɨ 40Ā Ī¼l of the elution buffer. Further DNase digestion of the RNA solution was carried out using RQ1 RNase-Free DNase (Promega, Southampton, UK) as per the manufacturerā€™s instructions with the sample incubated for 30Ā min at 37Ā Ā°C. In order to remove the DNase and reaction buffer from the purified RNA, it was passed through the RNeasy Mini Kit using the RNA clean-up protocol and was eluted in 2 Ɨ 40Ā Ī¼l of elution buffer (10Ā mM Tris HCl, pHĀ 8.4). The RNA concentration in the eluate was measured using the Qubit RNA Assay Kit (Invitrogen, Paisley, Scotland).

Primers and probes were designed using Primer3 [36] and M-Fold using the canine specific GenBank [37] sequences for GPx1, SelH, SepW1, TrxRd1, TrxRd2, DIO1, SelK, SelT, Sep15, tumor necrosis factor alpha (TNF-Ī±), and nuclear factor kappa-light-chain-enhancer of activated B cells (NFĪŗB) as described by Peters et al. [38]. The assays for the potential housekeeper genes were the same as described in Peters et al. [39].

Synthesis of complementary DNA (cDNA) was carried out with 750Ā ng of random hexamers using the ImProm-II Reverse Transcription System (Promega, Southampton, UK) and 750Ā ng of total RNA (as measured by the Qubit) in a final volume of 30Ā Ī¼l. All reactions were prepared according to the manufacturerā€™s instructions giving a final magnesium chloride concentration of 3Ā mM. The cDNA synthesis was carried out by mixing the RNA with the random primers in a reaction tube. Samples were heated to 70Ā Ā°C for 5Ā min in the PTC-200 DNA engine (Bio-Rad Laboratories, Hemel Hempstead, UK) before cooling to 4Ā Ā°C for 5Ā min. Tubes were placed in a cold block before addition of the reaction buffer, deoxynucleotides (dNTPā€™s), magnesium chloride, reverse transcriptase enzyme mix and water to make a total volume of 20Ā Ī¼l. Reverse transcription (RT) was undertaken by heating the samples to 25Ā Ā°C for 5Ā min, 42Ā Ā°C for 30Ā min and finally 75Ā Ā°C for 10Ā min in the PTC-200 DNA engine (Bio-Rad Laboratories, Hemel Hempstead, UK). Duplicate RT reactions were performed for each RNA sample. All cDNAs were diluted 1:5 (v/v) for the genes of interest and housekeeper genes using EB Buffer (10Ā mM Trisā€“HCl pH-8.4, Qiagen, Manchester, UK) and then stored at āˆ’20Ā Ā°C for future use. No template controls were performed by addition of nuclease-free water in place of RNA.

Quantitative polymerase chain reaction (qPCR) was performed using GoTaq Colourless Master Mix (Promega, Southampton, UK) as described by Van de Velde et al. [40]. The absence of genomic contamination of the RNA samples was confirmed prior to the RT reactions and none of the samples showed evidence of amplifiable genomic DNA with the succinate dehydrogenase complex, subunit A qPCR assay. One qPCR reaction was run for each RT repeat resulting in two threshold values (Ct) for each RNA sample. A mean Ct value was calculated for each sample using the two measured Ct values for each dog for each of the potential housekeeper genes. The mean Ct value was converted to a relative copy number value using the EĪ”Ct method (E: reaction efficiency as determined from a standard curve) with Ī”Ct values calculated relative to the sample with the largest Ct (fewest gene copies). The geNorm visual basics for applications applet for Microsoft Excel was used to determine the most stable genes from the set of tested genes [41]. The three most stable housekeeper genes for the blood samples were beta-2-microglobulin (B2M), glyceraldehyde 3-phosphate dehydrogenase (GAPDH), and hypoxanthine phosphoribosyltransferase 1 (HPRT1). The primer and probe sequences of the genes of interest and these housekeeper genes are shown in TableĀ 3.

Table 3 Primer and probe sequences used for positive control and qPCR assays

Relative copy number expression values were calculated for each sample and normalised against the housekeeper gene results using the qBase applet for Microsoft Excel which employs the methodology described by Vandesompele et al. [41]. The sample with the fewest gene copies (latest Ct value) is given a relative copy number of 1 and all other samples are given values relative to this sample.

In order to assess reaction efficiency of the newly designed assays, a set of primers were designed for the gene target to amplify a larger fragment, which included the portion amplified by the qPCR assay. These assays were tested against a cDNA obtained from RNA extracted from canine blood. Products were separated by 2Ā % agarose gel electrophoresis, purified by NucleoSpin Extract II kit (Macherey-Nagel, DĆ¼ren, Germany) and then quantified by QBit (Invitrogen, Paisley, Scotland). The number of copies per Ī¼l of purified product was calculated and then a 1:10 dilution series from 107 to 1 copy per qPCR was analysed in duplicate using the qPCR assay and the reaction efficiency calculated using the MxPro software.

Statistical analyses

The two primary parameters were whole blood GPx and urinary Se:CT ratio. The secondary parameters in this study were: Se intake, serum Se, serum GPx, serum CK, serum T3:T4 ratio, serum Cu relative to Cu intake, hair growth, and mRNA expression for GPx1, SelH, SepW1, TrxRd1, TrxRd2, DIO1, SelK, SelT, Sep15, TNF-Ī±, and NFĪŗB.

The average of the three baseline measurements were analysed between dog groups with two-sample Studentā€™s t-tests. Using the entire dataset,each parameter was analysed univariately with a linear mixed effects model in R version 3.1.1 using the nlme package (R Foundation for Statistical Computing, Vienna, Austria). The fixed effects were diet, week, diet x week interaction, and baseline of the parameter (the average of the dogsā€™ three pre-feed measurements). The random effect was dog. The assumption of normality was assessed for each model by visual inspection of the residuals and urinary Se:CT, serum Cu and all mRNA parameters were log transformed. Between diet contrasts were applied in each week and between week contrasts were applied within each diet. Contrasts on the two primary parameters were Bonferroni corrected to an overall significance level of 5Ā % (i.e. threshold p-value of 0.025).

Results

Se intake during the pre-feed period was similar in the two experimental groups (Group A: average 22.9Ā Ī¼g/kg BW0.75, range 19.0ā€“29.5Ā Ī¼g/kg BW0.75, Group B: average 20.6Ā Ī¼g/kg BW0.75, range 16.8ā€“25.2Ā Ī¼g/kg BW0.75). During the experimental period, average Se intake for the dogs on the adequate Se diet was 21.7Ā Ī¼g/kg BW0.75 (range 16.6ā€“28.2Ā Ī¼g/kg BW0.75) and for the dogs on the low Se diet was 2.7Ā Ī¼g/kg BW0.75 (range 2.2ā€“3.4Ā Ī¼g/kg BW0.75) which was significantly different comparing each week (pā€‰<ā€‰0.001). There were no significant differences between the groups for any of the parameters at baseline, except for the serum Cu:Cu intake ratio (pā€‰=ā€‰0.033).

Urinary Se:CT ratio significantly decreased in dogs on the low Se diet from week 1 onwards and on average decreased by 84Ā % (Fig.Ā 1). Whole blood GPx activity was significantly lower (by 7Ā %) in dogs fed the low Se diet at week 8 (Fig.Ā 2). Cumulative hair growth results are shown in Fig.Ā 3 and no differences between the diets were found. An overview of the results of all serum measurements is given in TableĀ 4. Serum Se significantly decreased by 7Ā % in dogs fed the low compared to the adequate Se diet from week 1 onwards. A significant difference in serum GPx was detected from week 6 (6Ā % mean decrease from first significance).

Fig. 1
figure 1

Urinary selenium to creatinine ratio of dogs fed a low or adequate selenium diet. The low selenium diet contained 6.5Ā Ī¼g/MJ and the adequate selenium diet 46.1Ā Ī¼g/MJ. Black triangles are dogs on the adequate selenium diet, open squares are dogs on the low selenium diet. Values at week zero indicate the average baseline values. Symbols represent the means and error bars indicate their standard errors, based on the raw data

Fig. 2
figure 2

Glutathione peroxidase activity (U/L) in whole blood of dogs fed a low or adequate selenium diet. The low selenium diet contained 6.5Ā Ī¼g/MJ and the adequate selenium diet 46.1Ā Ī¼g/MJ. Black triangles are dogs on the adequate selenium diet, open squares are dogs on the low selenium diet. Values at week zero indicate the average baseline values. Symbols represent the means and error bars indicate their standard errors, based on the raw data

Fig. 3
figure 3

Cumulative hair growth (mm) of dogs fed a low or adequate selenium diet. Hair growth was measured weekly in the groin of Labrador retrievers fed a diet containing a low (6.5Ā Ī¼g/MJ) or adequate (46.1Ā Ī¼g/MJ) concentration of selenium. Black triangles are dogs on the adequate selenium diet, open squares are dogs on the low selenium diet. Values at week zero indicate the average baseline values at weeks 2, 5 and 8 of the pre-feed period, where hair had been removed at week 0. Symbols represent the means and error bars indicate their standard errors, based on the raw data

Table 4 Serum biomarker concentrations in dogs fed a diet with a low or adequate selenium concentration

None of the mRNA measures were significantly changed by the consumption of the low Se diet (see TableĀ 5). Interestingly, there was an upregulation in both groups in mRNA expression of SelK, SelT, and Sep15 during the second half of the experimental period (weeks 2 & 4 vs. weeks 6 & 8) and a downregulation in the expression of SelH, TrxRd1, TrxRd2, TNF-Ī±, and NFĪŗB (pā€‰<ā€‰0.05).

Table 5 Relativea mRNA expression in dogs fed a diet with a low or adequate selenium concentration

Discussion

This study indicated that there was a potential difference in the reaction of biomarkers to changes in dietary Se concentration in dogs with an adequate Se status. Some of the measured biomarkers reacted within one week, whilst others did not significantly change within the 8Ā week study period. The estimated biomarker reaction time will be dependent on the power for each biomarker and so, more biomarkers may have been identified as significantly reacting if more dogs were included in the study. The number of dogs in this study (nā€‰=ā€‰24) was selected to ensure 80Ā % power to detect differences between the diets in the primary measures: urinary Se:CT and whole blood GPx. This is considered a realistic number to be repeated at other dog facilities, as larger numbers of dogs are rarely found in research facilities. This study did not evaluate the effects of Se on dog health, but does give an indication on which parameters may be used to assess Se associated long term health effects.

Urinary Se:CT ratio dramatically decreased within one week after changing to the low Se diet. Urine is the main route of excretion of excess Se [42, 43]. 24-h urinary Se concentrations have previously been demonstrated to correlate well with Se intake in humans [43, 44], cats [15, 20], and dogs [15] and this study demonstrated that the more practical single void sample, corrected for CT concentration, is a valuable indicator of dietary Se concentration in dogs. Urinary Se:CT ratio has previously been validated as a proxy measure of 24-h urinary Se excretion in humans [45] but has never been measured in diseased dogs (e.g. with cancer), so there is no dog specific estimate of a safe urinary Se:CT concentration. However, in humans there are indications that urinary Se reflects Se requirements, as excretion decreases in healthy children, pregnant women and in people with cancer, where it is likely that Se requirements are higher compared to healthy individuals [43]. This makes urinary Se:CT ratio an important biomarker to measure changes in dietary Se intake, but the safe minimum level of urinary Se:CT ratio that is associated with the prevention of long term health effects warrants further research.

Serum Se concentrations also decreased within one week after switching the dogs to the low Se diet. The average serum Se concentrations of the dogs in this study was 257Ā Ī¼g/L (range 213ā€“310Ā Ī¼g/L) in the adequate and 238Ā Ī¼g/L (range 172ā€“298Ā Ī¼g/L) in the low Se group, and fell within the range reported by Forrer et al. [46] (150ā€“340Ā Ī¼g/L) for healthy dogs. However, dogs in the study of Forrer et al. were healthy at the time of sampling, but were not monitored during a longer period, so disease may have developed at a later stage. Pilarczyk et al. [47] found a significantly lower (pā€‰<ā€‰0.01) serum Se concentration in dogs with malignant neoplasm (range 103ā€“265Ā Ī¼g/L) compared to healthy dogs (range 208ā€“346Ā Ī¼g/L). This association reinforces the need to verify whether this relationship is causal.

GPx measurements were taken as an indicator of Se bioactivity. Se requirements are often based on GPx concentrations reaching a plateau [29, 48]. However, we do not know if optimal selenoprotein concentrations are desirable for optimal health. In humans, a blood Se concentration of 79ā€“95Ā Ī¼g/L is considered sufficient to maximize GPx activity [48]. This seems to be higher in dogs, as the average serum Se concentration of dogs in the low Se group in the last week of the study was 224Ā Ī¼g/L, where both serum and whole blood GPx concentrations were significantly lower than in the adequate Se group, and thus not maximised.

In this study, inorganic Se (sodium selenite) was used as the primary source of Se. This can be converted to selenocysteine and incorporated into selenoproteins [49]. Selenomethionine cannot be synthesized from inorganic selenium, but may be non-specifically incorporated into body proteins [50], so using inorganic Se prevents selenomethionine mixing with the methionine pool and incorporation of selenomethionine into body proteins where it has no Se specific role [49]. The use of selenomethionine (organically-bound Se) would have likely resulted in a higher retention of Se in the body (i.e. lower excretion) without being bioactive, and a higher dietary Se requirement may thus apply for organic compared to inorganic Se species in order to maximize selenoprotein activity.

The body stores of Se may also have had an impact on all biomarkers measured in this study. In the first part of the study, Se may still have been present in body proteins and during protein turn-over may have been released to become available for selenoprotein incorporation (in this study measured as GPx), while in the last few weeks body stores of Se may have become depleted. It is likely that the amount of stored Se is animal species specific, as Todd et al. [15] found that cats retain less Se than dogs. This may be linked to the evolutionary ā€œfeast and famineā€ feeding approach of dogs, as described by Bosch et al. [51], in which dogs gorge feed, and during the famine stage, stored Se may be used for physiological processes.

Serum GPx, in this study, significantly reacted two weeks earlier to a change in dietary Se compared to whole blood GPx. Similar findings were reported in cats by Todd et al. [20]. Changes in plasma GPx were observed after feeding a diet containing 2, 1.43 and 0.98Ā mg sodium selenite/kg DM after 16, 32 and 24Ā days, respectively, compared to a basal diet containing 0.45Ā mg Se/kg DM. No changes in whole blood GPx were observed within the 32-day lasting study of Todd et al. [20]. This difference may simply be caused by the higher level and variation of GPx activity in whole blood compared to serum. GPx activity of dogs in whole blood is approximately five times higher than in serum. Another possible reasoning is that the surplus of tissue GPx is effluxed into the plasma, as indicated by the high correlation between tissue and plasma GPx concentrations [52]. The difference in reaction time between serum and whole blood GPx activity may also be explained by different isoforms of GPx. Plasma is known to contain mainly the plasma GPx (GPx3) enzyme [53], while whole blood may additionally contain other forms of GPx. However, the GPx activity measurement does not discriminate between various forms of GPx and it may be that the activity of serum GPx is more quickly affected by a reduction in dietary Se than other forms of GPx. In rats, it has been determined that plasma GPx3 activity is more highly regulated by dietary Se concentration than, for example, muscle GPx4 [54]. The same authors [54] have shown that although GPx1 mRNA expression was highly regulated by dietary Se concentration in rats, dietary Se requirements based on tissue (plasma, red blood cell, liver, kidney, and muscle) GPx activity were higher than when based on mRNA expression (i.e. mRNA expression plateaus at lower dietary Se concentrations). This is in accordance with the findings of the present study. The fact that GPx is affected, but the selenoprotein mRNA levels are not, indicates that the bioactive Se fraction of the low Se diet (total Se concentration: 6.5Ā Ī¼g/MJ or 0.11Ā mg/kg DM) is sufficient to maintain mRNA expression of the measured selenoproteins. Although mRNA expression is the basis of protein formation, the actual formation of the active selenoproteins were not maximized by the low Se diet. This may be due to effects on translation, like the structure of specific stem-loops which bind proteins [55]. Also, post-translational modifications (which affect the actual activity of the selenoproteins, e.g. protein folding, phosphorylation, etc.) may have played a role in the difference in reaction of GPx1 mRNA expression and GPx activity to a change in dietary Se concentration. Therefore, GPx activity may be a more reliable biomarker for Se status than the mRNA expression of selenoproteins. However, it may be that maximal selenoprotein activity is not necessary for the prevention of disease on the long term. Selenius et al. [56], for example, found that TrxRd1 mRNA expression was increased with increasing addition of sodium selenite concentrations (range 2.5ā€“10Ā Ī¼M) to lung cancer cells, while TrxRd activity decreased at high selenite concentrations, which indicates the impairment of selenoprotein formation. Therefore, it is hypothesized that selenoprotein mRNA expression may be a useful biomarker in the detection of high dietary Se concentrations, but that GPx activity is more useful in the estimation of minimum Se requirements.

The fact that the SelK, SelT, and Sep15 mRNA markers were increased and SelH, TrxRd1, TrxRd2, TNF-Ī±, and NFĪŗB were decreased in the second half compared to the first half of the study, point towards an event of inflammation during the first part of the study, as TNF-Ī±, and NFĪŗB are inflammation markers. The study did not include induction of inflammation and there are no indications of such events during or before the study, all circumstances stayed the same. This reinforces the idea that other factors than dietary Se have more impact on the selenoprotein mRNA expression than dietary Se concentration.

The dietary Se concentration did not result in a significant difference in hair growth between the diets. Previous studies which have used hair growth as a marker of Se status [27] used beagles, so there may be a breed variation in hair growth. Although no literature on speed of hair growth between different breeds could be found, there is evidence for genetic differences in coat formation between breeds [57]. However, the most likely explanation for the lack of difference in hair growth is that the current study has used sodium selenite as Se source and in the study of Yu et al. [27], selenomethionine was supplemented. Methionine is often a limiting amino acid and selenomethionine can be non-specifically incorporated into body proteins such as hair [50]. Therefore, it may be that the decrease in hair growth in the study of Yu et al. is not due to a reduction in Se, but in methionine.

As there is no clear adequate range of dietary Se for adult dogs, the low Se diet may have contained sufficient Se (6.5Ā Ī¼g/MJ or 0.11Ā mg/kg DM) to maintain health. In this 8Ā week study it did not have a negative effect on thyroid hormone metabolism (T3:T4). The conversion of T4 into the active form T3 has been reported to increase with increasing dietary Se concentrations in puppies within a range of 0ā€“0.52Ā mg/kg [29]. Also Cu is indicated to be involved in thyroid hormone metabolism, as a lower GPx activity and liver selenodeiodinase activity was found in Cu deficient rats compared to a control group [58]. These results indicate that the low Se diet is not deficient for thyroid hormone metabolism of adult dogs. Although the serum Cu:Cu intake ratio was significantly different between the groups at baseline, this is unlikely to have affected the results as including the baseline intake as a covariate in the analysis will normalize any differences between groups. Further studies are required to determine the long term health effects associated with a change in urinary Se:CT ratio, serum Se and serum GPx concentrations. This can be done by a retrospective study to determine adequate ranges of these markers and to establish which set of markers from our study is needed to provide a sufficient prediction of the risk of Se-induced disease. Consecutively, these ranges can be used in a study to determine the minimal Se requirement of dogs.

Conclusion

There were variations in the reaction time of Se biomarkers to a reduction in dietary Se concentration in dogs. This study has demonstrated that urinary Se:CT ratio, serum Se, and serum GPx activity react most quickly. These may be useful biomarkers in future long term studies to evaluate the minimum requirements for optimal health in dogs.

Abbreviations

B2M:

beta-2-microglobulin

BCS:

Body condition score

BW:

Body weight

BW0.75 :

Metabolic body weight

cDNA:

Complementary deoxyribonucleic acid

CK:

Creatine kinase

Ct:

Threshold value

CT:

Creatinine

Cu:

Copper

DIO:

Iodothyronine deiodinase

DM:

Dry matter

ELISA:

Enzyme-linked immunosorbent assay

GPx:

Glutathione peroxidase

GAPDH:

Glyceraldehyde 3-phosphate dehydrogenase

HPRT1:

Hypoxanthine phosphoribosyltransferase 1

MJ:

Mega joule

mRNA:

Messenger ribonucleic acid

NFĪŗB:

Nuclear factor kappa-light-chain-enhancer of activated B cells

qPCR:

Quantitative polymerase chain reaction

RT:

Reverse transcription

Se:

Selenium

SeCys:

Selenocysteine

SelH:

Selenoprotein H

SelK:

Selenoprotein K

SelT:

Selenoprotein T

SeMet:

Selenomethionine

Sep15:

15Ā kDa selenoprotein

SepW1:

Selenoprotein W

T3:

Triiodothyronine

T4:

Thyroxine

TNF-Ī±:

Tumor necrosis factor alpha

TrxRd:

Thioredoxin reductase

References

  1. National Research Council. Nutrient requirements of dogs and cats, Rev. ed. Animal nutrition series. Washington D.C: National Academies Press; 2006.

    Google ScholarĀ 

  2. European Food and Safety Authority. Scientific opinion on the substantiation of health claims related to selenium and protection of DNA, proteins and lipids from oxidative damage, function of the immune system, thyroid function, function of the heart and blood vessels, prostate function, cognitive function and spermatogenesis pursuant to article 13 of Regulation (EC) No 1924/2006. EFSA Journal. 2009;7(9):1ā€“24. doi:10.2903/j.efsa.2009.1220.

    Google ScholarĀ 

  3. Duntas LH. Selenium and the thyroid: a close-knit connection. J Clin Endocrinol Metab. 2010;95(12):5180ā€“8. doi:10.1210/jc.2010-0191.

    ArticleĀ  PubMedĀ  CASĀ  Google ScholarĀ 

  4. McKenzie RC, Arthur JR, Miller SM, Rafferty TS, Beckett GJ. Selenium and the immune system. In: Calder PC, Field CJ, Gill HS, editors. Nutrition and immune function. Wallingford: CABI Publishing; 2002. p. 229ā€“50.

    ChapterĀ  Google ScholarĀ 

  5. Kumar MS, Selvam R. Supplementation of vitamin E and selenium prevents hyperoxaluria in experimental urolithic rats. J Nutr Biochem. 2003;14(6):306ā€“13. http://www.jnutbio.com/article/S0955-2863%2803%2900033-0/pdf.

    ArticleĀ  Google ScholarĀ 

  6. Liu Y, Xu H, Zhong W, Shen Q, Zhuang T, Huang K. Organic Selenium Alleviated the Formation of Ethylene Glycol-Induced Calcium Oxalate Renal Calculi by Improving Osteopontin Expression and Antioxidant Capability in Dogs. Biol Trace Elem Res. 2015:1ā€“9. doi:10.1007/s12011-015-0373-9.

  7. Fico ME, Poirier KA, Watrach AM, Watrach MA, Milner JA. Differential effects of selenium on normal and neoplastic canine mammary cells. Cancer Res. 1986;46(7):3384ā€“8.

    PubMedĀ  CASĀ  Google ScholarĀ 

  8. Waters DJ, Shen S, Glickman LT, Cooley DM, Bostwick DG, Qian J, et al. Prostate cancer risk and DNA damage: translational significance of selenium supplementation in a canine model. Carcinogenesis. 2005;26(7):1256ā€“62. doi:10.1093/carcin/bgi077.

    ArticleĀ  PubMedĀ  CASĀ  Google ScholarĀ 

  9. Aldrich G. Rendered products in pet food. In: Meekers DL, editor. Essential rendering. All about the animal by-product industry. Virginia: National Renderers Association; 2006. p. 159ā€“77.

    Google ScholarĀ 

  10. Zentek J. A changing landscape: the pet food market in Europe. 2007. http://en.engormix.com/MA-feed-machinery/formulation/articles/changing-landscape-pet-food-t423/p0.htm. Accessed 13 July 2015.

    Google ScholarĀ 

  11. Thiry C, Ruttens A, de Temmerman L, Schneider Y-J, Pussemier L. Current knowledge in species-related bioavailability of selenium in food. Food Chem. 2012;130(4):767ā€“84. doi:10.1016/j.foodchem.2011.07.102.

    ArticleĀ  CASĀ  Google ScholarĀ 

  12. Gibson RS. Assessment of trace element status in humans. Prog Food Nutr Sci. 1989;13(2):67ā€“111.

    PubMedĀ  CASĀ  Google ScholarĀ 

  13. Alfthan G, Aro A, Arvilommi H, Huttunen JK. Selenium metabolism and platelet glutathione peroxidase activity in healthy Finnish men: effects of selenium yeast, selenite, and selenate. Am J Clin Nutr. 1991;53(1):120ā€“5.

    PubMedĀ  CASĀ  Google ScholarĀ 

  14. Wedekind KJ, Cowell C, Combs Jr GF. Bioavailability of selenium in petfood ingredients. FASEB J. 1997;11:A360.

    Google ScholarĀ 

  15. Todd SE, Thomas DG, Bosch G, Hendriks WH. Selenium status in adult cats and dogs fed high levels of dietary inorganic and organic selenium. J Anim Sci. 2012;90(8):2549ā€“55. doi:10.2527/jas.2011-3911.

    ArticleĀ  PubMedĀ  CASĀ  Google ScholarĀ 

  16. Roman M, Jitaru P, Barbante C. Selenium biochemistry and its role for human health. Metallomics. 2014;6:25ā€“54.

    ArticleĀ  PubMedĀ  CASĀ  Google ScholarĀ 

  17. Wedekind KJ, Bever RS, Combs Jr GF. Is Selenium addition necessary in pet foods? FASEB J. 1998;12:A823.

    Google ScholarĀ 

  18. FEDIAF. Nutritional guidelines for complete and complementary pet food for cats and dogs. Bruxelles: European Pet Food Industry Federation; 2014.

    Google ScholarĀ 

  19. Levander OA, Sutherland B, Morris VC, King JC. Selenium balance in young men during selenium depletion and repletion. Am J Clin Nutr. 1981;34(12):2662ā€“9.

    PubMedĀ  CASĀ  Google ScholarĀ 

  20. Todd SE, Thomas DG, Hendriks WH. Selenium balance in the adult cat in relation to intake of dietary sodium selenite and organically bound selenium. J Anim Physiol Anim Nutr. 2011;96:148ā€“58. doi:10.1111/j.1439-0396.2011.01132.x.

    ArticleĀ  Google ScholarĀ 

  21. Sunde RA, Raines AM, Barnes KM, Evenson JK. Selenium status highly-regulates selenoprotein mRNA levels for only a subset of the selenoproteins in the selenoproteome. Biosci Rep. 2009;29(5):329ā€“38. doi:10.1042/bsr20080146.

    ArticleĀ  PubMedĀ  CASĀ  PubMed CentralĀ  Google ScholarĀ 

  22. Sunde RA. Molecular biomarker panels for assessment of selenium status in rats. Exp Biol Med. 2010;235(9):1046ā€“52. doi:10.1258/ebm.2010.010111.

    ArticleĀ  CASĀ  Google ScholarĀ 

  23. Read R, Bellew T, Yang JG, Hill KE, Palmer IS, Burk RF. Selenium and amino acid composition of selenoprotein P, the major selenoprotein in rat serum. J Biol Chem. 1990;265(29):17899ā€“905.

    PubMedĀ  CASĀ  Google ScholarĀ 

  24. Hill KE, Zhou J, McMahan WJ, Motley AK, Atkins JF, Gesteland RF, et al. Deletion of selenoprotein P alters distribution of selenium in the mouse. J Biol Chem. 2003;278(16):13640ā€“6. doi:10.1074/jbc.M300755200.

    ArticleĀ  PubMedĀ  CASĀ  Google ScholarĀ 

  25. Sunde RA. Selenoproteins: hierarchy, requirements, and biomarkers. In: Hatfield DL, Berry MJ, Gladyshev VN, editors. Selenium. Itā€™s molecular biology and role in human health. 3rd ed. Dordrecht: Springer Science and Business Media; 2012. p. 137ā€“52.

    Google ScholarĀ 

  26. Evenson JK, Wheeler AD, Blake SM, Sunde RA. Selenoprotein mRNA is expressed in blood at levels comparable to major tissues in rats. J Nutr. 2004;134(10):2640ā€“5.

    PubMedĀ  CASĀ  Google ScholarĀ 

  27. Yu S, Wedekind KJ, Kirk CA, Nachreiner RF. Primary hair growth in dogs depends on dietary selenium concentrations. J Anim Physiol Anim Nutr. 2006;90(3ā€“4):146ā€“51. doi:10.1111/j.1439-0396.2005.00575.x.

    ArticleĀ  CASĀ  Google ScholarĀ 

  28. Kralik A, Eder K, Kirchgessner M. Influence of zinc and selenium deficiency on parameters relating to thyroid hormone metabolism. Horm Metab Res. 1996;28(5):223ā€“6.

    ArticleĀ  PubMedĀ  CASĀ  Google ScholarĀ 

  29. Wedekind KJ, Yu S, Combs Jr GF. The selenium requirement of the puppy. J Anim Physiol Anim Nutr. 2004;88(9ā€“10):340ā€“7. doi:10.1111/j.1439-0396.2004.00489.x.

    ArticleĀ  CASĀ  Google ScholarĀ 

  30. Yu S, Howard KA, Wedekind KJ, Morris JG, Rogers QR. A low-selenium diet increases thyroxine and decreases 3,5,3ā€™triiodothyronine in the plasma of kittens. J Anim Physiol Anim Nutr. 2002;86:36ā€“41.

    ArticleĀ  CASĀ  Google ScholarĀ 

  31. Fontaine M, Valli VE, Young LG, Lumsden JH. Studies on vitamin E and selenium deficiency in young pigs. I. Hematological and biochemical changes. Can J Comp Med. 1977;41(1):41ā€“51.

    PubMedĀ  CASĀ  PubMed CentralĀ  Google ScholarĀ 

  32. German AJ, Holden SL, Moxham GL, Holmes KL, Hackett RM, Rawlings JM. A simple, reliable tool for owners to assess the body condition of their dog or cat. J Nutr. 2006;136:2031Sā€“3S.

    PubMedĀ  CASĀ  Google ScholarĀ 

  33. AbrĆ moff MD, Magalhaes PJ, Ram SJ. Image processing with ImageJ. Biophotonics Intern. 2004;11(7):36ā€“42.

    Google ScholarĀ 

  34. van Zelst M, Hesta M, Alexander LG, Gray K, Bosch G, Hendriks WH, et al. In vitro selenium accessibility in dog foods is affected by diet composition and type. Br J Nutr. 2015;113:1888ā€“94.

    ArticleĀ  PubMedĀ  Google ScholarĀ 

  35. Lavu RVS, Willekens K, Vandecasteele B, Tack F, Du Laing G. Fertilizing soil with Selenium fertilizers: impact on concentration, speciation and bioaccessibility of Selenium in leek (Allium ampeloprasum). J Agric Food Chem. 2012;60:10930ā€“5.

    ArticleĀ  PubMedĀ  CASĀ  Google ScholarĀ 

  36. Rozen S, Skaletsky H. Primer3 on the WWW for general users and for biologist programmers. Methods Mol Biol. 2000;132:365ā€“86.

    PubMedĀ  CASĀ  Google ScholarĀ 

  37. GenBank. Genetic Sequence Data Bank. National Center for Biotechnology Information. 2015. http://www.ncbi.nlm.nih.gov/nuccore/. Accessed 6 July 2015.

  38. Peters I, Helps CR, Hall EJ, Day MJ. Realt-time RT-PCR: considerations for efficient and sensitive assay design. J Immunol Methods. 2004;286:203ā€“17.

    ArticleĀ  PubMedĀ  CASĀ  Google ScholarĀ 

  39. Peters I, Peeters D, Helps CR, Day MJ. Development and application of multiple internal reference (housekeeper) gene assays for accurate normalisation of canine gene expression studies. Vet Immunol Immunopathol. 2007;117(1ā€“2):55ā€“66.

    ArticleĀ  PubMedĀ  CASĀ  Google ScholarĀ 

  40. Van de Velde H, Janssens GPJ, de Rooster H, Polis I, Peters I, Ducatelle R, et al. The cat as a model for human obesity: insights into depot-specific inflammation associated with feline obesity. Br J Nutr. 2013;110(07):1326ā€“35.

    ArticleĀ  PubMedĀ  Google ScholarĀ 

  41. Vandesompele J, De Preter K, Pattyn F, Poppe B, Van Roy N, De Paepe A, et al. Accurate normalization of real-time quantitative RT-PCR data by geometric averaging of multiple internal control genes. Genome Biol. 2002;3(7):research 0034.

    ArticleĀ  Google ScholarĀ 

  42. Francesconi KA, Pannier F. Selenium metabolites in urine: a critical overview of past work and current status. Clin Chem. 2004;50(12):2240ā€“53. doi:10.1373/clinchem.2004.039875.

    ArticleĀ  PubMedĀ  CASĀ  Google ScholarĀ 

  43. Sanz Alaejos M, DĆ­az RC. Urinary selenium concentrations. Clin Chem. 1993;39(10):2040ā€“52.

    PubMedĀ  CASĀ  Google ScholarĀ 

  44. Hawkes W, Richter BD, Alkan Z, Souza E, Derricote M, Mackey B, et al. Response of selenium status indicators to supplementation of healthy North American Men with High-Selenium Yeast. Biol Trace Elem Res. 2008;122(2):107ā€“21. doi:10.1007/s12011-007-8066-7.

    ArticleĀ  PubMedĀ  CASĀ  Google ScholarĀ 

  45. Hojo Y. Evaluation of the expression of urinary selenium level as ng Se/mg creatinine and the use of single-void urine as a sample for urinary selenium determination. Bull Environ Contam Toxicol. 1981;27:213ā€“20.

    ArticleĀ  PubMedĀ  CASĀ  Google ScholarĀ 

  46. Forrer R, Gautschi K, Lutz H. Comparative determination of selenium in the serum of various animal species and humans by means of electrothermal atomic absorption spectrometry. J Trace Elem Electrolytes Health Dis. 1991;5(2):101ā€“13.

    PubMedĀ  CASĀ  Google ScholarĀ 

  47. Pilarczyk B, Tomza-Marciniak A, Pilarczyk R, Bąkowska M, Gaik M, Wilk M, et al. Relationship between serum Se concentration in dogs and incidence of some disease conditions. Cent Eur J Biol. 2013;8(6):527ā€“33. doi:10.2478/s11535-013-0163-8.

    CASĀ  Google ScholarĀ 

  48. Thomson CD. Assessment of requirements for selenium and adequacy of selenium status: a review. Eur J Clin Nutr. 2004;58(3):391ā€“402.

    ArticleĀ  PubMedĀ  CASĀ  Google ScholarĀ 

  49. Suzuki KT. Metabolomics of selenium: Se metabolites based on speciation studies. J Health Sci. 2005;51(2):107ā€“14. doi:10.1248/jhs.51.107.

    ArticleĀ  CASĀ  Google ScholarĀ 

  50. Schrauzer GN. Selenomethionine: a review of its nutritional significance, metabolism and toxicity. J Nutr. 2000;130:1653ā€“6.

    PubMedĀ  CASĀ  Google ScholarĀ 

  51. Bosch G, Hagen-Plantinga EA, Hendriks WH. Dietary nutrient profiles of wild wolves: insights for optimal dog nutrition? Br J Nutr. 2015;113(SupplementS1):S40ā€“54. doi:10.1017/S0007114514002311.

    ArticleĀ  PubMedĀ  CASĀ  Google ScholarĀ 

  52. Combs Jr GF, Combs SB. Biochemical functions of selenium. The role of selenium in nutrition. New York: Academic; 1986. p. 205ā€“63.

    BookĀ  Google ScholarĀ 

  53. Avissar N, Ornt DB, Yagil Y, Horowitz S, Watkins RH, Kerl EA, et al. Human kidney proximal tubules are the main source of plasma glutathione peroxidase. Am J Physiol. 1994;266(2):C367ā€“C75.

    PubMedĀ  CASĀ  Google ScholarĀ 

  54. Barnes KM, Evenson JK, Raines AM, Sunde RA. Transcript analysis of the selenoproteome indicates that dietary selenium requirements of rats based on selenium-regulated selenoprotein mRNA levels are uniformly less than those based on glutathione peroxidase activity. J Nutr. 2009;139(2):199ā€“206.

    ArticleĀ  PubMedĀ  CASĀ  PubMed CentralĀ  Google ScholarĀ 

  55. Hesketh JE, Vasconcelos MH, Bermano G. Regulatory signals in messenger RNA: determinants of nutrientā€“gene interaction and metabolic compartmentation. Br J Nutr. 1998;80(04):307ā€“21. doi:10.1017/S0007114598001378.

    ArticleĀ  PubMedĀ  CASĀ  Google ScholarĀ 

  56. Selenius M, Fernandes AP, Brodin O, Bjƶrnstedt M, Rundlƶf A-K. Treatment of lung cancer cells with cytotoxic levels of sodium selenite: Effects on the thioredoxin system. Biochem Pharmacol. 2008;75(11):2092ā€“9. http://www.sciencedirect.com/science/article/pii/S0006295208001391.

    ArticleĀ  PubMedĀ  CASĀ  Google ScholarĀ 

  57. Cadieu E, Neff MW, Quignon P, Walsh K, Chase K, Parker HG, et al. Coat variation in the domestic dog is governed by variants in three genes. Science. 2009;326(5949):150ā€“3. doi:10.1126/science.1177808.

    ArticleĀ  PubMedĀ  CASĀ  PubMed CentralĀ  Google ScholarĀ 

  58. Olin KL, Walter RM, Keen CL. Copper deficiency affects selenoglutathione peroxidase and selenodeiodinase activities and antioxidant defense in weanling rats. Am J Clin Nutr. 1994;59(3):654ā€“8.

    PubMedĀ  CASĀ  Google ScholarĀ 

Download references

Acknowledgements

The hair growth measurements were done by Robyn Bednall and Tim Oā€™Brine of the WALTHAMĀ® Centre for Pet Nutrition, which was very much appreciated. The authors also express their thanks to Joachim Neri of the Department of Applied Analytical and Physical Chemistry at the Faculty of Bioscience Engineering of Ghent University for the Se analyses. Donna Vanhauteghem of the Laboratory of Animal Nutrition at Ghent University is thanked for the Cu, CK, T3 and T4 analyses and Iain Peters of the Innovation Centre (Exeter, UK) is gratefully acknowledged for his expertise and analysis of the mRNA samples.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Geert P. J. Janssens.

Additional information

Competing interests

This study is part of a PhD project of MvZ funded by the WALTHAM Centre for Pet Nutrition. KG and RS were employed by the WALTHAM Centre for Pet Nutrition at the time of the study. None of the other authors have conflicts of interest to declare.

Authorsā€™ contributions

MvZ, MH, KG, and GPJJ conceived and designed the study. MvZ collected the data. RS did the statistical analysis. MvZ and GDL were involved in the chemical analysis. MvZ, MH, KG, RS, GDL, and GPJJ were involved in the data interpretation. MvZ drafted the manuscript and MH, KG, RS, GDL, and GPJJ critically read and edited the manuscript. All authors have read and approved the final version of the manuscript.

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

van Zelst, M., Hesta, M., Gray, K. et al. Biomarkers of selenium status in dogs. BMC Vet Res 12, 15 (2016). https://doi.org/10.1186/s12917-016-0639-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12917-016-0639-2

Keywords