Open Access
How to translate text using browser tools
18 January 2023 Segment-Dependent Gene Expression Profiling of the Cartilaginous Fish Nephron Using Laser Microdissection for Functional Characterization of Nephron at Segment Levels
Takashi Horie, Wataru Takagi, Naotaka Aburatani, Manabu Yamazaki, Mayu Inokuchi, Masaya Tachizawa, Kataaki Okubo, Ritsuko Ohtani-Kaneko, Kotaro Tokunaga, Marty Kwok-Sing Wong, Susumu Hyodo
Author Affiliations +
Abstract

For adaptation to a high salinity marine environment, cartilaginous fishes have evolved a ureosmotic strategy. They have a highly elaborate “four-loop nephron” in the kidney, which is considered to be important for reabsorption of urea from the glomerular filtrate to maintain a high concentration of urea in the body. However, the function and regulation, generally, of the “four-loop nephron” are still largely unknown due to the complicated configuration of the nephron and its many subdivided segments. Laser microdissection (LMD) followed by RNA-sequencing (RNA-seq) analysis is a powerful technique to obtain segment-dependent gene expression profiles. In the present study, using the kidney of cloudy catshark, Scyliorhinus torazame, we tested several formaldehyde-free and formaldehyde-based fixatives to optimize the fixation methods. Fixation by 1% neutral buffered formalin for 15 min resulted in sufficient RNA and structural integrities, which allowed LMD clipping of specific nephron segments and subsequent RNA-seq analysis. RNA-seq from the LMD samples of the second-loop, the fourth-loop, and the five tubular segments in the bundle zone revealed a number of specific membrane transporter genes that can characterize each segment. Among them, we examined expressions of the Na + -coupled cotransporters abundantly expressed in the second loop samples. Although the proximal II segment of the second loop is known for the elimination of excess solutes, the present results imply that the PII segment is also crucial for reabsorption of valuable solutes. Looking ahead to future studies, the segment-dependent gene expression profiling will be a powerful technique for unraveling the renal mechanisms and regulation in euryhaline elasmobranchs.

INTRODUCTION

The marine environment is a high salinity environment, and organisms have adapted to this environment by adopting various physiological strategies. The strategy adopted by cartilaginous fish (sharks, rays, skates, and chimaeras) is well recognized as urea-based osmoregulation (also referred to as a ureosmotic strategy) (Pang et al., 1977; Hazon et al., 2003). The ureosmotic animals maintain their plasma ions at levels approximately half of seawater (SW), while plasma osmolality is slightly hyperosmotic to the surrounding SW through active production and retention of urea. As a result, they do not suffer dehydration even in the high salinity marine environment (Anderson et al., 2007).

Multiple organs are involved in maintaining body-fluid homeostasis of cartilaginous fishes. Among these, the kidney is one of the most important organs to achieve the ureosmotic strategy. Although urea is a key osmolyte to maintain the high plasma osmolality of cartilaginous fish, this osmolyte is freely filtered by the glomerulus. Despite the potential loss, more than 90% of filtered urea is reabsorbed from the glomerular filtrate by the renal tubules (Smith, 1936; Kempton, 1953; Boylan, 1967). Therefore, renal excretion of urea accounts for less than 20% of total urea loss (Goldstein and Forster, 1971; Payan et al., 1973). Meanwhile, the pioneering works by Lacy and Reale (1985a, b) and Hentschel et al. (1993) described the highly elaborated renal nephron as follows. Each nephron begins at a renal corpuscle and forms four loops that traverse repeatedly between two distinct regions, the bundle and sinus zones, within the kidney. Beginning from Bowman's capsule, each nephron consists of the neck, proximal, intermediate, distal, and collecting tubules. The proximal, intermediate, and distal tubules are further subdivided into multiple segments with different morphological characteristics (Lacy and Reale, 1985b; Hentschel, 1987; Kakumura et al., 2015). The characteristic structure of the “four-loop nephron” is considered to be important to perform dedicated roles in each segment and contribute to its overall function, such as urea reabsorption from the glomerular filtrate (see Hyodo et al., 2014).

In order to clarify the function of elasmobranch nephrons, we have conducted mapping studies of various ion pumps, channels, and transporters in the nephron segments. For instance, a facilitative urea transporter (UT; Slc14a1/2) is expressed in the collecting tubule, the final segment of the nephron, while Na+/K+-ATPase (NKA) and Na+,K+,2Cl cotransporter (NKCC2; Slc12a2) are co-localized in the early and late distal tubules, implying that the distal tubules could be segments for desalination (Hyodo et al., 2004; Kakumura et al., 2015; Imaseki et al., 2019; Aburatani et al., 2020, 2022). Based on these results, our group proposed a model for urea reabsorption where urea is reabsorbed in the collecting tubule through the coordinated reabsorption of water and NaCl in the first and/or the third loops in the bundle zone (Hyodo et al., 2014). Likewise, based on our findings, expressions of sulfate transporters (Slc26a1, Slc26a6) were detected in the proximal II (PII) segment, indicating that the PII segment in the sinus zone is involved in sulfate excretion (Hasegawa et al., 2016). In these investigations, expression analyses of target genes or RNA-sequencing (RNA-seq) were conducted using a portion of whole kidney. These whole tissue analyses lack the needed resolution to allow comprehensive gene expression analysis at segment levels, and thus it limits advancement in functional and regulatory studies of each nephron segment.

Laser microdissection (LMD) clips a particular cell(s) or a piece of tissue from a tissue section (Espina et al., 2006; Murray, 2007). Recently, proximal and distal tubule cells were dissected separately by laser capture microdissection and relative mRNA expression levels of segment-specific solute and water transporters were quantified in Atlantic salmon acclimated to fresh water and SW (Madsen et al., 2020). By using the LMD technique, we expect that segment-specific gene expression profiling can also be obtained in the highly elaborate cartilaginous fish nephron. In the LMD clipping, unfixed tissue sections instantly frozen are recommended to preserve RNA integrity for the subsequent gene expression analysis. However, in the case of cartilaginous fish kidney, multiple nephron segments are mixed in a small area. To identify the segment accurately, it is necessary to stain the fixed tissue briefly before conducting LMD. In general, formaldehyde is used in many fixation procedures because it preserves morphology. However, formaldehyde also cross-links nucleic acids and proteins (Park et al., 1996; Chung et al., 2008) and covalently modifies RNA, hindering the RNA extraction, reverse transcription, and quantitation (Auerbach et al., 1977; Masuda et al., 1999).

In the present study, we therefore aimed to optimize the fixation conditions and RNA extraction method, which allowed LMD clipping of a specific nephron segment for subsequent RNA-seq analysis. To confirm the segment-specific clipping by LMD, four established marker genes (sodium-glucose cotransporter 2 [sglt2, slc5a2], nkcc2, Na + -Cl–cotransporter [ncc, slc12a3], and ut) were used to evaluate the clipping accuracy of each segment. Immunoreactive SGLT2 is abundantly localized in the proximal I segment (the first loop) in spiny dogfish Squalus acanthias (Althoff et al., 2006), while nkcc2 and ncc are expressed in the third and fourth loops, and the fourth loop, respectively (Kakumura et al., 2015; Imaseki et al., 2019; Aburatani et al., 2022). ut is almost exclusively found in the final segment (collecting tubule) in all cartilaginous fish nephrons thus far examined (Hyodo et al., 2004; Kakumura et al., 2015; Imaseki et al., 2019; Aburatani et al., 2020). Following the segment-specific gene profiling, selected co-transporter genes that were highly expressed in the PII segment (second loop of nephron) were further examined by in situ hybridization. Since the function of PII segment is less well described, our approach successfully increases the knowledge of the functions of this nephron segment de novo.

MATERIALS AND METHODS

Animals

Adult cloudy catsharks (Scyliorhinus torazame) were transported from Ibaraki Prefectural Oarai Aquarium to the Atmosphere and Ocean Research Institute at The University of Tokyo. They were kept in 3000 L tanks with recirculating natural SW under a constant photoperiod (light: dark; 12 h: 12 h) at 16°C. Fertilized eggs were collected from captive individuals, and eggs and hatched juveniles were maintained in 1000 L tanks under the same environmental conditions. Both adults and juveniles were fed with chopped squid or sardine to satiation twice a week. All experimental procedures were approved by the Animal Ethics Committee of the Atmosphere and Ocean Research Institute of The University of Tokyo (P19-2). The present study was carried out in compliance with the ARRIVE guidelines.

Tissue preparation for laser microdissection (LMD)

Juvenile catsharks were euthanized with 0.02% (w/v) ethyl 3-aminobenzoate methanesulfonate (Sigma-Aldrich, MO, USA), and whole kidneys were dissected out. The dissected kidneys were sliced transversely into approximately 2 mm-thick kidney blocks. Some blocks were frozen immediately in liquid N2, and the other blocks were fixed using various protocols summarized in Table 1. The fixatives used in the present study are as follows: Methacarn (methanol: chloroform: acetic acid = 6: 3: 1) (Puchtler et al., 1970); UMFIX (methanol: polyethylene glycol 300 = 17: 3) (Vincek et al., 2003); Z7 (0.5% zinc chloride, 17.16 mM zinc trifluoroacetate, 0.05% calcium acetate in 0.1 M Tris-HCl, pH 6.4 to 6.7) (Lykidis et al., 2007); 4% PFA (4% paraformaldehyde in 0.1 M phosphate buffer [pH 7.4] containing 150 mM NaCl and 350 mM urea); 10% NBF (commercial 10% formalin neutral buffer solution [Wako Pure Chemical Industries, Osaka, Japan]), and 1% NBF (1% NBF in 20% sucrose).

Table 1.

Fixatives used in the present study and corresponding RNA integrity numbers.

ta_zs220092_001.gif

Fixed blocks of kidney tissue were washed once in Optimal Cutting Temperature (O.C.T.) compound (Sakura Finetek Japan, Tokyo, Japan), and embedded in new O.C.T. compound. All of the embedded blocks were stored at –80°C until sectioning. Frozen blocks were cut at 10 µm using a cryostat maintained at –20°C and consecutive sections were mounted alternatively onto MembraneSlide NF 1.0 PEN (Carl Zeiss Micro Imaging, Jena, Germany) for LMD use or onto MAS-GP-coated glass slides (Matsunami Glass, Osaka, Japan) for reference. This allows a reference histological image of the adjacent section for each LMD-slide. The slides were dried in a cryostat for at least 2 min, followed by fixation in 70% ethanol for 5 min, and then washed with distilled water (DW) for 1 min twice.

Sections on the LMD-slides were stained with Mayer's hematoxylin solution (Wako Pure Chemical Industries) for 1 min, and hematoxylin coloration was enhanced by incubation in DW at 40°C for 1 min. Sections were washed with 70% ethanol for 5 min, dehydrated through graded ethanol of 70%, 95%, and 100% (1 min each), and rapidly dried with a hair dryer. After spraying Liquid Cover Glass (Carl Zeiss Micro Imaging) on the sections, the slides were placed in 50 ml tubes and stored at –80°C until use. All of the staining procedures were performed on ice, except for the coloration enhancement step after hematoxylin staining.

Sections on the reference-slides were stained with the Periodic acid Schiff (PAS) method. Sections were stained with 0.5% Periodic acid solution (Wako Pure Chemical Industries) for 5 min, washed with tap water for 3 min once and then with DW for 5 min three times. Sections were then incubated with Schiff's reagent (Wako Pure Chemical Industries) for 15 min. After rinsing with sulfurous acid solution (10% sodium bisulfite: 1 N HCl: DW = 6: 5: 100) for 3 min three times, the sections were washed with DW for 5 min three times, and then counterstained with Mayer's hematoxylin solution for 3 min. To promote hematoxylin coloration, the sections were washed with running tap water for 15 min. The sections were washed with DW for 3 min, and then dehydrated through the graded ethanol series, cleared in xylene, and mounted with Permount (Thermo Fisher Scientific, Waltham, MA, USA).

RNA extraction and evaluation of the quality of RNA

Total RNA was extracted from two sections (10 µm each) directly from the cryostat, or six scraped-off sections after hematoxylin staining with a FastGene RNA Basic Kit according to the manufacturer's protocol (Nippon Genetics, Tokyo, Japan). RNA quality was assessed by microcapillary electrophoresis with the RNA 6000 Pico LabChip kit (Agilent Biotechnologies, Tokyo, Japan) and the Agilent 2100 bioanalyzer (Agilent Biotechnologies), which assigns an RNA integrity number (RIN) to each RNA electropherogram. This number ranges from 1 (completely degraded RNA sample) to 10 (intact RNA sample) (Schroeder et al., 2006).

LMD and Quartz-Seq

The PALM MicroBeam System (Carl Zeiss Micro Imaging) at Toyo University Itakura Campus was used to excise the selected tissue areas on the LMD slides. The selected tissue areas were determined according to the morphological identification in the reference slides that contain the adjacent sections after PAS-staining. The frozen LMD slides were transferred to the laser microscope (PALM) immediately after being brought to room temperature. The narrow beam ultraviolet laser cut out the areas of interest from surrounding tissue, and the laser was then used to catapult the micro-dissected areas into collecting caps of AdhesiveCap 500 clear tubes (Carl Zeiss Micro Imaging). In the present study, the bundle zone (“Bundle”; a mixture of the first loop, third loop and final segment), the PII segment of the second loop (“2nd”) and the late distal tubule (LDT) of the fourth loop (“4th”) were excised from 12 cryosections, and stored at –80°C until being extracted.

Total RNA from LMD-excised renal segments was extracted using an RNeasy® Micro Kit (Qiagen, Tokyo, Japan) according to the manufacturer's protocol. Total RNA concentration was measured using the QuantiFluor RNA System (Promega, Madison, WI, USA). Quartz-Seq (Sasagawa et al., 2013) was performed to amplify cDNA from the low-yield total RNA. Firstly, total RNA was added to Quartz-RT primer (5′-TATAGAATTCGCGGCCGCTCGCG ATAATACGACTCACTATAGGGCG(T)24-3′), dNTPs (Thermo Fisher Scientific), and RNasin Plus (Promega) at 12 µl solution and the mixtures were treated at 70°C for 90 s and at 35°C for 15 s to prime the RNA with RT primers. Then, SuperScript III (Thermo Fisher Scientific) master mix was added to a final volume of 20 µl and the reverse transcription reaction was performed at 35°C for 5 min and 45°C for 20 min, followed by a 10 min treatment at 70°C to inactivate the reverse transcriptase. The resulting first-strand cDNA was purified with Agencourt AMPure XP beads (Beckman Coulter, Brea, CA, USA) and treated with Exonuclease I (Takara Bio, Shiga, Japan) at 37°C for 30 min to digest the excess RT primers. Exonuclease I was inactivated by incubation at 80°C for 20 min. After that, the Poly-A tailing reaction was performed by adding reaction mixture containing dATP (Thermo Fisher Scientific), RNase H (Thermo Fisher Scientific), and Terminal Transferase (Roche Applied Science, Mannheim, Germany). A reaction mix containing Quartz-Tagging primer (5′-TATAGAATTCGCGGCCGCTCGCGA(T)24-3′) and MightyAmp DNA polymerase (Takara Bio) was prepared on ice and added to the cDNA templates. The double-stranded DNA synthesis reaction was carried out at 98°C for 2 min 10 sec, 40°C for 1 min, and 68°C for 5 min. After the reaction, the solution was placed on ice to stop the reaction. cDNA was then amplified in a 107 µl reaction system with Quartz-PCR primer (5′-NH2-GTATAGAATTCGCGGCCGCTCGCGAT-3′). cDNA amplification was carried out in 18 amplification cycles at 98°C for 10 s, 65°C for 15 s, and 68°C for 5 min. To verify the quality of the amplified cDNA, 4 µl of the reaction product was electrophoresed on a 1.5% agarose gel. A successful cDNA amplification resulted in a smear of DNA at various lengths. Subsequently, the product was purified using Min-Elute PCR Purification Kit (Qiagen) and the cDNA concentration was measured using Qubit 4 (Thermo Fisher Scientific) and the Qubit dsDNA HS Assay Kit (Thermo Fisher Scientific) and Qubit Assay Tubes (Thermo Fisher Scientific).

RNA-seq analysis

RNA-seq library preparation, sequencing, mapping, and gene expression analysis were performed by DNAFORM (Yokohama, Kanagawa, Japan).

Double-stranded cDNA libraries (RNA-seq libraries) were prepared from the cDNA amplified by Quartz-Seq using the SMART-Seq Stranded Kit (Takara Bio) according to the manufacturer's instruction. The “Bundle”, the “2nd”, and the “4th” samples of catshark kidney were analyzed (n = 3). RNA-seq libraries were sequenced using 150 bp paired end reads on a NextSeq 500 instrument (Illumina, San Diego, CA, USA). Qualified reads were mapped to the catshark Genome assembly: Storazame_v1.0 using STAR (version 2.7.2) (Dobin et al., 2013). Reads on annotated genes were counted using featureCounts (version 1.6.1) (Liao et al., 2014). Transcripts per million (TPM) values were calculated from the mapped reads and normalized to total counts of transcript, and the TPM values for each segment were arranged in a descending order. The expression levels among the three segments (Bundle, 2nd, and 4th) were then statistically analyzed. All of the coding nucleotide sequences predicted on the catshark genome were subjected to NCBI blastx (Expect threshold: 0.05) against the non-redundant protein sequences to identify the putative proteins. The gene names, protein names, and Gene Ontology (GO) terms corresponding to all of the identified proteins were acquired through Uniprot database. The genes containing GO terms of molecular function “transporter activity” and cellular component “plasma membrane” were then short-listed. The raw sequenced reads were deposited in the DNA Data Bank of Japan (DDBJ) under the accession numbers DRR413293–DRR413301.

cDNA cloning

Complementary DNA cloning was performed as previously described in detail (Takagi et al., 2017) using frozen kidney of juvenile catshark. Total RNA was extracted using ISOGEN (Nippon Gene, Toyama, Japan). DNase-treated total RNA (1 µg) was reverse-transcribed into cDNA using a High Capacity cDNA Reverse Transcription Kit (Thermo Fisher Scientific), following the manufacturer's instructions. Putative nucleotide sequences of mRNAs encoding catshark SGLT2, NKCC2, NCC, UT, Slc1a3, Slc5a8, Slc5a11, Slc13a3, and β-actin were obtained by BLAST search on catshark transcriptome database Squalomix ( https://transcriptome.riken.jp/squalomix/) (Hara et al., 2018). Partial cDNA fragments, which were amplified with Kapa Taq Extra DNA polymerase (Kapa Biosystems, Boston, MA, USA) and gene-specific primer sets (see  Supplementary Table S1 (zs220092_TableS1.pdf)), were ligated into pGEMT Easy plasmid vector (Promega). The nucleotide sequences were determined using a DNA sequencer (ABI PRISM 3130, Life Technologies, Carlsbad, CA, USA). The sequence data were deposited in the DDBJ under the accession numbers LC739024 (ncc), LC739025 (sglt2), LC739026 (nkcc2), LC739027 (ut), LC739028 (slc1a3), LC739029 (slc5a8), LC739030 (slc5a11), and LC739031 (slc13a3).

In situ hybridization

Kidneys of adult female catsharks fixed in Bouin's solution without acetic acid were embedded in Paraplast (McCormick Scientific, Richmond, IL, USA). Serial sections were prepared at 7 µm and mounted onto MAS-coated slides (Matsunami Glass). The cloned cDNAs were amplified from the plasmids with Primestar GXL DNA polymerase (Takara Bio) using M13 forward and reverse primers. The PCR products were subsequently purified using Wizard SV Gel and PCR Clean-up System (Promega) and used to synthesize digoxigenin (DIG)-labeled antisense and sense cRNA probes with a DIG RNA labeling kit (Roche Applied Science), following the manufacturer's protocols. Hybridization and washing were conducted using a previously described protocol (Aburatani et al., 2022). Stained sections were counterstained with Kernechtrot Stain Solution (MUTO PURE CHEMICALS, Tokyo, Japan) and mounted using Permount (Thermo Fisher Scientific) (sglt2, nkcc2, ncc, and ut), or mounted using CC/Mount (Diagnostic BioSystems, Pleasanton, CA, USA) without counterstaning for weakly stained sections (slc1a3, slc5a8, slc5a11, and slc13a3).

Quantitative real-time PCR (qPCR)

The gene expression levels were measured by qPCR using 7900HT Fast Real Time PCR System (Applied Biosystems, Forster city, CA, USA) with a KAPA SYBR FAST qPCR Kit (Kapa Biosystems) and gene-specific primer sets (see  Supplementary Table S1 (zs220092_TableS1.pdf)). The cloned plasmids containing the target sequences were serially diluted as standard templates for quantification. The expression levels of β-actin mRNA were used as internal control.

Statistical analysis

Values are presented as means ± S.E.M. The qPCR results were analyzed by one-way ANOVA followed by Tukey's multiple comparison test or Kruskal-Wallis test followed by Dunn's multiple comparison test. The TPM values among the three segments (Bundle, 2nd, and 4th) were analyzed by one-way ANOVA followed by Tukey's multiple comparison test. P values < 0.05 were considered as statistically significant. All analyses were performed using Prism Ver. 9 for Windows (GraphPad, San Diego, CA, USA).

RESULTS

Evaluation of tissue preparation methods

For successful LMD dissection and subsequent RNA-seq, an optimal fixation method that provides substantial histological resolution and RNA integrity is required. To avoid possible cross-linking damage to RNA integrity by formaldehyde, several formaldehyde-free fixatives, Methacarn, UMFIX, and Z7, were examined (Table 1). The RNA quality was assessed by the RNA integrity number (RIN) in an RNA electropherogram. For each fixation method, RNA samples were prepared from 1) the tissues following cryostat sectioning without mounting on glass slides (“Cryostat”) and 2) the tissues scraped off from glass slides following hematoxylin staining (“Scraped off”). The RNA samples prepared from the unfixed tissues had RIN values between 5.73 ± 0.66 (“Cryostat”) and 4.60 ± 1.10 (“Scraped off”). Most RNA samples from fixed tissues with formaldehyde-free fixatives had RIN values higher than 5, but some samples from the Methacarn-fixed for 1 h, UMFIX-fixed for 3 h, and Z7 for 12h had RIN values lower than 5 (Table 1). On the other hand, RNA samples prepared from the formalin-fixed paraffin-embedded (4% PFA overnight) tissues had RIN values less than 3.5, which was considerably lower than those of formaldehyde-free fixatives.

Figure 1A shows the transverse section of the formaldehyde-fixed kidney of adult catshark. The kidney consists of the sinus zone and the bundle zone, with single nephrons traversing repeatedly between the two zones (see Fig. 1D). The sinus zone is filled with the second and the fourth loops (Fig. 1B), while five tubular segments (the descending and ascending segments of the first and third loops and the collecting tubule) are aggregated in the bundle zone. The five bundle segments were closely wrapped by a sac-like peritubular sheath (circled in Fig. 1C). Figure 1E, G represent the sinus zone of UMFIX- and Z7-fixed kidney samples, respectively, and histological preservation by the formaldehyde-free fixatives was too low for identifying specific segments for the subsequent LMD. In addition, during the preparation of the adjacent sections for PAS staining, we noticed that histological structures were easily distorted in the formaldehyde-free fixed sections, and thus confident identification of specific tubular segments on the reference slides was not possible (Fig. 1F, H).

Fig. 1.

Anatomical features of the cartilaginous fish kidney and histological preservation by the formaldehyde-free and formaldehyde-based fixatives for LMD. (A) Transverse section of the formaldehyde-fixed adult catshark kidney stained with PAS. SZ, sinus zone; BZ, bundle zone. (B) Magnified image of the sinus zone. The number labels 2 and 4 mark the PII segment and the LDT, respectively. Note that, in the case of PII segment, the brush border of the apical surface is stained with PAS. (C) Magnified view of the bundle zone. The circle represents five tubular segments of a single nephron. (D) Schematic illustration of cartilaginous fish nephron. CT, collecting tubule; EDT, early distal tubule; IS, intermediate segment; LDT, late distal tubule; NS, neck segment; PI and PII, proximal segment I and II, respectively; RC, renal corpuscle. (E–L) Histological images of catshark kidney for LMD (hematoxylin-stained [E, G, I, K]) and reference (PAS-stained [F, H, J, L]) slides. Tissues were fixed with UNFIX for 12 h (E, F), Z7 for 12 h (G, H), 10% NBF for 1 min (I, J), or 1% NBF for 15 min (K, L). The number label 2 represents the PII segment of second loop. Scale bars: 1 mm (A), 100 µm (B, C), 300 µm (E–L).

fi_zs220092_001.jpg

The above results suggest that the formaldehyde-free fixatives examined so far are not suitable for excising specific nephron segments using LMD. Therefore, the use of formaldehyde-based fixatives was examined (Table 1). The low RIN score in the 4% PFA overnight samples might be due to excessive cross-linking and modification of nucleic acids by formaldehyde, and possible damage to nucleic acids by formic acid produced by the oxidation of formaldehyde. Therefore, we next examined various fixation conditions for frozen tissues using a commercial 10% neutral buffered formalin (NBF; approximately 3.7% paraformaldehyde) solution. Four conditions were examined: 1) 10% NBF for 5 min under negative pressure, 2) 10% NBF for 1 min, 3) 1% NBF in 20% sucrose for 1 min, and 4) 1% NBF in 20% sucrose for 15 min (Table 1). Fixation in 10% NBF under negative pressure is considered to facilitate formaldehyde penetration into the tissues, thereby minimizing the exposure time to formaldehyde. However, the RIN scores were 5.80 and 1.00, indicating that the negative pressure protocol was not compatible with RNA extraction. On the other hand, the reduced fixation time for 1 min in 10% NBF achieved high RIN scores of 7.90 and 8.00. Furthermore, a low concentration of NBF at 1% for 1 min or 15 min also improved RIN score to over 7.0 except for one sample fixed in 1% NBF for 15 min (Table 1).

Cryostat sections with formaldehyde-based fixatives improved histological preservation compared to those with formaldehyde-free fixatives (Fig. 1I, K). When adjacent sections were stained with PAS, the proximal segment of the second loop and the distal segment of the fourth loop were clearly distinguishable by the presence and absence, respectively, of brush border on the apical surface of lumen (Fig. 1J, L). Although both fixation times of 1 min and 15 min resulted in satisfactory RIN values, the longer fixation time reduced the hastiness of the procedures and thus produced a more consistent fixation condition. Collectively, fixation in a low-concentration formaldehyde (1% NBF) for 15 min maintains a reasonable tissue morphology and RNA quality, and thus allows LMD clipping of specific renal tubules for subsequent RNA-seq.

Fig. 2.

Expression of the marker genes used in the present study. (A–D) Na+-coupled glucose transporter (sglt2mRNA), (E–H) Na+,K+,2Clcotransporter (nkcc2 mRNA), (I–L) Na+,Cl cotransporter (ncc mRNA), (M–P) urea transporter (ut mRNA). (A, E, I, M) Low magnification of catshark kidney showing both the bundle and sinus zones. Magnified views of the bundle zone (B, F, J, N) and the sinus zone (C, G, K, O). The marks represent the following: open arrowheads, the ascending part of the PIa segment (“ii” in the tubules); arrows, the EDT of the third loop (“iv” in the tubules); arrowheads, the collecting tubule (“v” in the tubules); aLDT and pLDT, anterior and posterior LDT, respectively. (D, H, L, P) Schematic images representing localization of sglt2 (D), nkcc2 (H), ncc (L), and ut (P) mRNAs that are shown in red color. aPIa, ascending part of PIa segment. Scale bars: 200 µm (A, E, I, M) and 50 µm (B, C, F, G, J, K, N, O).

fi_zs220092_002.jpg

Evaluation of segment specificity of LMD samples

Four membrane transporter genes were selected as markers as they are specifically and abundantly expressed in particular segments of the nephron according to previous studies. Segment-specific expressions of these marker genes were confirmed in the catshark kidney (Fig. 2). Signals for sglt2 mRNA were detected in the bundle zone, but not in the sinus zone (Fig. 2A–C). In the bundle zone, sglt2 mRNA signals were observed only in a single nephron segment (open arrowheads in Fig. 2B). Observation using serial sections revealed that the sglt2 mRNA-expressing tubule is the ascending (posterior) limb of the PIa segment (Fig. 2D). On the other hand, nkcc2 mRNA signals were found both in the bundle and sinus zones (Figs. 2E–G). In the bundle zone, intense nkcc2 mRNA signals were detected in tubules comprised of the largest columnar cells, indicating that the early distal tubule (EDT; filled arrows) of the third loop expresses nkcc2 mRNA (Fig. 2F, H). In the sinus zone, nkcc2 mRNA was detected in tubules comprised of small and thin squamous cells, indicating that nkcc2 mRNA is expressed in the fourth loop (the late distal tubule; LDT) (Fig. 2G, H). The ncc mRNA signals were observed in the LDT and the transitional region between LDT and CT (Fig. 2I, K). In the sinus zone, however, the posterior half of LDT was intensely stained by ncc mRNA signals, which was different from the anterior half of the LDT that was intensely stained by nkcc2 mRNA (Fig. 2G, H, K, L). The ut mRNA signals were found in the tubules in bundle zone (Fig. 2M), and the ut mRNA positive segment (filled arrowheads in Fig. 2N) did not overlap with segments expressing sglt2 (open arrowheads) or nkcc2 (filled arrows) mRNAs, indicating that the ut mRNA-positive segment is the collecting tubule and preceding transitional region between LDT and CT (Fig. 2P). These results confirmed that the four marker genes are effective for evaluating segment specificity of LMD samples.

In the present LMD study, three regions of catshark nephron were excised: 1) five tubular segments wrapped by the peritubular sheath in the bundle zone (“Bundle”, the descending and ascending segments of the first and third loops, and the collecting tubule), 2) the PII segment of the second loop (“2nd”), and 3) the LDT of the fourth loop (“4th”). Although no significant difference was observed in marker gene expressions among the LMD samples due to the large variations in observed data, qPCR analysis revealed that the expression of four marker genes in the LMD samples corresponded to the local abundance of mRNA shown by the results of in situ hybridization (Fig. 3). Expressions of sglt2, nkcc2, and ut were observed in the “Bundle” samples, in which the first loop, third loop, and collecting tubule were included. Expression of nkcc2 mRNA was also found in the “4th” samples. Expression of ncc mRNA was detected in the “4th” samples, while none of the four genes had high expression in the “2nd” samples. These results further demonstrated that contamination of other segments is negligible for the obtained LMD samples.

Expression profiles of membrane transporter mRNAs in the LMD samples

Segment-dependent mRNA expression profiles were examined using three individuals. In the present study, we focused on membrane transport proteins, including the solute carrier (Slc) family, for deducing the reabsorbing and/or secreting functions of each nephron segment. By searching GO terms as described in the Methods section, we shortlisted the top 30 highly expressed membrane transporter genes in each LMD segment (Tables 2 to 4), and we statistically analyzed their expression levels among the three LMD segments. As a result, five, 13, and four genes were found to be significantly expressed in the “Bundle”, “2nd”, and “4th” segments, respectively, which characterize the functions of those segments (Tables 2 to 4).

Fig. 3.

Expression of the marker genes in the LMD samples for “Bundle” (five tubular segments in the bundle zone), “2nd” (the PII segment of second loop), and “4th” (the LDT of fourth loop). Data are presented as means ± S.E.M. with each sample value as an open circle (n = 3). No significant differences were observed in expressions of the marker genes among the LMD samples due to the large variations.

fi_zs220092_003.jpg

Many segment-dependent genes were detected in the second loop samples, but generally these have not been investigated yet in the elasmobranch kidneys. Therefore, we selected four genes (slc13a3, slc5a8, slc5a11, and slc1a3) and further investigated their localization by in situ hybridization (Fig. 4). slc13a3 has the highest TPM values among Slc family genes in the second loop, and its expression was intensely detected in the sinus zone (Fig. 4A). The slc13a3-positive segments were relatively large and did not express nkcc2 and ncc mRNAs (Fig. 4P), indicating that slc13a3 gene expression is specific to the PII segment of second loop. The intense signals observed in the second loop by in situ hybridization were consistent with the results of RNA-seq, where the TPM value of slc13a3 mRNA was significantly higher in the “2nd” sample than that in the “4th” and “Bundle” samples.

Two genes encoding SLC5 members (slc5a8 and slc5a11) were listed as genes abundantly expressed in the second loop based on RNA-seq. However, the localization patterns of slc5a8 and slc5a11 mRNAs in the nephron were different. The slc5a11 mRNA signals were found only in the sinus zone (Fig. 4F). The slc5a11-positive tubules were the same tubules expressing slc13a3 mRNA, although the signal intensity of slc5a11 mRNA was much lower than that of slc13a3 mRNA. Meanwhile, slc5a8 mRNA signals were faintly detected in one segment within the bundle zone (arrows in Fig. 4M). The slc5a8 mRNA-positive segment co-expressed sglt2 mRNA (Fig. 4R) but was negative for nkcc2 (Fig. 4Q) and ut (Fig. 4S) mRNAs, suggesting that slc5a8 is expressed in the ascending part of the PIa segment of the first loop.

Fig. 4.

Expression of slc13a3 (A–E), slc5a11 (F–J) and slc5a8 (K–O) mRNAs in the sinus zone (A, B, F, G, K, L) and the bundle zone (C, D, H, I, M, N) of catshark kidney. Sections were hybridized with antisense (A, C, F, H, K, M) or sense (B, D, G, I, L, N) probes. (P–S) Adjacent sections were stained with ncc ([P] sinus zone), nkcc2 ([Q] bundle zone), sglt2 ([R] bundle zone), or ut ([S] bundle zone) mRNA. The labels ii, iv, and v mark the ascending part of PIa segment, the EDT, and the collecting tubule, respectively. Both slc13a3 and slc5a11 mRNA signals were localized in the PII segment of the second loop (A, F), while the weak mRNA signals of slc5a8 were found in the ascending part of PIa segment (arrows in [M]), where mRNA signals of sglt2 were co-localized (arrows in [R]). Schematic images represent localization of slc13a3 (E), slc5a11 (J), and slc5a8 (O) mRNAs that are shown in red color. Scale bars, 50 µm.

fi_zs220092_004.jpg

Although slc1a3 mRNA was listed as a gene characteristic of the second loop, expression levels were too low to be detected by in situ hybridization. In all in situ hybridization experiments, negative control slides using sense cRNA probes resulted in no signals (Fig. 4B, D, G, I, L, N).

DISCUSSION

The cartilaginous fish nephron forms four loops that traverse repeatedly between the bundle and sinus zones. The single nephron is composed of over 10 morphologically distinct segments (Lacy and Reale, 1985b; Hentschel, 1987; Kakumura et al., 2015), and those multiple segments most likely have separate secretory and/or reabsorption functions (Hyodo et al., 2014; Hasegawa et al., 2016). However, due to technical limitations, comprehensive gene expression analyses have been conducted only using whole kidney (or a portion of whole kidney) homogenates, which has limited the search for segment-dependent gene expression that can characterize each segment. This study is the first to apply the LMD technique to analyze gene expression in the cartilaginous fish nephron, and we successfully obtained gene expression profiles that characterized three segments of the catshark nephron.

Successful clipping and gene expression profiling in each segment requires that both RNA integrity and histological preservation be relatively high. Since formaldehyde-based fixatives form cross-linkages between mRNAs and surrounding proteins, RNA integrity is often higher when the tissue is treated with formaldehyde-free fixatives. Indeed, our results showed that alcohol-based and zinc-based fixatives such as Methacarn, UMFIX, and T7 maintained high nucleic acid qualities (Table 1). However, the histological preservation by these fixatives was too low to provide essential guidance for clipping of specific nephron segments by LMD. Therefore, we tested several formaldehyde-based fixatives with modifications of fixation strength and time. Although an initial trial with 10% NBF for 5 min under negative pressure resulted in low RIN score (less than 6.0), reduction of fixation time without negative pressure or reduction of fixative concentration to 1% NBF improved the RIN scores (higher than 7.0). For RNA-seq analysis, RIN scores higher than four (Romero et al., 2014) or five (Fleige and Pfaffl, 2006) are recommended. Indeed, consistent PCR amplification of segment-specific marker genes was successfully obtained using the RNA samples excised by LMD from the sections fixed in 1% NBF for 15 min.

Table 2.

List of top 30 highly expressed membrane transporter genes in the “Bundle” samples.

ta_zs220092_002.gif

In the present study, the five tubular segments (the descending and ascending segments of first and third loops and the collecting tubule) were excised together by LMD. EDT of the third loop, comprised of large columnar cells, is distinguishable from the other four segments within the bundle. Except for EDT, however, morphological characteristics under general staining methods (e.g., HE or PAS) are not sufficient to identify other segments. To identify each segment in the tubular bundle, it is necessary to stain the reference section or the LMD section for marker genes/proteins. The present study confirmed that SGLT2, NKCC2, and UT can be used as specific marker genes/proteins for the ascending segment of first loop, the EDT of third loop, and the collecting tubule, respectively. The formaldehyde-based fixative protocol established in the present study is appropriate for in situ hybridization and immunohistochemistry for improving the LMD protocol in the future.

In the kidney of cloudy catshark, the localization of sglt2 mRNA signals was confined to the ascending segment of PIa of the first loop. To date, there has been no report showing that the PIa segment is further subdivided into segments with different functions. In the kidney of spiny dogfish, immunoreactive SGLT2 is widely localized in the PIa, PIb, PIIb, LDT, and collecting tubule (Althoff et al., 2006). The observed difference between cloudy catshark and spiny dogfish could be due to methodology, where the former relied on probe specificity for the mRNA and the latter depended on antibody specificity for the protein. The localization of sglt2 mRNA in the catshark nephron was consistent with that in mammalian nephrons, where SGLT2 is expressed predominantly on the luminal surface of cells of the first part of the proximal tubule (S1 and S2 segments) (Gerich, 2010). The SGLT2 transporter protein is responsible for reabsorbing approximately 90% of the glucose filtered at the glomerulus in mammalian nephrons (see Marsenic, 2009), thus it is reasonable to consider that glucose is predominantly reabsorbed in the early proximal tubules in different vertebrate lineages, including cartilaginous fishes.

Table 3.

List of top 30 highly expressed membrane transporter genes in the “2nd” samples.

ta_zs220092_003.gif

Segment-dependent gene expression profiles

In the present study, we successfully obtained lists of genes abundantly expressed in the second loop, fourth loop, and the tubular bundle from the RNA-seq analyses on the LMD samples. We focused on membrane transporter genes, including the slc family, as they play crucial roles in reabsorption and secretion of solutes across the epithelial cells of renal nephrons (Hediger et al., 2004). We found that the marker genes used in the present study were on the lists: nkcc2, sglt2, and ut genes on the “Bundle” list, and nkcc2 and ncc on the “4th” list.

Among the genes on the lists, we further focused on the Na+-coupled cotransporters abundantly expressed in the second loop because of the following reasons: 1) Na+coupled cotransporters are the secondarily active transporters directly contributing to reabsorption or secretion of solutes in nephron segments, and 2) little information has been available on the function of proximal segment(s) in the second loop except possible divalent ion secretion (Hasegawa et al., 2016). The most abundantly expressed slc gene in the second loop was slc13a3. In situ hybridization clearly showed that slc13a3 mRNA is abundantly expressed in the PII segment of the second loop. The Slc13 family is comprised of five members that encode electrogenic Na+coupled cotransporters for anions or di- and tri-carboxylates (Markovich and Murer, 2004). In mammalian kidneys, Slc13a3 is localized on the basolateral membrane of the proximal tubule, and transport intermediates of the Krebs cycle such as succinate, citrate, and α-ketoglutarate (Bergeron et al., 2013). Slc13a3 also plays an important role in the renal elimination of xenobiotics by secreting organic anions into urine.

Table 4.

List of top 30 highly expressed membrane transporter genes in the “4th” samples.

ta_zs220092_004.gif

The renal secretion of organic anions is considered to be achieved with several steps as follows (Breljac et al., 2016): 1) primary active NKA creates an inwardly directed Na+ gradient, 2) secondary active Na+-coupled dicarboxylate cotransporter (Slc13a3) creates an outwardly directed gradient of α-ketoglutarate, and 3) tertiary active organic anion transporters (such as Slc22a6 and Slc22a8) transport organic anions into epithelial cells by exchanging organic anions with α-ketoglutarate. All three components (NKA, Slc13a3, and Slc22a6 or Slc22a8) were localized on the basolateral membrane (see Breljac et al., 2016). On the apical membrane, 4) multiple ATP-binding cassette (ABC) and Slc transporters secrete organic anions into the lumen of the proximal tubule (Schlatter et al., 2006). In this regard, our RNA-seq data supported such a scheme for the shark nephron since the transcripts of slc22a6 and abcc1 genes were on the list of highly expressed membrane transporter genes in the second loop. These results suggest that the PII segment of the second loop contributes to secretion of organic anions in the elasmobranch nephron, in addition to divalent ion secretion.

In the list of highly expressed transporters in the second loop, two Slc5 members were found, slc5a8 and slc5a11. The slc5a11 mRNA was localized in the PII segment of the second loop, while slc5a8 mRNA signals were co-localized with sglt2 mRNA in the ascending part of the PIa segment in the bundle zone (Figure 4). The Slc5 family is known as the Na + -coupled glucose cotransporter family, but these proteins also transport monocarboxylates such as lactate and pyruvate, short-chain fatty acids, and ketone bodies (Plata et al., 2007). Slc5a8 transports monocarboxylates and ketone bodies, and thus was named Na + -coupled monocarboxylate transporter (SMCT1) (Gopal et al., 2004). In mammalian kidneys, the reabsorption fraction of lactate from the glomerular filtrate is over 95%. In the renal tubular system, lactate and ketone bodies are also used as the major fuels for active reabsorption of electrolytes and nutrients (Yanase et al., 2008). Since an increased reliance on ketone bodies as an energy source was observed following prolonged starvation in elasmobranchs (Zammit and Newsholme, 1979), Slc5a8 could have a regulatory role in energy metabolism in the elasmobranch nephron. Since both sglt2 and slc5a8 mRNAs were expressed in the ascending part of PIa segment, it is likely that this segment is important for reabsorption of energy substrates such as glucose, lactate, and ketone bodies. On the other hand, Slc5a11 (SGLT6) is a Na + -coupled myo-inositol cotransporter in the apical membrane of renal tubule cells, where it mediates myo-inositol reabsorption from the primary urine (Kottgen et al., 2018). Myo-inositol and its derivatives have functions including regulation of ion channels, metabolic flux, mRNA transcription and translation, stress response, and embryonic development (Kiani et al., 2021). In the kidney, myo-inositol is also important as a protective osmolyte against osmotic stress.

slc1a3, which is another slc gene examined in the present study, is a member of glutamate/neutral amino acid transporters. The TPM value of slc1a3 was less than 5% of that of slc13a3, and we could not confirm the gene expression site by in situ hybridization due to the low expression level. Slc1a3 is a glial-type glutamate transporter (also known as a glutamate-aspartate transporter, GLAST, or an excitatory amino acid transporter, EAAT1) abundantly expressed in the cerebellum in mammals (Kanai and Hediger, 2004). Slc1a3 is responsible for glutamate uptake in the brain, which is crucial for glutamatergic neurotransmission by removing glutamate from the synaptic cleft (Berger and Hediger, 2006). However, expression of the Slc1a3 gene has not been proven in mammalian renal tubules. Instead, another glutamate transporter, Slc1a1, is localized in the apical membrane of proximal tubules in the mammalian kidneys (Berger and Hediger, 2006). Slc1a1-knockout mice exhibited dicarboxylic aminoaciduria, confirming its crucial role for reabsorbing glutamate in the proximal tubule (Kanai and Hediger, 2004). Further investigations on slc1a1 and slc1a3 mRNAs are required to understand their roles in the kidney of elasmobranchs.

Seawater is an environment of excess divalent ions, and thus the kidneys of marine teleosts and cartilaginous fishes must excrete divalent ions. Excretion of NaCl is achieved by the branchial ionocytes (teleost fishes) and the rectal gland (cartilaginous fishes), whereas it has been well documented that kidneys are responsible for the excretion of divalent ions (see Takvam et al., 2021). We previously revealed co-localization of sulfate transporters Slc26a1 and Slc26a6 in the basolateral and apical membranes, respectively, in the PII segment of the elephant fish nephron, implying that the PII segment of the second loop is a secretory segment for excess solutes (Hasegawa et al., 2016). In the present study, we found intense expression of slc13a3 mRNA in the PII segment. As mentioned above, Slc13a3 most likely contributes to the renal elimination of xenobiotics by secreting organic anions into urine, further supporting the secretory role of the PII segment. On the other hand, expression of Na+-coupled myo-inositol transporter (slc5a11) mRNA was also found in the PII segment. Although the cellular localization of Slc5a11 is uncertain in the cartilaginous fish nephron, distribution of these transporters in the apical membrane has been well documented in the mammalian nephron. These results imply that the PII segment may not be a simple secretory segment, but it may also be involved in reabsorption of valuable solutes. In the catshark nephron, at least four segments can be identified in the proximal tubule: the descending PIa (the first loop with no expression of sglt2 mRNA), the ascending PIa (the first loop with sglt2 and slc5a8 mRNAs), PIb, and PII (with slc13a3 and slc5a11 mRNAs). To understand further the function of each segment, finer-scale clipping of each segment by the LMD is required, with the guidance of reference section stained with in situ hybridization and/or immunohistochemistry.

Among the four slc transporters examined in the second loop, expressions of slc5a11 and slc13a3 were only detected in the PII segment by in situ hybridization, confirming the quantitative results of RNA-seq from the LMD samples. However, expression of slc5a8 mRNA was detected in the ascending part of PIa segment in the bundle zone, which is inconsistent with the low TPM values of slc5a8 in the “Bundle” samples (4.4 ± 1.0). Since the TPM values of sglt2 (slc5a2) that marked the PIa segment were significantly high in the “Bundle” samples, it is uncertain why the TPM values of slc5a8 mRNA were low in the “Bundle” samples. Perhaps unannotated contig(s) encoding Slc5a8 exists in the RNA-seq data of the “Bundle” samples, leading to incomplete counting of transcripts.

Conclusion and perspectives

In the present study, we developed a protocol for clipping specific nephron segments of catshark kidney by LMD, and successfully conducted subsequent RNA-seq analysis for profiling the segment-dependent gene expressions. Most cartilaginous fishes are obligatory marine species, while a limited number of sharks and rays have the ability to inhabit low salinity environments (Ballantyne and Robinson, 2010). During the acclimation to low salinity environments, euryhaline sharks and rays produce large volumes of dilute urine to excrete excess water; they do this by increasing glomerular filtration and solute reabsorption in the renal tubules (Janech et al., 2006; Imaseki et al., 2019; Aburatani et al., 2022). Indeed, we found marked increase in the expression of the ncc gene in the LDT of FW-acclimated bull shark, Carcharhinus leucas (Imaseki et al., 2019), and of the nkcc2 gene in the EDT of FW-acclimated red stingray, Hemitrygon akajei (Aburatani et al., 2020). It is likely that water reabsorption in renal tubules would also be decreased in low salinity environments. Currently, regulatory mechanisms of renal function remain poorly resolved in euryhaline elasmobranchs, and therefore, the segment-dependent gene expression profiling that we report here will be a powerful technique for unraveling the renal mechanisms and regulation in euryhaline elasmobranchs.

ACKNOWLEDGMENTS

We are grateful to Ms. Kiriko Ikeba and Dr. Taro Watanabe of AORI for help with keeping the animals, and Ms. Sugako Watanabe and Mr. Koya Shimoyama of AORI for preparing illustrations and figures. We are grateful to Prof. Christopher A. Loretz of the State University of New York at Buffalo for his critical reading of this manuscript. This study was supported by Grants-in-Aid for Scientific Research from the Japan Society for the Promotion of Science to SH (JSPS KAKENHI 17H03868 and 19K22414) and to WT (JSPS KAKENHI 22K15153), and a Grant-in-Aid for JSPS Fellows to NA (21J20882). NA is supported by JSPS Research Fellowships for Young Scientists.

COMPETING INTERESTS

The authors have no competing interests to declare.

AUTHOR CONTRIBUTIONS

TH and SH designed the study. TH, WT, NA, MY, MI, and MT performed the experiments and collected data. TH, WT, NA, and KT kept the animals and contributed to sample collection. WT, MI, KO, RO-K, and MK-SW supervised the experiments. TH and SH wrote the first draft of the manuscript, and WT, NA, MK-SW, and SH largely contributed to the revision step. All of the authors contributed substantial input to the final version of the manuscript.

SUPPLEMENTARY MATERIALS

Supplementary material for this article is available online. (URL:  https://doi.org/10.2108/zs220092)

 Supplementary Table S1 (zs220092_TableS1.pdf). Primer sets used for quantitative PCR and in situ hybridization.

REFERENCES

1.

Aburatani N, Takagi W, Wong MKS, Kadota M, Kuraku S, Tokunaga K, et al. (2020) Facilitated NaCl uptake in the highly developed bundle of the nephron in Japanese red stingray Hemitrygon akajei revealed by comparative anatomy and molecular mapping. Zool Sci 37: 458–466 Google Scholar

2.

Aburatani N, Takagi W, Wong MKS, Kuraku S, Tanegashima C, Kadota M, et al. (2022) Molecular and morphological investigations on the renal mechanisms enabling euryhalinity of red stingray Hemitrygon akajei. Front Physiol 13: 953665 Google Scholar

3.

Althoff T, Hentschel H, Luig J, Schutz H, Kasch M, Kinne RKH (2006) Na+-D-glucose cotransporter in the kidney of Squalus acanthias: molecular identification and intrarenal distribution. Am J Physiol Regul Integr Physiol 290: R1094–R1104 Google Scholar

4.

Anderson WG, Taylor JR, Good JP, Hazon N, Grosell M (2007) Body fluid volume regulation in elasmobranch fish. Comp Biochem Physiol A 148: 3–13 Google Scholar

5.

Auerbach C, Moutschen-Dahmen M, Moutschen J (1977) Genetic and cytogenetical effects of formaldehyde and related compounds. Mutat Res 39: 317–362 Google Scholar

6.

Ballantyne JS, Robinson JW (2010) Freshwater elasmobranchs: a review of their physiology and biochemistry. J Comp Physiol B 180: 475–493 Google Scholar

7.

Berger UV, Hediger MA (2006) Distribution of the glutamate transporters GLT-1 (SLC1A2) and GLAST (SLC1A3) in peripheral organs. Anat Embryol 211: 595–606 Google Scholar

8.

Bergeron MJ, Clemencon B, Hediger MA, Markovich D (2013) SLC13 family of Na+-coupled di- and tri-carboxylate/sulfate transporters. Mol Aspects Med 34: 299–312 Google Scholar

9.

Boylan JW (1967) Gill permeability in Squalus acanthias. In“Sharks, Skates and Rays” Ed by PW Gilbert, RF Mathewson, DP Rall, Johns Hopkins Press, Baltimore, pp 197–206 Google Scholar

10.

Breljak D, Ljubojevic MI, Hagos Y, Micek V, Eror DB, Madunic IV, et al. (2016) Distribution of organic anion transporters NaDC3 and OAT1-3 along the human nephron. Am J Physiol Renal Physiol 311: F227–F238 Google Scholar

11.

Chung JY, Braunschweig T, Williams R, Guerrero N, Hoffmann KM, Kwon M, et al. (2008) Factors in tissue handling and processing that impact RNA obtained from formalin-fixed, paraffin-embedded tissue. J Histochem Cytochem 56: 1033–1042 Google Scholar

12.

Dobin A, Davis CA, Schlesinger F, Drenkow J, Zaleski C, Jha S, et al. (2013) STAR: ultrafast universal RNA-seq aligner. Bioinformatics 29: 15–21 Google Scholar

13.

Espina V, Wulfkuhle JD, Calvert VS, VanMeter A, Zhou W, Coukos G, et al. (2006) Laser-capture microdissection. Nat Protoc 1: 586–603 Google Scholar

14.

Fleige S, Pfaffl MW (2006) RNA integrity and the effect on the real-time qRT-PCR performance. Mol Aspects Med 27: 126–139 Google Scholar

15.

Gerich JE (2010) Role of the kidney in normal glucose homeostasis and in the hyperglycaemia of diabetes mellitus: therapeutic implications. Diabet Med 27: 136–142 Google Scholar

16.

Goldstein L, Forster RP (1971) Osmoregulation and urea metabolism in the little skate Raja erinacea. Am J Physiol 224: 367–372 Google Scholar

17.

Gopal E, Fei YJ, Sugawara M, Miyauchi S, Zhuang L, Martin P, et al. (2004) Expression of slc5a8 in kidney and its role in Na+coupled transport of lactate. J Biol Chem 279: 44522–44532 Google Scholar

18.

Hara Y, Yamaguchi K, Onimaru K, Kadota M, Koyanagi M, Keeley SD, et al. (2018) Shark genomes provide insights into elasmobranch evolution and the origin of vertebrates. Nat Ecol Evol 2: 1761–1771 Google Scholar

19.

Hasegawa K, Kato A, Watanabe T, Takagi W, Romero MF, Bell JD, et al. (2016) Sulfate transporters involved in sulfate secretion in the kidney are localized in the renal proximal tubule II of the elephant fish (Callorhinchus milii). Am J Physiol 311: R66–R78 Google Scholar

20.

Hazon N, Wells A, Pillans RD, Good JP, Anderson WG, Franklin CE (2003) Urea based osmoregulation and endocrine control in elasmobranch fish with special reference to euryhalinity. Comp Biochem Physiol B 136: 685–700 Google Scholar

21.

Hediger MA, Romero MF, Peng JB, Rolfs A, Takanaga H, Bruford EA (2004) The ABCs of solute carriers: physiological, pathological and therapeutic implications of human membrane transport proteins. Eur J Physiol 447: 465–468 Google Scholar

22.

Hentschel H (1987) Renal architecture of the dogfish Scyliorhinus caniculus (Chondrichthyes, Elasmobranchii). Zoomorphology 107: 115–125 Google Scholar

23.

Hentschel H, Mähler S, Herter P, Elger M (1993) Renal tubule of dogfish, Scyliorhinus caniculus: A comprehensive study of structure with emphasis on intramembrane particles and immunoreactivity for H+-K+-adenosine triphosphatase. Anat Rec 235: 511–532 Google Scholar

24.

Hyodo S, Katoh F, Kaneko T, Takei Y (2004) A facilitative urea transporter is localized in the renal collecting tubule of the dogfish Triakis scyllia. J Exp Biol 207: 347–356 Google Scholar

25.

Hyodo S, Kakumura K, Takagi W, Hasegawa K, Yamaguchi Y (2014) Morphological and functional characteristics of the kidney of cartilaginous fishes: with special reference to urea reabsorption. Am J Physiol 307: R1381–R1395 Google Scholar

26.

Imaseki I, Wakabayashi M, Hara Y, Watanabe T, Takabe S, Kakumura K, et al. (2019) Comprehensive analysis of genes contributing to euryhalinity in the bull shark, Carcharhinus leucas; Na+-Cl co-transporter is one of the key renal factors upregulated in acclimation to low-salinity environment. J Exp Biol 222: jeb201780 Google Scholar

27.

Janech MG, Fitzgibbon WR, Ploth DW, Lacy ER, Miller DH (2006) Effect of low environmental salinity on plasma composition and renal function of the Atlantic stingray, a euryhaline elasmobranch. Am J Physiol 291: F770–F780 Google Scholar

28.

Kakumura K, Takabe S, Takagi W, Hasegawa K, Konno N, Bell JD, et al. (2015) Morphological and molecular investigations of the holocephalan elephant fish nephron: the existence of a countercurrent-like configuration and two separate diluting segments in the distal tubule. Cell Tissue Res 362: 677–688 Google Scholar

29.

Kanai Y, Hediger MA (2004) The glutamate/neutral amino acid transporter family SLC1: molecular, physiological and pharmacological aspects. Eur J Physiol 447: 469–479 Google Scholar

30.

Kempton R (1953) Studies on the elasmobranch kidney-II. Reabsorption of urea by smooth dogfish, Mustelus canis. Biol Bull 104: 45–56 Google Scholar

31.

Kiani AK, Paolacci S, Calogero AE, Cannarella R, Di Renzo GC, Gerli S, et al. (2021) From Myo-inositol to D-chiro-inositol molecular pathways. Eur Rev Med Pharmacol Sci 25: 2390–2402 Google Scholar

32.

Kottgen A, Raffler J, Sekula P, Kastenmuller G (2018) Genome-wide association studies of metabolite concentrations (mGWAS): relevance for nephrology. Semin Nephrol 38: 151–174 Google Scholar

33.

Lacy ER, Reale E (1985a) The elasmobranch kidney. I. Gross anatomy and general distribution of the nephrons. Anat Embryol 173: 23–24 Google Scholar

34.

Lacy ER, Reale E (1985b) The elasmobranch kidney. II. Sequence and structure of the nephrons. Anat Embryol 173: 163–186 Google Scholar

35.

Liao Y, Smyth GK, Shi W (2014) FeatureCounts: an efficient general purpose program for assigning sequence reads to genomic features. Bioinformatics 30: 923–930 Google Scholar

36.

Lykidis D, Van Noorden S, Armstrong A, Spencer-Dene B, Li J, Zhuang Z, et al. (2007) Novel zinc-based fixative for high quality DNA, RNA and protein analysis. Nucleic Acids Res 35: e85 Google Scholar

37.

Madsen SS, Bollinger RJ, Brauckhoff M, Engelund MB (2020) Gene expression profiling of proximal and distal renal tubules in Atlantic salmon (Salmo salar) acclimated to fresh water and seawater. Am J Physiol Renal Physiol 319: F380–F393 Google Scholar

38.

Markovich D, Murer H (2004) The SLC13 gene family of sodium sulphate/carboxylate cotransporters. Eur J Physiol 447: 594–602 Google Scholar

39.

Marsenic O (2009) Glucose control by the kidney: an emerging target in diabetes. Am J Kid Dis 53: 875–883 Google Scholar

40.

Masuda N, Ohnishi T, Kawamoto S, Monden M, Okubo K (1999) Analysis of chemical modification of RNA from formalin-fixed samples and optimization of molecular biology applications for such samples. Nucleic Acids Res 27: 4436–4443 Google Scholar

41.

Murray GI (2007) An overview of laser microdissection technologies. Acta Histochem 109: 171–176 Google Scholar

42.

Pang PK, Griffith RW, Atz JW (1977) Osmoregulation in elasmobranchs. Am Zool 17: 365–377 Google Scholar

43.

Park YN, Abe K, Li H, Hsuih T, Thung SN, Zhang DY (1996) Detection of hepatitis C virus RNA using ligation-dependent polymerase chain reaction in formalin-fixed, paraffin-embedded liver tissues. Am J Pathol 149: 1485–1491 Google Scholar

44.

Payan P, Goldstein L, Forster RP (1973) Gills and kidneys in ureosmotic regulation in euryhaline skates. Am J Physiol 224: 367–372 Google Scholar

45.

Plata C, Sussman CR, Sindic A, Liang JO, Mount DB, Josephs ZM, et al. (2007) Zebrafish Slc5a12 encodes an electroneutral sodium monocarboxylate transporter (SMCTn). J Biol Chem 282: 11996–12009 Google Scholar

46.

Puchtler H, Waldrop FS, Meloan SN, Terry MS, Conner HM (1970) Methacarn (methanol-Carnoy) fixation. Histochemie 21: 97–116 Google Scholar

47.

Romero IG, Pai AA, Tung J, Gilad Y (2014) RNA-seq: impact of RNA degradation on transcript quantification. BMC Biol 12: 42 Google Scholar

48.

Sasagawa Y, Nikaido I, Hayashi T, Danno H, Uno KD, Imai T, Ueda HR (2013) Quartz-Seq: a highly reproducible and sensitive single-cell RNA sequencing method, reveals non-genetic gene-expression heterogeneity. Genome Biol 14: R31 Google Scholar

49.

Schlatter P, Gutmann H, Drewe J (2006) Primary porcine proximal tubular cells as a model for transepithelial drug transport in human kidney. Eur J Pharm Sci 28: 141–154 Google Scholar

50.

Schroeder A, Mueller O, Stocker S, Salowsky R, Leiber M, Gassmann M, et al. (2006) The RIN: an RNA integrity number for assigning integrity values to RNA measurements. BMC Mol Biol 7: 3 Google Scholar

51.

Smith HW (1936) The retention and physiological role of urea in the elasmobranchii. Biol Rev 11: 49–82 Google Scholar

52.

Takabe S, Teranishi K, Takaki S, Kusakabe M, Hirose S, Kaneko T, Hyodo S (2012) Morphological and functional characterization of a novel Na + /K + -ATPase-immunoreactive, follicle-like structure on the gill septum of Japanese banded houndshark, Triakis scyllium. Cell Tissue Res 348: 141–153 Google Scholar

53.

Takagi W, Kajimura M, Tanaka H, Hasegawa K, Ogawa S, Hyodo S (2017) Distributional shift of urea production site from the extra-embryonic yolk sac membrane to the embryonic liver during the development of cloudy catshark (Scyliorhinus torazame). Comp Biochem Physiol A 211: 7–16 Google Scholar

54.

Takvam M, Wood CM, Kryvi H, Nilsen TO (2021) Ion transporters and osmoregulation in the kidney of teleost fishes as a function of salinity. Front Physiol 12: 664588 Google Scholar

55.

Vincek V, Nassiri M, Nadji M, Morales AR (2003) A tissue fixative that protects macromolecules (DNA, RNA, and protein) and histomorphology in clinical samples. Lab Invest 83: 1427–1435 Google Scholar

56.

Yanase H, Takebe K, Nio-Kobayashi J, Takahashi-Iwanaga H, Iwanaga T (2008) Cellular expression of a sodium-dependent monocarboxylate transporter (Slc5a8) and the MCT family in the mouse kidney. Histochem Cell Biol 130: 957–966 Google Scholar

57.

Zammit VA, Newsholme EA (1979) Activities of enzymes of fat and ketone-body metabolism and effects of starvation on blood concentrations of glucose and fat fuels in teleost and elasmobranch fish. Biochem J 184: 313–322 Google Scholar
© 2023 Zoological Society of Japan
Takashi Horie, Wataru Takagi, Naotaka Aburatani, Manabu Yamazaki, Mayu Inokuchi, Masaya Tachizawa, Kataaki Okubo, Ritsuko Ohtani-Kaneko, Kotaro Tokunaga, Marty Kwok-Sing Wong, and Susumu Hyodo "Segment-Dependent Gene Expression Profiling of the Cartilaginous Fish Nephron Using Laser Microdissection for Functional Characterization of Nephron at Segment Levels," Zoological Science 40(2), 91-104, (18 January 2023). https://doi.org/10.2108/zs220092
Received: 18 October 2022; Accepted: 21 November 2022; Published: 18 January 2023
KEYWORDS
cartilaginous fish
gene expression profiling
kidney
laser microdissection
membrane transporter
RNA-Sequencing
Back to Top