triadins modulate intracellular ca2+ homeostasis but are not essential for excitation-contraction...

16
Triadins Modulate Intracellular Ca 2 Homeostasis but Are Not Essential for Excitation-Contraction Coupling in Skeletal Muscle * S Received for publication, July 11, 2007, and in revised form, September 20, 2007 Published, JBC Papers in Press, November 2, 2007, DOI 10.1074/jbc.M705702200 Xiaohua Shen , Clara Franzini-Armstrong § , Jose R. Lopez , Larry R. Jones , Yvonne M. Kobayashi ¶1 , Ying Wang , W. Glenn L. Kerrick , Anthony H. Caswell**, James D. Potter**, Todd Miller**, Paul D. Allen , and Claudio F. Perez ‡2 From the Department of Anesthesiology, Brigham and Women’s Hospital, Boston, Massachusetts 02115, the § Deptartment of Cell and Developmental Biology, University of Pennsylvania, Philadelphia, Pennsylvania 19104, the Krannert Institute of Cardiology, Department of Medicine, Indiana University School of Medicine, Indianapolis, Indiana 46202, and the Departments of Physiology and Biophysics and **Molecular and Cellular Pharmacology, University of Miami Miller School of Medicine, Miami, Florida 33101 To unmask the role of triadin in skeletal muscle we engi- neered pan-triadin-null mice by removing the first exon of the triadin gene. This resulted in a total lack of triadin expression in both skeletal and cardiac muscle. Triadin knockout was not embryonic or birth-lethal, and null mice presented no obvious functional phenotype. Western blot analysis of sarcoplasmic reticulum (SR) proteins in skeletal muscle showed that the absence of triadin expression was associated with down-regula- tion of Junctophilin-1, junctin, and calsequestrin but resulted in no obvious contractile dysfunction. Ca 2 imaging studies in null lumbricalis muscles and myotubes showed that the lack of tria- din did not prevent skeletal excitation-contraction coupling but reduced the amplitude of their Ca 2 transients. Additionally, null myotubes and adult fibers had significantly increased myo- plasmic resting free Ca 2 .[ 3 H]Ryanodine binding studies of skeletal muscle SR vesicles detected no differences in Ca 2 acti- vation or Ca 2 and Mg 2 inhibition between wild-type and tria- din-null animals. Subtle ultrastructural changes, evidenced by the appearance of longitudinally oriented triads and the pres- ence of calsequestrin in the sacs of the longitudinal SR, were present in fast but not slow twitch-null muscles. Overall, our data support an indirect role for triadin in regulating myoplas- mic Ca 2 homeostasis and organizing the molecular complex of the triad but not in regulating skeletal-type excitation-contrac- tion coupling. In skeletal muscle, depolarization-initiated calcium release, and the subsequent muscle contraction, depends on the direct interaction of the slow voltage-gated L-type Ca 2 channel (dihydropyridine receptor (DHPR) 3 ) and the sarcoplasmic reticulum Ca 2 release channel (ryanodine receptor, RyR1) in the triad junction. However, an increasing number of other proteins associated with RyRs have been identified in the triadic region and have emerged as important regulators of the Ca 2 release mediated by these two channels (for review see Refs. 1– 4). These proteins, including calsequestrin (Csq), calmodu- lin, triadin, junctin, Junctophilins 1 and 2, MG-29, FKBP12, and others yet to be discovered, along with the RyR and DHPR make up the so-called calcium release units (CRUs) (5). Triadins, a multimember family of proteins that are the prod- uct of alternative splicing from a single gene and expressed almost exclusively in striated muscle (3, 6), have generated sig- nificant attention in recent years for their involvement in a vari- ety of cellular events in muscle cells, but their precise role in muscle function is mostly unknown. Triadin was first identified in skeletal muscle as a 94- to 95-kDa transmembrane protein (7, 8) that is abundantly expressed on the junctional sarcoplasmic reticulum (jSR), were it colocalizes with RyR1 and DHPR (9). Early studies of binding assays of solubilized SR proteins showed that triadin could not only be coimmunoprecipitated with other triadic proteins (10) but also could associate into macromolecular complexes with both the DHPR and RyR1 (7, 11, 12). Based in this association triadin was proposed as the key molecular linker mediating the DHPR/RyR1 communication during muscle contraction. Although functional interactions between triadin and the DHPR in skeletal muscle have proven difficult to confirm, functional and structural interactions between triadin and RyR1 have been documented by several investigators. In vitro studies have shown that the SR luminal domain of triadin not only interacts with RyR1 but appears to anchor Csq to it, mediating the functional coupling between these two proteins via specific domains (13–17). Several studies have suggested a major role for triadin 95 in modulating RyR channel properties. Both an anti-triadin anti- * This work was supported by American Heart Association Grant 0530250N (to C. F. P.) and National Institute of Health Grant PO1AR47605 (to P. D. A.). The costs of publication of this article were defrayed in part by the pay- ment of page charges. This article must therefore be hereby marked advertisement” in accordance with 18 U.S.C. Section 1734 solely to indi- cate this fact. S The on-line version of this article (available at http://www.jbc.org) contains supplemental Figs. S1–S3. 1 Present address: Dept. of Molecular Physiology and Biophysics, University of Iowa, Iowa City, IA 52242. 2 To whom correspondence should be addressed: Brigham and Women’s Hospital, 75 Francis St., Boston, MA 02115. Tel.: 617-525-6486; Fax: 617- 732-6927; E-mail: [email protected]. 3 The abbreviations used are: DHPR, dihydropyridine receptor; RyR1, ryano- dine receptor 1; Csq, calsequestrin; CRU, calcium release unit; SR, sarco- plasmic reticulum; jSR, junctional SR; EC, excitation contraction; FDB, flexor digitorum brevis; TA, tibialis anterior; Sol, soleus; EDL, extensor digitorum longus; CHAPS, 3-[(3-cholamidopropyl)dimethylammonio]-1-propanesul- fonic acid; SERCA, sarco/endoplasmic reticulum Ca-ATPase. THE JOURNAL OF BIOLOGICAL CHEMISTRY VOL. 282, NO. 52, pp. 37864 –37874, December 28, 2007 © 2007 by The American Society for Biochemistry and Molecular Biology, Inc. Printed in the U.S.A. 37864 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 282 • NUMBER 52 • DECEMBER 28, 2007 by guest on February 10, 2016 http://www.jbc.org/ Downloaded from by guest on February 10, 2016 http://www.jbc.org/ Downloaded from by guest on February 10, 2016 http://www.jbc.org/ Downloaded from by guest on February 10, 2016 http://www.jbc.org/ Downloaded from by guest on February 10, 2016 http://www.jbc.org/ Downloaded from by guest on February 10, 2016 http://www.jbc.org/ Downloaded from

Upload: independent

Post on 22-Nov-2023

0 views

Category:

Documents


0 download

TRANSCRIPT

Triadins Modulate Intracellular Ca2� Homeostasis but AreNot Essential for Excitation-Contraction Coupling inSkeletal Muscle*□S

Received for publication, July 11, 2007, and in revised form, September 20, 2007 Published, JBC Papers in Press, November 2, 2007, DOI 10.1074/jbc.M705702200

Xiaohua Shen‡, Clara Franzini-Armstrong§, Jose R. Lopez‡, Larry R. Jones¶, Yvonne M. Kobayashi¶1, Ying Wang�,W. Glenn L. Kerrick�, Anthony H. Caswell**, James D. Potter**, Todd Miller**, Paul D. Allen‡, and Claudio F. Perez‡2

From the ‡Department of Anesthesiology, Brigham and Women’s Hospital, Boston, Massachusetts 02115, the §Deptartmentof Cell and Developmental Biology, University of Pennsylvania, Philadelphia, Pennsylvania 19104, the ¶Krannert Instituteof Cardiology, Department of Medicine, Indiana University School of Medicine, Indianapolis, Indiana 46202,and the Departments of �Physiology and Biophysics and **Molecular and Cellular Pharmacology,University of Miami Miller School of Medicine, Miami, Florida 33101

To unmask the role of triadin in skeletal muscle we engi-neered pan-triadin-null mice by removing the first exon of thetriadin gene. This resulted in a total lack of triadin expression inboth skeletal and cardiac muscle. Triadin knockout was notembryonic or birth-lethal, and null mice presented no obviousfunctional phenotype. Western blot analysis of sarcoplasmicreticulum (SR) proteins in skeletal muscle showed that theabsence of triadin expression was associated with down-regula-tion of Junctophilin-1, junctin, and calsequestrin but resulted innoobvious contractile dysfunction.Ca2� imaging studies in nulllumbricalis muscles and myotubes showed that the lack of tria-din did not prevent skeletal excitation-contraction coupling butreduced the amplitude of their Ca2� transients. Additionally,null myotubes and adult fibers had significantly increasedmyo-plasmic resting free Ca2�. [3H]Ryanodine binding studies ofskeletal muscle SR vesicles detected no differences in Ca2� acti-vation or Ca2� andMg2� inhibition betweenwild-type and tria-din-null animals. Subtle ultrastructural changes, evidenced bythe appearance of longitudinally oriented triads and the pres-ence of calsequestrin in the sacs of the longitudinal SR, werepresent in fast but not slow twitch-null muscles. Overall, ourdata support an indirect role for triadin in regulating myoplas-mic Ca2� homeostasis and organizing themolecular complex ofthe triad but not in regulating skeletal-type excitation-contrac-tion coupling.

In skeletal muscle, depolarization-initiated calcium release,and the subsequent muscle contraction, depends on the directinteraction of the slow voltage-gated L-type Ca2� channel

(dihydropyridine receptor (DHPR)3) and the sarcoplasmicreticulum Ca2� release channel (ryanodine receptor, RyR1) inthe triad junction. However, an increasing number of otherproteins associatedwithRyRs have been identified in the triadicregion and have emerged as important regulators of the Ca2�

release mediated by these two channels (for review see Refs.1–4). These proteins, including calsequestrin (Csq), calmodu-lin, triadin, junctin, Junctophilins 1 and 2,MG-29, FKBP12, andothers yet to be discovered, alongwith theRyR andDHPRmakeup the so-called calcium release units (CRUs) (5).Triadins, amultimember family of proteins that are the prod-

uct of alternative splicing from a single gene and expressedalmost exclusively in striated muscle (3, 6), have generated sig-nificant attention in recent years for their involvement in a vari-ety of cellular events in muscle cells, but their precise role inmuscle function ismostly unknown. Triadinwas first identifiedin skeletalmuscle as a 94- to 95-kDa transmembrane protein (7,8) that is abundantly expressed on the junctional sarcoplasmicreticulum (jSR), were it colocalizes with RyR1 and DHPR (9).Early studies of binding assays of solubilized SR proteins

showed that triadin could not only be coimmunoprecipitatedwith other triadic proteins (10) but also could associate intomacromolecular complexes with both the DHPR and RyR1 (7,11, 12). Based in this association triadinwas proposed as the keymolecular linker mediating the DHPR/RyR1 communicationduring muscle contraction. Although functional interactionsbetween triadin and the DHPR in skeletal muscle have provendifficult to confirm, functional and structural interactionsbetween triadin and RyR1 have been documented by severalinvestigators. In vitro studies have shown that the SR luminaldomain of triadin not only interacts with RyR1 but appears toanchor Csq to it, mediating the functional coupling betweenthese two proteins via specific domains (13–17).Several studies have suggested a major role for triadin 95 in

modulating RyR channel properties. Both an anti-triadin anti-

* This work was supported by American Heart Association Grant 0530250N(to C. F. P.) and National Institute of Health Grant PO1AR47605 (to P. D. A.).The costs of publication of this article were defrayed in part by the pay-ment of page charges. This article must therefore be hereby marked“advertisement” in accordance with 18 U.S.C. Section 1734 solely to indi-cate this fact.

□S The on-line version of this article (available at http://www.jbc.org) containssupplemental Figs. S1–S3.

1 Present address: Dept. of Molecular Physiology and Biophysics, University ofIowa, Iowa City, IA 52242.

2 To whom correspondence should be addressed: Brigham and Women’sHospital, 75 Francis St., Boston, MA 02115. Tel.: 617-525-6486; Fax: 617-732-6927; E-mail: [email protected].

3 The abbreviations used are: DHPR, dihydropyridine receptor; RyR1, ryano-dine receptor 1; Csq, calsequestrin; CRU, calcium release unit; SR, sarco-plasmic reticulum; jSR, junctional SR; EC, excitation contraction; FDB, flexordigitorum brevis; TA, tibialis anterior; Sol, soleus; EDL, extensor digitorumlongus; CHAPS, 3-[(3-cholamidopropyl)dimethylammonio]-1-propanesul-fonic acid; SERCA, sarco/endoplasmic reticulum Ca-ATPase.

THE JOURNAL OF BIOLOGICAL CHEMISTRY VOL. 282, NO. 52, pp. 37864 –37874, December 28, 2007© 2007 by The American Society for Biochemistry and Molecular Biology, Inc. Printed in the U.S.A.

37864 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 282 • NUMBER 52 • DECEMBER 28, 2007

by guest on February 10, 2016http://w

ww

.jbc.org/D

ownloaded from

by guest on February 10, 2016

http://ww

w.jbc.org/

Dow

nloaded from

by guest on February 10, 2016http://w

ww

.jbc.org/D

ownloaded from

by guest on February 10, 2016

http://ww

w.jbc.org/

Dow

nloaded from

by guest on February 10, 2016http://w

ww

.jbc.org/D

ownloaded from

by guest on February 10, 2016

http://ww

w.jbc.org/

Dow

nloaded from

body (18) and overexpression of triadin (19) have been shownto inhibit calcium release from SR vesicles. In addition, purifiedtriadin 95 has been shown to inhibit [3H]ryanodine binding insolubilized heavy SR and to reduce the probability of aperture ofpurified RyR1s incorporated into the planar lipid bilayers (20).Together these studies suggest a negative regulatory role fortriadin on RyR1 Ca2� regulation and support the hypothesisthat triadin 95 plays an important role in excitation contraction(EC) coupling in skeletal muscle.In addition to the 95-kDa isoform described in rabbit and rat

skeletal muscle, where it has been renamed Trisk 95, a 51-kDatriadin isoform (Trisk 51) was also identified in rat (21) andhuman skeletal muscle (22) and two new 49-kDa and 32-kDaisoforms (Trisk 49 and Trisk 32) have recently been identifiedin rat skeletal muscle (23). Trisk 51, which was originallythought to be an anomalous mobility form of the 95-kDa iso-form, is also recognizable in rabbit and canine skeletal muscle(24, 25). Unlike the two larger isoforms, Trisk 49 and Trisk 32are not located in the CRU but are found on the I-Z-I level SR.All isoforms share common N-terminal domains but have aunique C-terminal regions (3, 21, 22, 25). Both the differentlocations of the smaller isoforms and the different functionaleffects caused by overexpression of triadin 95 kDa and triadin51 kDa (19) suggest specific roles for each of the four isoforms.In an attempt to unmask the role of triadins in skeletal mus-

cle, we engineered a pan-triadin-null mouse that shows nomeasurable expression levels of any known isoforms of triadin.Muscles and myotubes derived from triadin-null mice werethen subjected to an extensive physiologic, biochemical, andultrastructural characterization. Surprisingly, we find thathomozygous triadin-null mice have no obvious functional phe-notype and are able to breed and survive into adulthood. Mostimportantly we detected almost normal EC coupling and noobvious differences in RyR1 channel regulation between nulland wild-typemice.Minor ultrastructural alterations of the tri-adic junction organization were observed in fast but not slowtwitch triadin-null muscle fibers. Overall, our data support acontributory role for triadin 95 and triadin 51 in the structuralorganization of the triad junction and regulation of intracellularCa2� homeostasis but indicate that triadins are not essentialcomponents of skeletal-type EC coupling.

EXPERIMENTAL PROCEDURES

Generation of Triadin-null Mice—A mouse genomic DNAlibrary was screened with a rabbit triadin cDNA fragment toyield a 14-kb DNA fragment containing the complete Exon-1and a partial fragment of Intron-1 of Trdn gene. A targetingconstruct was designed by flanking the entire Exon-1 and itsputative promoter regulatory region with two loxP sites. Thefloxed Exon was then inserted into pTKLNCL vector (kindlyprovided by Dr. Rick Mortensen), downstream of the cytocinedeaminase/neomycin expression cassette (Fig. 1A). The vectorwas electroporated into 129/SvJ ES cells, and the recombinedcells were then microinjected into C57BL/6N blastocysts togenerate chimeric mice. Heterozygous germ line-transmittedmice carrying the floxed allele (Trdn�/flxd) were then bred withEIIa-Cre transgenic mice (Jackson Laboratories, Bar Harbor,ME) to excise the loxP-floxed exon (26). The resulting het-

erozygousTrdn-nullmice, which lack both Exon-1 and theCD-Neo selection cassette, were then interbred to obtain homozy-gous Trdn-null mice. Genotyping was performed by PCR andSouthern blotting analysis using genomic DNA (Fig. 1B).Cell Culture and Ca2� Imaging—Primary myotubes from

triadin-null and wild-type mice were isolated according to themethod of Rando and Blau (27). Themyoblasts were grown anddifferentiated as described previously (28). Calcium imagingwas performed 4–5 days after differentiation, in myotubesloaded with 2 �M Fluo-4 AM (Molecular Probes, OR). The cellswere imaged in imaging buffer (mM: 125 NaCl, 5 KCl, 2 CaCl2,1.2 MgSO4, 6 glucose, and 25 HEPES, pH 7.4) at 490–500 nmusing a DG4multiwavelength light source with a Stanford Pho-tonics 12 bit digital intensified charge-coupled device, and thedata were displayed and analyzed using QED imaging software(QED Software, Pittsburgh PA). Depolarization-initiated Ca2�

release was tested by exposing the cells to increased voltagesteps using electrical field stimulation in imaging buffer with-out CaCl2 and supplemented with 0.5 mM CdCl2 and 0.1 mMLaCl3. Chemical depolarization was achieved by a 10-s expo-sure to highK�buffer (28), supplementedwith orwithout 2mMCa2� plus Cd2� and La3�, using a multivalve perfusion system(AutoMate Scientific, Inc., Oakland, CA).SR Ca2� Loading Content Determination—Relative SR Ca2�

content levels of myotubes and flexor digitorum brevis (FDB)fibers were estimated from the magnitude of the Ca2� releaseinduced by 40 mM caffeine in Fluo-4AM-loaded cells. Caffeinewas supplementedwith 1�M thapsigargin and 0.5mMCd2�/0.1mM La3� to block the SERCA pump and avoid Ca2� entry fromthe extracellular medium, respectively. Single muscle fibersfrom FDB muscle were enzymatically dissociated with colla-genase as described previously (29) and were imaged at 1 fps inthe presence of 100 �M n-benzyl-p-toluenesulfonamide to pre-vent muscle contraction. Total SR calcium content wasexpressed as the area under the curve of the Ca2� release tran-sient induced by 70- to 90-s exposure to the caffeine mixture.Resting Free Ca2� Measurements—Determination of myo-

plasmic resting free Ca2� in myotubes and intact muscles wasperformed with double-barreled Ca2�-selective microelec-trodes assembled with ETH129 resin as described previously(28, 30).Intact Adult Fiber Studies—Intact lumbricalis muscles (1.0–

2.0� � 0.2–0.3 mm) from the hind foot were dissected in ice-cold Krebs-Henseleit solution (mM: 119 NaCl, 4.6 KCl, 1.8CaCl2, 1.2 MgSO4, 25 NaHCO3, 1.2 KH2PO4, and 11 glucose)containing 30 mM 2,3-butanedione monoxime (Sigma, St.Louis, MO), mounted in the Guth muscle research apparatusand loaded with 5 �M Fura 2-AM (Molecular Probes, Eugene,OR) in an oxygenated Krebs-Henseleit solution containing0.5% cremophore for 1 h, as described previously (31). Fura-2fluorescence, expressed as a 340 nm/380 nm excitation ratio,was collected with a 4-ms resolution timeframe.Membrane Vesicle Preparation and Immunoblotting—Mi-

crosomal vesicles were prepared from lower limb and individ-ual muscles (tibialis anterior (TA), soleus (Sol), and extensordigitorum longus (EDL)) of wild-type and triadin-null mice,6–8 weeks old. Muscles were homogenated in a Polytron celldisrupter in 5 mM imidazole, pH 7.4, 300 mM sucrose supple-

Triadin Knockout and Skeletal Muscle Function

DECEMBER 28, 2007 • VOLUME 282 • NUMBER 52 JOURNAL OF BIOLOGICAL CHEMISTRY 37865

by guest on February 10, 2016http://w

ww

.jbc.org/D

ownloaded from

mented with protease inhibitor (CompleteTM, Roche AppliedScience) and collected as described elsewhere (28). Proteinswere separated in SDS-polyacrylamide gel electrophoresis (32)and transferred to polyvinylidene difluoride membrane.Expression of specific proteins was assessed by incubation ofimmunoblots with poly- or monoclonal antibodies against;RyR1 (34C, Sigma), skeletal DHPR (MA3–920), SERCA (MA3–911), calsequestrin (MA3–913), triadin (MA3–927, fromAffin-ity BioReagents,Golden,CO), junctin (1E6), Junctophilin-1 and-2 (kindly provided by Dr. H. Takeshima), and glyceraldehyde-3-phosphate dehydrogenase (FL-335, Santa Cruz Biotechnol-ogy, Santa Cruz, CA). Membranes were then incubated witheither anti-mouse or anti-rabbit horseradish peroxidase-con-jugated goat, secondary antibody and developed with Super-Signal ultra chemiluminescent substrate (Pierce). Triadinexpression was also confirmed by using PA-TRN6 antibody,which was raised to residues 146–160 in the luminal domaincommon to all skeletal muscle and cardiac triadins. Immu-noblotting of transferred proteins was conducted using 125I-Protein A for visualization of antibody-binding proteins asreported previously (25, 33).[3H]Ryanodine Binding Assay—[3H]Ryanodine binding to

crude membrane extracts (25–30 �g/ml) was performed atequilibrium for 90min at 37 °C in 20mMHEPES, pH 7.4, 0.25 M

KCl, and 5 nM [3H]Ryanodine (PerkinElmer Life Sciences) asdescribed previously (28).

Immunolabeling—Sternomastoidmuscles were fixed in situ in 1%p-formaldehyde prepared in phos-phate-buffered saline. Small bun-dles of cells were teased out, washedand then blocked in phosphate-buffered saline/bovine serum albu-min containing 10% goat serum and0.5% Triton X-100 for 1 h, at roomtemperature. Cells were then incu-bated in the primary antibody over-night at 4 °C, and then for 1–2 hwith the secondary antibody atroom temperature. The primaryantibodies used were: anti-CSQ(VIIID12, Affinity BioReagents),anti-triadin (GE 4.90), anti-junctin(1E6), and anti-RyR (34C). Second-ary antibodies were Cy3-conjugatedgoat anti-mouse and Texas red-conjugated goat anti-rabbit (MPBiomedical, Solon, OH). Imageswere obtained in a confocal ZeissLSM 510 microscope (Carl Zeiss,Switzerland).Electron Microscopy—Sternomas-

toid, EDL, and soleus muscles werefixed with 3.5% glutaraldehyde in0.1 M sodium cacodylate buffer, pH7.2. Muscles were post-fixed in 2%OsO4 in 0.1 M sodium cacodylate for1 h at room temperature, stained en

bloc with saturated uranyl acetate in 70% ethanol, andembedded in Epon. Ultrathin sections (70–90 nm) werestained with saturated uranyl acetate solution in 50% ethanoland Sato lead solution (34). The frequency of triads in differ-ent orientations was counted in images from longitudinalsections covering a 36-�m2 area. Two or three images weretaken for each fiber.Data Analysis—Unless indicated otherwise statistical signif-

icant differences among the datawere evaluated usingOne-wayTukey’s analysis of variance (nonparametric) analysis and Stu-dent’s t test (GraphPad Software, San Diego, Ca). Data are pre-sented as mean � S.D. or means � S.E.

RESULTS

Generation of Triadin-null Mice—For the 5� sequence thetriadin 95 cDNAwas used as probe to isolate the 5� fragment ofthe triadin gene (Trdn) in a 129svJ genomic DNA library andused to create a targeting construct. Because of the suspectedkey role of triadin in EC coupling, a conditional knockout strat-egy was attempted. The targeting vector was designed to con-ditionally delete Exon-1, and its putative 5�-untranslatedregion, in a tissue-specificmanner by “floxing” the exon and theNeo/CD selection cassette with three loxP sites (Fig. 1A). EScells carrying the appropriately targeted allele were injectedinto C57BL6 blastocysts and introduced into pseudo pregnantfemale mice. Germ line chimeric mice carrying the floxed alle-

FIGURE 1. Targeted disruption of Exon-1 of mouse triadin gene. A, restriction map of mouse triadin geneshowing the wild-type allele (top), targeting vector (first middle), recombinant allele before Cre recombination(second middle), and recombinant null allele after Cre recombination (bottom). The Exon-1 (filled boxes), cyto-neo resistance cassette (arrow), and loxP sites (open boxes) are indicated. Dashed lines indicate the backbone ofthe vector. Localization of hybridization probes is indicated by open arrows. B, BamHI; N, NdeI; P, PvuII; P*,deleted PvuII site. B, Southern blot analysis of mouse tail DNA from floxed littermates. Digestion with NdeI andhybridization with probe P1 (left panel) generated a 6.5- and a 4.3-kb fragment, for the wild-type and mutantallele, respectively. Digestion with BamHI and hybridization with probe P2 (middle panel) generated a 8.2-kband a 6.0-kb fragment, for the wild-type and mutant allele, respectively. Tail DNA from hetero- and homozy-gous triadin-null mice (right panel) was digested with PvuII/BamHI and then hybridized with probe P3. Thewild-type allele is 5.5 kb, and the null allele is 6.7 kb. C, Western blot analysis of triadin expression in skeletalmuscles. 40 �g/lane of microsomal vesicles from wild-type (�/�) and triadin hetero (�/�), and homozygous(�/�) lower limb muscles were analyzed with triadin-specific antibodies GE 4.90. D, triadin expression patternof cardiac (lanes a and b) and skeletal muscle (lanes c and d) from wild-type (�/�) and triadin-null (�/�) micewas compared with canine cardiac (lane e) and skeletal (lane f) muscle using the site-specific PA-TRN6 antibody.E, profile of triadin expression in fast and slow twitch muscle was analyzed in microsomal fractions (40 �g/lane)of EDL, TA, and Sol muscles from wild-type (�/�) and triadin-null (�/�) using PA-TRN6 antibody.

Triadin Knockout and Skeletal Muscle Function

37866 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 282 • NUMBER 52 • DECEMBER 28, 2007

by guest on February 10, 2016http://w

ww

.jbc.org/D

ownloaded from

les were backcrossed with C57BL6 EIIA cre mice in an attemptto either remove the Neo/CD selection cassette (to generateconditional knockouts) or to remove both Exon-1 and theNeo/CD selection cassette (to generate traditional knockouts).Progeny were screened using PCR (data not shown) and South-ern blot (Fig. 1B). Despite several attempts it was never possibleto obtain mice with the conditional allele. Heterozygous micecarrying the traditional knockout allele were then backcrossed,and surprisingly the homozygous mice (Trdn�/�) did notexhibit embryonic or birth lethality nor did they demonstrateany obvious gross functional phenotype.Triadin-null Muscles Lack Expression of All CRU-related

Triadin Isoforms—Analyses for expression of triadin isoformswere performed by immunoblot in crude membrane extractsfrom lower limb muscle using two different anti-triadin anti-bodies. Fig. 1C compares the triadin expression pattern of wild-type, heterozygous, and homozygous null muscles detectedwith MA3–297 antibody, specific for the 95-kDa isoform oftriadin. As expected, skeletal muscle from heterozygous nullmice (�/�) expressed significant amounts of the 95-kDa tria-din isoform, although, to a lesser extent than wild-type (�/�)muscle. However, no detectable expression levels of this iso-form were observed in membrane extracts from homozygousnullmuscle (�/�). The expression of the other triad-associatedtriadin isoform was further investigated by using a site-specificantibody (PA-TRN6) raised against residues 146–160 in theluminal domain of mouse cardiac and skeletal muscle triadins(25). In microsomal fractions from wild-type mouse skeletalmuscle (Fig. 1D, lane c) and from canine skeletal muscle (Fig.1D, lane f), this antibody detected two prominent bands of�60kDa and 95 kDa, corresponding to the higher molecular weightisoforms of triadin. No detectable levels of either isoform wereevident in fractions from triadin-null mice (Fig. 1D, lane d). Inaddition this antibody did not detect either of the two lowermolecular weight forms of triadin (35–40 kDa) that are primar-ily detected in cardiac ventricle (Fig. 1D, lane e). These twobands, although of different intensities, were readily observedin cardiac ventricular microsomal preparations from wild-typemice (Fig. 1D, lane a) but were absent in similar preparationfrom triadin-null animals (Fig. 1D, lane b), thus confirming thattriadin-null skeletal and cardiac muscles lack any detectableexpression levels of triadins.Triadin expression was further separately tested in fast

twitch (EDL and TA) and slow twitch (Sol) muscle. Immuno-blot analysis ofmicrosomal fractions fromwild-typemice showthat eachmuscle type had a slightly different triadin expressionpattern: whereas EDL and Sol muscles primarily expressed the60-kDa isoform and, to a lesser extent, the 95-kDa isoform oftriadin, TA muscles express only the 60-kDa isoform (Fig. 1E).Unlike their wild-type counterparts triadin-null musclesshowed no detectable expression levels of either triadin iso-form. However, a weak band in the 75- to 80-kDa range wasevident in all muscle types tested (Fig. 1, D and E). This banddoes not correlate with any known isoform of triadin and isassumed to be the result of nonspecific antibody cross-reactiv-ity, because it was present in both wild-type and null mice.

Expression of Other Triadic Proteins in Triadin-null SkeletalMuscle—To address whether or not expression of triadinaffected the expression pattern of other CRU proteins that areinvolved in the organization of the CRU, expression levels ofRyR1, DHPR �1S, Csq, junctin, JP-1 and JP-2 were compared inwild-type and triadin-null EDL and soleusmuscles (Fig. 2A). Toquantify differences a densitometric analysis of the immuno-blotswas performed, and banddensitieswere normalized to theexpression of glyceraldehyde-3-phosphate dehydrogenase tocorrect for possible loading differences. Densitometric datashow no significant differences in expression levels of DHPR�1S subunit or JP-2 between triadin-null and wild-type EDL orSol muscles (Fig. 2, A and B, mean � S.D., p � 0.05). However,a significant reduction in expression of JP-1 and Csq wasdetected in both EDL and Sol fibers in triadin-null mice.Expression of JP-1 was reduced to 76 � 7% and 28 � 4% ofwild-type EDL and Sol fibers, respectively, andCsqwas reducedto 76� 2% and 62� 15% of the wild-type EDL and Solmuscles,respectively (Fig. 2B). Expression of junctin and RyR1 was dif-ferentially affected in Sol and EDL muscles. Junctin was signif-icantly down-regulated in EDL muscles (72 � 2% of wild-typeexpression) but not in Sol fibers (99 � 4% of wild-type expres-sion, Fig. 2B), and RyR1 expression showed a slight but signifi-cant up-regulation in Sol (133 � 8% of wild-type expression,Fig. 2B) but not in EDL (114 � 4% of wild-type expression, Fig.2B). Interestingly, unlike the adult muscles we detected no sig-nificant differences (p � 0.05) in Csq or junctin expressionbetween wild-type and triadin-null myotubes (Fig. 2, A and B).Immunolabeling for calsequestrin, junctin, andRyR1 showed

no detectable changes in the position of these molecules in fasttwitch (sternomastoid) muscles of null versus wild-type mice(Fig. 2C). In all cases these proteins were located at spots cor-responding to the location of the triads in two transverse bandson either side of the Z lines (note: a positional shift of Csq,affecting only a small percentage of themolecules aswas observedby electron microscopy (see below), would not be detectable inthese confocal images). Unfortunately anti-triadin antibody GE4.90, which was used forWestern blots, gave nonspecific stainingwhen used on p-formaldehyde fixed tissue in our attempts at tria-din immunolabeling (data not shown).Effect of the Absence of Triadin on the Organization of Junc-

tional and Longitudinal SR—To evaluate the role of triadin onthe subcellular architecture of skeletal muscle we analyzed theultrastructure of junctional and free SR in wild-type and tria-din-null fibers from fast twitch (sternomastoid, EDL) and slowtwitch (Sol) muscles. The muscle ultrastructure from the nullmice was basically unaltered in regards to the disposition andordering of themyofibrils and cross-striation but showed subtleultrastructural alterations in the triads and lateral SR that reveala role for triadin both in retaining Csq within the jSR domainsand, perhaps indirectly, in the arrangement of triads within thefiber.Two structural alterations were observed in triadin-null

muscles. One is a disorder in the orientation of the triads.Although located opposite to the edges of the A band as in wildtype, some of the triads in null muscles lost their predominanttransverse orientation, which is more or less perpendicular tothe long axis of the fibers (Fig. 3A, asterisks) and acquired an

Triadin Knockout and Skeletal Muscle Function

DECEMBER 28, 2007 • VOLUME 282 • NUMBER 52 JOURNAL OF BIOLOGICAL CHEMISTRY 37867

by guest on February 10, 2016http://w

ww

.jbc.org/D

ownloaded from

oblique or longitudinal orientation (Fig. 3A, double arrows).Groups of transversely and longitudinally oriented triads occu-pied adjacent domains within the same fibers, as shown in Fig.3B (transverse triads, asterisks; longitudinal triads, doublearrows). The frequency of longitudinally oriented triads variedin different muscles, and even in different areas of individualfibers, and was affected by the absence of triadin. In wild-typeEDL and sternomastoid muscles, triads are almost exclusivelytransverse, although a small percentage of the areas countedshowed a low frequency (5–15%) of longitudinal triads (Fig. 3and supplemental Fig. S1). Wild-type soleus had no longitudi-nal triads. Triadin-null EDL and sternomastoid fibers show anincreased frequency of areas with only a moderate frequency oftriads (5–20%), and more than a third of the areas counted hada higher percentage of longitudinal triads (20 and 70%) thanwild-type controls. In triadin-null soleus, on the other hand,longitudinal triads were present in very few areas and only atlow frequency (Fig. 3B and supplemental Fig. S1). Average val-ues and statistical analysis are shown in Table 1.Triads of muscles from adult null and wild-type mice,

whether longitudinally or transversely oriented, were undistin-guishable from each other in regard to the structure of Csq. Inall triads Csq appeared as a finely granular tightly aggregatedmeshworkwithin the terminal cisternae of the SR (Fig. 3). How-

FIGURE 2. Levels of triadic protein in homogenates from fast and slow twitch skeletal muscle. A, identical mounts of microsomal fraction of fast twitch(EDL) and slow twitch (Sol) muscle and myotubes from wild-type and triadin-null mice were loaded and immunoblotted for several antibody. Blots were probedwith antibodies specific for ryanodine receptor (RyR1), DHPRs, Csq, Junctophilin-1 (JP-1), Junctophilin-2 (JP-2), and junctin, as described under “ExperimentalProcedures.” Myotube preparations were tested with antibodies against Csq and anti junctin only. B, normalized band intensity of triadic proteins expressedin EDL and Sol muscles (upper panel) and primary myotubes (lower panel) from triadin-null mice. The band intensity for each protein was the normalized by itscorresponding wild-type counterpart (dotted line). The numbers in the bars indicate the number of blots analyzed per protein. Data presented as mean � S.D.*, p � 0.05; **, p � 0.01. C, immunohistochemical detection of triadic proteins in skeletal muscles from wild-type (�/�) and triadin-null (�/�) mice. Sterno-mastoid muscles were fixed in p-formaldehyde and stained with anti-Csq, anti-junctin (Jct), and anti-RyR antibodies.

FIGURE 3. SR ultrastructure in fast twitch fibers from the sternomastoid oftriadin-null (A–E) and wild-type (B, D, and insets) muscles. Longitudinally ori-ented triads (double arrows) as well as the normal transversely oriented triads(asterisks) are seen in the images. Enlarged longitudinal SR containing mislocal-ized Csq (arrowheads) are illustrated in transversal (B) and longitudinal (C–E) thinsections from triadin-null muscles. Insets inBandDshowSRfromwild-typemuscleswith an apparently empty lumen. The longitudinal SR from triadin-null muscle (C)shows a normal structure at the left of the Z line (Z), but is filled with Csq on the right.

Triadin Knockout and Skeletal Muscle Function

37868 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 282 • NUMBER 52 • DECEMBER 28, 2007

by guest on February 10, 2016http://w

ww

.jbc.org/D

ownloaded from

ever, a detectable structural alteration of null muscles was theoccasional presence of Csq in parts of the SR that are usuallydevoid of the protein. The misplaced Csq protein was found intubes of the longitudinal SR that were dilated and located eitherin close proximity to the junctional SR, or at some distancefrom it, opposite the middle of either the A band or the I-Z-Iregion (Fig. 3, B–E, arrowheads). The randomness of this event isbest shown inFig. 3C, inwhich theSR tubes on the leftof theZ linehad the normal narrow diameter and an apparently empty lumen,whereas those on the right, even though continuous with the for-mer, had a wider diameter and a visible granular content. Easilyseen examples of displaced Csq are not frequent, and they wererevealedonly by intensive scrutiny of large areas of sectionedmus-cle. For that reason a quantitative assessment of the changes wasnotpossible. It is likely thatminormovementsofCsq,belowdetec-tion threshold by electron microscopy and also by immunolabel-ing (see above) arewidespread but of smallmagnitude. The soleusmuscle did not show this phenotype.As expected, freeze-fracture analysis of the diaphragm in

nine images from a triadin-null pup at E18 showed DHPRs

grouped into tetrads indicating that in the absence of triadin,RyR1, and the DHPR maintained their appropriate link (sup-plemental Fig. S2). The finding is consistent with the Ca2�

imaging data that support near normal EC-coupling propertiesof the triadin-null muscles.ECCoupling in Triadin-nullMyotubes—The effect of triadin

ablation on depolarization-initiated Ca2� release was tested inFluo4-loaded cultured myotubes from wild-type and triadin-null mice using high speed Ca2� imaging in nominal Ca2�-freesolutions and in Fura-2-loaded adult lumbricalis musclesfreshly dissected fromnull andwild-typemice. The presence orabsence of triadins in these myotubes (Fig. 4A, inset) was con-firmed by immunoblot. Field stimulation of wild-type myo-tubes elicited transient elevations of intracellular calcium in theabsence of extracellular Ca2� (i.e. skeletal-type EC coupling).Subsequent addition of 40 mM caffeine induced an even largerCa2� release from internal stores. The same protocol elicitedsimilar electrically stimulated and caffeine-induced transientsin triadin-null myotubes, but both transients were smaller inamplitude (Fig. 4A) than seen in wild-type myotubes.To probe for possible differences in the sensitivity of wild-

type and triadin-null myotubes to depolarization we comparedthe ability of these cells to respond to increased depolarizationsteps induced by KCl. In addition to effectively clamping themembrane potential, K� depolarization has the advantage ofallowing prolonged plasmalemma depolarization (seconds)without deleterious effects on the integrity of the plasmamem-brane. As shown in Fig. 4B, both cell types respond to K� in aconcentration-dependent manner. However, triadin-null myo-tubes had a small but statistically significant reduction intheir sensitivity to K� depolarization (EC50 15.1 � 0.4 mM

versus 20.4 � 0.7 mM for wild-typeand nulls myotubes, respectively)and had Ca2� transients that weresmaller in amplitude than wild-type myotubes at all KCl concen-trations (p � 0.05, Fig. 4C).Calcium transients induced by 40

mM KCl, in the presence of eitherextracellular Ca2� or calcium chan-nel blockers, are compared in Fig. 5(A and B). Triadin-null myotubeshad Ca2� transients that were onaverage 30% smaller in magnitudethan those observed in wild-typecells. Normalization of the imagingdata to the peak Ca2� transient ineach group shows that Ca2� releaseand uptake kinetics were nearlyidentical (Fig. 5,C andD) both in thepresence and in the absence ofextracellular Ca2�.ForceGeneration Studies in Intact

Muscles—The contractile proper-ties of skeletal muscles from hetero-and homozygous triadin-null micewere studied in intact lumbricalismuscles. Muscles were mounted in

FIGURE 4. Effect of triadin-null on skeletal EC coupling. A, action potential-induced Ca2� transients in Fluo-4-loaded primary myotubes. Both wild-type and triadin-null cells responded to 1-ms stimuli (1 Hz) and 40 mM

caffeine with Ca2� transients that persisted after blocking currents with 0.5 mM Cd2� and 0.1 mM La3�. Expres-sion of triadin in myotubes was confirmed by immunoblotting with GE 90.1 antibody (inset). B, representativetraces of wild-type and triadin-null myotubes loaded with Fluo-4 and exposed to increased depolarizationsteps with KCl solutions (10 s). C, comparison of average magnitude of the KCl-induced Ca2� transients (areaunder the curve) between wild-type and triadin-null myotubes. Data presented as mean � S.D. (n 48 cells).

TABLE 1Longitudinally oriented triads as percentage of total triads in fastand slow twitch muscles

Wild type No. ofimages

No. oftriads Null No. of

imagesNo. oftriads

Mean � S.D. mean � S.D.Sternomastoid 1.0 � 1.8%a,b 66 2331 18.0 � 15.8%a,b 71 1406EDL 3.4 � 6.2%b 27 517 30.2 � 20.9%b 46 680Soleus 0%c 45 744 1.5 � 3.5%c 35 563a Data were collected from three mice; date for all others were collected from twomice.

b Mann-Whitney test, medians, p � 0.0001.c Medians, p � 0.05.

Triadin Knockout and Skeletal Muscle Function

DECEMBER 28, 2007 • VOLUME 282 • NUMBER 52 JOURNAL OF BIOLOGICAL CHEMISTRY 37869

by guest on February 10, 2016http://w

ww

.jbc.org/D

ownloaded from

Guth muscle research apparatusand loaded with 5 �M Fura-2 forsimultaneousmeasurement of forcegeneration and global Ca2� releasein response to electrical stimulation.As expected, both heterozygous andhomozygous muscle preparationshad contractile responses evoked byelectrical stimulation confirmingour studies in myotubes. Analysis ofthe tension levels developed inresponse to single twitch revealed alarge variation within each musclesgroup, which resulted in no recog-nizable differences in force genera-tion between wild-type and triadin-null muscle. Normalization of thedeveloped force (P/Po) by their peakvalues did not show any differencesin kinetic of isometric force devel-opment between null and wild-typemuscles (Fig. 6, upper). However,measurement of global Ca2� release(Fura-2, 340/380 ratio) in responseto single twitch revealed that tria-din-null muscles had a significant

reduction in the amplitude of their peak Ca2� transient whencompared with wild-type muscles (expressed as fraction ofwild-type Ca2� peak ratio, Fig. 6, lower).Effect of Triadin on SR Ca2� Content—Based in our Ca2�

imaging data, it appeared that both myotubes and adult fibersfrom triadin-null mice have smaller SR Ca2� loading capacitythan their wild-type counterparts. In an attempt to quantify thelevel of SR calcium content we used 40mM caffeine in the pres-ence of 1 �M thapsigargin plus Cd2�and La3� to deplete SRstores in Fluo-4-loaded cells. Calcium release was measured inthe presence of thapsigargin, Cd2�, and La3� to avoid Ca2�

reloading into the SR stores and Ca2� entry from the extracel-lularmedium, respectively. Fig. 7A shows that under these con-ditionsmyotubes had a biphasic response to caffeine with a fastCa2� release component followed by a slower and sustainedrelease event, most likely representing the release of free Ca2�

and Ca2� bound to Csq, respectively. Although, wild-type andtriadin-null cells display similar Ca2� release kinetics, theamplitude of the fast Ca2� release component in triadin-nullmyotubes was significantly decreased compared with wild-typecells. This difference in Ca2� release very likely represents realdifferences in the SR calcium pool as both wild-type and tria-din-null cells had identical sensitivity to caffeine (see Fig. 8). Asa result, the quantification of the total Ca2� released (areaunder the curve) showed that there was modest but significantreduction in total SR calcium content in triadin-null cells (15%,Fig. 7B). FDB fibers displayed a similar Ca2� release pattern asmyotubes butwith a less prominent fast component and amorerobust slow release phase (Fig. 7C). Unlike myotubes, however,the total Ca2� released in triadin-null FDBs was much smaller(40% reduction in SR Ca2� content) than wild-type fibers (Fig.7D). The larger reduction in Ca2� loading capacity observed in

FIGURE 5. Comparison of the magnitude of K�-induced Ca2� transients of wild-type and triadin-nullmyotubes. Average magnitude of Ca2� transients induced by 40 mM KCl in wild-type (black) and triadin-null(gray) myotubes. Transients were measured in the presence of 2 mM extracellular calcium (left panels) and in theabsence of calcium plus Cd2� and La3� (right panels). Records were normalized by their peak Ca2� amplitude(bottom panels) to allow for comparison of their Ca2� transient kinetics. Data are presented as mean � S.E. Thehorizontal scale bar indicates 10 s.

FIGURE 6. Force generation and Ca2� release from skeletal muscle. Lum-bricalis muscles from wild-type, heterozygous (�/�), and homozygous(�/�) triadin-null mice were stretched in a Guth muscle apparatus untilthe maximum active twitch force was obtained and were allowed to equil-ibrate and develop a stable force for one hour before study. Upper, aver-age normalized force and lower, Ca2� transients amplitude, normalized bywild-type peak amplitude, in response to 1.0 Hz stimulation. Data arepresented as mean � S.E. (n 4).

Triadin Knockout and Skeletal Muscle Function

37870 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 282 • NUMBER 52 • DECEMBER 28, 2007

by guest on February 10, 2016http://w

ww

.jbc.org/D

ownloaded from

triadin-null FDBs seems consistent with the sizable reductionof Csq expression observed in triadin-null adult muscle, whichwas not seen in myotubes (see Fig. 2, A and B).

Resting Free Ca2� Concentrationin Triadin-null Myotubes—Be-cause of the alleged role of Csq andintraluminal Ca2� levels in modu-lating RyR1 activity and, therefore,Ca2� homeostasis, it is not unlikelythat the reduction inCsq expressionobserved in adult fibers but not inmyotubes of triadin-null mice couldaffect Ca2� regulation in myotubesand muscle fibers differently. Toassess this possibility resting freeCa2� concentration was measuredin wild-type and triadin-null pri-mary myotubes and TA musclesusing calcium selective microelec-trodes. As shown in Table 2, myo-plasmic resting free Ca2� in wild-type myotubes was 111 � 2 nM,similar to the resting Ca2� concen-tration previously reported in adultskeletal muscle fibers and RyR1-knockout myotubes expressingrecombinant RyR1 (28). By compar-ison, primary myotubes from tria-din-null mice had a substantialincrease (182 � 3 nM) in myoplas-mic resting free Ca2�. Likewise, insitu determinations of myoplasmicCa2� concentration in intact TAmuscle from wild-type and triadin-null mice showed similar differ-ences in resting Ca2� concentra-tions as those seen in myotubepreparations (see Table 2). How-ever, triadin-null TA muscles dis-played a seemingly lower restingcalcium concentration than nullmyotubes. These results are consist-ent with our imaging data showingthat triadin-null muscles havesignificantly less SR Ca2� loadingcapacity than myotubes and sug-gest that triadins may play a rolein regulating myoplasmic Ca2�

homeostasis.Effects of Triadin Expression on RyR1 Ca2� Release Channel

Function—Triadin/RyR1 interaction has been suggested tomodulate RyR1 channels either directly or indirectly (20, 35).To evaluate the impact of the absence of triadin on RyR1 activ-ity, [3H]ryanodine binding was measured in crude SR mem-brane preparations in the presence of RyR1 agonists (caffeineand Ca2�) and inhibitors (Ca2� and Mg2�). Fig. 8A shows thatmembrane preparations from both wild-type and triadin-nulllimbmuscles had a classical bell-shaped dose response to Ca2�,in which sub-millimolar concentrations of Ca2� increase RyR1channel activity and millimolar Ca2� concentrations inhibit it.Average EC50 values, calculated from the Ca2� activation curve

FIGURE 7. Comparison of SR calcium loading content in myotubes and adult muscle fibers. SR calciumstores of Fluo-4 loaded primary myotubes (A) and isolated FDB fibers (C) were depleted with 40 mM caffeine inthe presence of 1 �M thapsigargin, 0.5 mM Cd2�, 0.1 mM La3�, and nominally free (�7 �M) extracellular Ca2�.Total SR Ca2� content was estimated from the amplitude of the cytosolic calcium transients (area under thecurve, B and D). Data are presented as mean � S.E. of 36 –50 FDB fibers and 126 –149 myotubes. *, p � 0.05;***, p � 0.001.

FIGURE 8. [3H]Ryanodine binding to wild-type and triadin-null skeletal muscle homogenates. Ca2�

dependence (A), caffeine activation (B), and Mg2� inhibition curves (C) of specific [3H]ryanodine binding tomicrosomal vesicles from wild-type (E) and triadin-null (F) mice skeletal muscles. Values are depicted asdose-response curves with each point representing mean � S.D. for 5–10 measurements. EC50 and log IC50values were determined as described under “Experimental Procedures.” Statistical analyses using analysis ofvariance yielded no significant differences for any condition tested.

TABLE 2Resting free Ca2� concentration in myotubes and TA muscles fromwild type and Triadin-null mice

Wild type Null No. of determinationsnM (mean � S.E.) nM (mean � S.E.) Wild type/null

Myotubes 111 � 2 182 � 3a 26/55TA in situ 123 � 2 157 � 5a 9/9

a One-way analysis of variance (Tukey analysis), p � 0.001 in comparison to wild type.

Triadin Knockout and Skeletal Muscle Function

DECEMBER 28, 2007 • VOLUME 282 • NUMBER 52 JOURNAL OF BIOLOGICAL CHEMISTRY 37871

by guest on February 10, 2016http://w

ww

.jbc.org/D

ownloaded from

(0.1–100 �M Ca2�) of three independent measurements donein duplicate, indicates that wild-type and triadin-null prepara-tions had no significant difference in their sensitivity for Ca2�

activation (0.98 � 0.09 �M and 0.79 � 0.04 �M, for wild-typeand null mice, respectively, p � 0.05) or Ca2� inhibition. Caf-feine (0.5–50 mM) caused an increase in channel activity in bothwild-type and triadin-null membrane preparations, but there wasnodifference in their sensitivity to caffeine (Fig. 8C; EC50; 9.7�1.2mM and 9.3 � 0.2 mM, respectively, n 2, p � 0.05). Similarly,Mg2� inhibited [3H]ryanodine binding in crudemembrane prep-arations from both wild-type and null muscle in a concentration-dependent manner (Fig. 8B) with almost identical sensitivity(IC50; 0.54 � 0.03 mM and 0.54 � 0.04 mM, respectively, n 2,p � 0.05). Overall these results suggest that the absence of tria-din expression does not significantly affect RyR1’s sensitivity toCa2� and Mg2�, a result consistent with Scatchard analysis ofthe data that show no changes in high affinity [3H]ryanodinebinding between wild-type and triadin-null muscle prepara-tions (see supplemental Fig. S3).

DISCUSSION

Direct biochemical evidence suggests that triadins havestrong interactions with several key components of the CRU(12, 35–37), and several authors have suggested that they maybe critical for skeletal muscle function (3, 19, 21, 23). In thepresent study we have addressed this by creating pan-triadin-null mice. Triadin-specific antibodies failed to detect the pres-ence of skeletal (95 and 60 kDa) and cardiac-specific (32 and 40kDa) isoforms of triadin in skeletal and cardiacmuscles of thesemice. The antibody used for these studies would be expected tocross-react with all triadin isoforms, because it was raisedagainst a peptide specific for an intraluminal domain shared byall triad-localized triadin isoforms in skeletal (Trisk 95 andTrisk 51) and cardiac muscle (CT1: 32 kDa, CT2: 35 kDa, andCT3: 75 kDa) (25, 38). In addition this antibody would also beexpected to bind to the recently reported non-triad-localized49- and 32-kDa skeletal muscle-specific isoforms (23). Its spec-ificity herewas demonstrated by the fact that the same antibodywas highly effective in detecting and differentiating betweenthese isoforms in wild-type muscles and heart and confirmingthat triadin-1, the predominant cardiac isoform (25), wasabsent in cardiac muscle from null animals. Therefore, the factthat the null mice have a seemingly normal phenotype, repro-ducing, moving, and surviving normally suggests that if triadinshave a role in muscle function, it is either a minor one, or onethat can be taken over by some compensatory mechanism.Direct testing of EC-coupling properties in wild-type versus

null myotubes and adult muscle failed to demonstrate an oblig-atory role for triadin in regulating skeletal-type EC coupling.No changes in the kinetics of Ca2� release were detectable, andthe only significant change in Ca2� release properties is a clearreduction in the magnitude of the calcium transient elicited bydepolarization, which is 25–30% smaller than wild type in bothmyotubes and adult muscles. This result is in agreement with arecent study showing that the expression of triadin-bindingdeficient RyR1 in dyspedic myotubes resulted in electricallyevoked Ca2� transients with up to 50% smaller amplitude thanwild-type RyR1 (39). Despite the reduction in Ca2� release our

data showed no clear differences in force development betweenwild-type and triadin-null muscles. This result, however, can-not rule out either the existence of amore subtle phenotype thatcould be masked by the large variation in our tension measure-ments or an altered phenotype that could only be exposed byheavy exercise or other forms of stress.The role of triadins on skeletal EC coupling have been stud-

ied using several experimental approaches, including overex-pression of different skeletal isoforms of triadin (19), expressionof triadin binding-deficient RyRs in dyspedic myotubes (39),and knockdown of the expression of the 95-kDa isoform usingsiRNAs in C2C12 myotubes (40). All these studies suggest thattriadin plays an important role in regulating skeletal EC cou-pling, because all of them demonstrated a significant reductionin the amplitude of the depolarization-evoked Ca2� transient.In agreement with these studies here we show that total abla-tion of triadin expression resulted in a noticeable reduction inskeletal EC coupling gain, but the reduction was considerablysmaller in magnitude than was reported in the former studies.Here the reduced Ca2� release appears to be explained on thebasis of a smaller SR Ca2� pool. Interestingly, studies in myo-tubes expressing triadin binding-deficient RyRs reveal similarreduction in caffeine-induced Ca2� transient amplitude (39,41), suggesting the possibility that the lack of triadin interactionwith RyR1 results in a similar reduction of the SR calcium poolsas we report here. It is worth mentioning that, in all of theprevious studies, the effect of triadin on EC coupling has beenassayed after acute disruption of RyR1-triadin interaction, insome cases even with the endogenous triadin still present (19,39). It is likely that our triadin-null myotubes andmuscles havebeen exposed to a long term adaptation process that hasallowed them to gradually compensate for the lack of triadins.This adaptation process may not take place during transientdisruption experiments where the cells are prompted torespond to acute changes in protein expression. Towhat extentthe short versus long term disruption of RyR1-triadin interac-tion can account for the discrepancies in severity of theobserved phenotypes is unknown. However, from our data it isclear that ablation of the expression of all currently known iso-forms of triadin does not prevent the occurrence of nearly nor-mal EC coupling in mouse skeletal muscles.Previous hypotheses for a role of triadin in controlling RyR

activity, either by itself or as amediator between Csq and RyR1,have come primarily from indirect evidence, based on the pre-sumed disruption of the RyR-triadin or RyR-triadin-Csq asso-ciation in vitro by antibodies (18, 35), peptides (35), triadin itself(20), and changes in ionic strength and/or intra-luminal cal-cium concentration (42). Our findings that absence of triadinshas no apparent effect on [3H]ryanodine binding activity to iso-lated SR vesicles seems contrary to previous [3H]ryanodinebinding and bilayer studies showing that triadin itself caninhibit RyRs activity (20) and to the effect that mutations in theRyR1 putative triadin binding site have on caffeine-inducedCa2� transients (41). This discrepancymay be explained in partby the differences in experimental binding condition used inour study (crude membrane homogenates) and those in Ohku-ra’s study performed in CHAPS-purified RyR1 in the presenceof soybean lecithin and dithiothreitol, two conditions known to

Triadin Knockout and Skeletal Muscle Function

37872 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 282 • NUMBER 52 • DECEMBER 28, 2007

by guest on February 10, 2016http://w

ww

.jbc.org/D

ownloaded from

influence RyR channel behavior. In addition, pretreatment oftriadin 95 with dithiothreitol in Ohkura’s study prevented theinhibitory effect of triadin on [3H]ryanodine binding, suggest-ing that functionally meaningful interactions between RyR1and triadins may be highly dependent on the oxidation state ofboth proteins. Despite the fact that our [3H]ryanodine bindingstudies failed to detect any functional differences betweenwild-type and triadin-null muscles, the small but significant increasein themyoplasmic resting freeCa2� levels observed in both nullmyotubes and adult muscles do suggest the possibility thatthere is an abnormally higher basal activity of RyR1 in triadin-null muscle.An alternative possibility is that another regulatory protein

with redundant or synergistic function with triadin could befunctionally compensating for the lack of triadins in our mem-brane studies. Based on the major reshaping of the expressionprofile of several SR proteins observed in null muscles, thispossibility seems likely. In this regard, junctin (26 kDa), whichhas a similar single membrane-spanning domain, a short cyto-plasmicN terminus, and a longer luminal C-terminal tail that ishighly conserved between both proteins (33, 43) and containsthe joint putative interaction site (junctin with triadin and viceversa) and the interaction site with Csq and RyRs (43, 44),would be a likely candidate. In cardiac tissue junctin and triadinhave been shown to have synergistic function in associatingCsqwith the junctional SR membrane (45) and to positively modu-late the activity of RyR channels fused into lipid bilayers (46).However, this does not seem to be the case in the triadin-nullmice. Our data show that, whereas in myotubes the ablation oftriadin expression resulted in no significant changes in junctinexpression in adult muscle fibers, it caused a significant reduc-tion of junctin expression, revealing a rather negative compen-sation. These results were summed to the fact that triadin-nullmyotubes appear to have a higher RyR1 basal activity than adultfibers and seem to support an inhibitory role for 95- and/or60-kDa triadin on RyR1 function. This is consistent with previ-ous biochemical (1, 20, 42) and physiological data (19) and sug-gests that unlike cardiac muscle in skeletal muscle triadin andjunctin may have opposite influences on Ca2� release regula-tion. Thus a mismatch of the absence of triadin expression (anegative regulator) summed with an incomplete down-regula-tion of junctin (a positive regulator) could result in a net acti-vating effect onCRU function. This in turn could be responsiblefor an augmented myoplasmic resting free Ca2� such as weobserved in both cultured myotubes and intact muscle. Thefacts, that junctin-knockout mice, like the triadin-null mice, donot seem to have a lethal phenotype (47) but siRNAknockdownof both junctin and triadin causes significantly impaired ECcoupling in myotubes (40), seem to support the idea that bothproteins have coordinated functions.In viewof the possible role of triadin and junctin in anchoring

Csq within the junctional SR, it is somewhat surprising thatmovement of Csq away from the triad is rarely observed in thetriadin-null muscles that not only totally lack triadin but alsohave a reduced amount of junctin. This may indicate that just afew anchoring sites may be sufficient to hold polymerized Csqin place. The occasional presence of Csq at sites over than triadshas been detected in normal frog muscle (48), and thus it is

likely that the same Csq displacement in triadin-null muscledoes not have a strong functional effect. It was also interestingthat any structural alterations seen in the triadin-null muscleswere restricted only to fast twitch fibers. Somewhat similarstructural alterations have been reported in both JP-1 (49, 50)and Csq1 (51) knockout mice, and the expression of both ofthese proteins is down-regulated in triadin-nullmice.However,down-regulation of JP-1 and Csq1 cannot account for the dif-ferential response of fast muscles, because a comparable reduc-tion in Csq and an even greater reduction in JP-1 were alsoobserved in triadin-null Sol muscles that showed no ultrastruc-tural alterations. A major difference between the two muscletypes was the level of junctin expression. Unlike Sol muscle,which displayed unchanged expression levels of junctin, EDLmuscles had a 30% reduction of junctin expression. Thus, it ispossible that the deletion of triadin combined with the reducedexpression of junctin could account for the mislocalization ofCsq and in the structural organization of the jSR observedbetween these two muscle types.Nonetheless, we cannot rule out that the altered orientation

of triads may have a different origin. In this respect it is impor-tant to note that fast but not slow fibers seem to readily respondto a variety of stimuli (denervation and lack of Csq) with similardisarrangement of the triads (51, 52).Overall, our data suggest that (i) 95- and 60-kDa triadin are

indirectly involved in organizing the molecular complex of thetriad, although they are not essential components; (ii) triadinsare not required to support EC coupling in skeletal muscle; and(iii) triadins seem to play a negative regulatory role in RyR1activity and thereby contribute to the modulation of globalCa2� homeostasis as evidenced by an increased myoplasmicresting free Ca2� in triadin-null muscles and myotubes.

REFERENCES1. Beard, N. A., Casarotto, M. G., Wei, L., Varsanyi, M., Laver, D. R., and

Dulhunty, A. F. (2005) Biophys. J. 88, 3444–34542. Franzini-Armstrong, C., and Protasi, F. (1997) Physiol. Rev. 77, 699–7293. Marty, I. (2004) Cell Mol. Life Sci. 61, 1850–18534. Meissner, G. (2002) Front. Biosci. 7, d2072–20805. Flucher, B. E., and Franzini-Armstrong, C. (1996) Proc. Natl. Acad. Sci.

U. S. A. 93, 8101–81066. Peng, M., Fan, H., Kirley, T. L., Caswell, A. H., and Schwartz, A. (1994)

FEBS Lett. 348, 17–207. Brandt, N. R., Caswell, A. H., Wen, S. R., and Talvenheimo, J. A. (1990) J.

Membr. Biol. 113, 237–2518. Knudson, C. M., Stang, K. K., Jorgensen, A. O., and Campbell, K. P. (1993)

J. Biol. Chem. 268, 12637–126459. Carl, S. L., Felix, K., Caswell, A. H., Brandt, N. R., Brunschwig, J. P., Meiss-

ner, G., and Ferguson, D. G. (1995)Muscle Nerve 18, 1232–124310. Marty, I., Robert, M., Villaz, M., De Jongh, K., Lai, Y., Catterall, W. A., and

Ronjat, M. (1994) Proc. Natl. Acad. Sci. U. S. A. 91, 2270–227411. Motoike, H. K., Caswell, A. H., Smilowitz, H. M., and Brandt, N. R. (1994)

J. Muscle Res. Cell Motil. 15, 493–50412. Kim, K. C., Caswell, A. H., Talvenheimo, J. A., and Brandt, N. R. (1990)

Biochemistry 29, 9281–928913. Guo, W., and Campbell, K. P. (1995) J. Biol. Chem. 270, 9027–903014. Caswell, A. H., Motoike, H. K., Fan, H., and Brandt, N. R. (1999) Biochem-

istry 38, 90–9715. Glover, L., Culligan, K., Cala, S., Mulvey, C., and Ohlendieck, K. (2001)

Biochim. Biophys. Acta 1515, 120–13216. Lee, J. M., Rho, S. H., Shin, D. W., Cho, C., Park, W. J., Eom, S. H., Ma, J.,

and Kim, D. H. (2004) J. Biol. Chem. 279, 6994–7000

Triadin Knockout and Skeletal Muscle Function

DECEMBER 28, 2007 • VOLUME 282 • NUMBER 52 JOURNAL OF BIOLOGICAL CHEMISTRY 37873

by guest on February 10, 2016http://w

ww

.jbc.org/D

ownloaded from

17. Kobayashi, Y. M., Alseikhan, B. A., and Jones, L. R. (2000) J. Biol. Chem.275, 17639–17646

18. Brandt, N. R., Caswell, A. H., Brunschwig, J. P., Kang, J. J., Antoniu, B., andIkemoto, N. (1992) FEBS Lett. 299, 57–59

19. Rezgui, S. S., Vassilopoulos, S., Brocard, J., Platel, J. C., Bouron,A., Arnoult,C., Oddoux, S., Garcia, L., De Waard, M., and Marty, I. (2005) J. Biol.Chem. 280, 39302–39308

20. Ohkura,M., Furukawa, K., Fujimori, H., Kuruma, A., Kawano, S., Hiraoka,M., Kuniyasu, A., Nakayama, H., and Ohizumi, Y. (1998) Biochemistry 37,12987–12993

21. Marty, I., Thevenon, D., Scotto, C., Groh, S., Sainnier, S., Robert, M.,Grunwald, D., and Villaz, M. (2000) J. Biol. Chem. 275, 8206–8212

22. Thevenon, D., Smida-Rezgui, S., Chevessier, F., Groh, S., Henry-Berger, J.,Beatriz Romero, N., Villaz, M., DeWaard, M., and Marty, I. (2003) Bio-chem. Biophys. Res. Commun. 303, 669–675

23. Vassilopoulos, S., Thevenon, D., Rezgui, S. S., Brocard, J., Chapel, A.,Lacampagne, A., Lunardi, J., Dewaard, M., and Marty, I. (2005) J. Biol.Chem. 280, 28601–28609

24. Marty, I., Robert, M., Ronjat, M., Bally, I., Arlaud, G., and Villaz, M. (1995)Biochem. J. 307, 769–774

25. Kobayashi, Y. M., and Jones, L. R. (1999) J. Biol. Chem. 274, 28660–2866826. Lakso, M., Pichel, J. G., Gorman, J. R., Sauer, B., Okamoto, Y., Lee, E., Alt,

F. W., and Westphal, H. (1996) Proc. Natl. Acad. Sci. U. S. A. 93,5860–5865

27. Rando, T. A., and Blau, H. M. (1997)Methods Cell Biol. 52, 261–27228. Perez, C. F., Lopez, J. R., and Allen, P. D. (2005) Am. J. Physiol. 288,

C640–C64929. Brown, L. D., Rodney, G. G., Hernandez-Ochoa, E., Ward, C. W., and

Schneider, M. F. (2007) Am. J. Physiol. 292, C1156–C116630. Lopez, J. R., Contreras, J., Linares, N., and Allen, P. D. (2000) Anesthesiol-

ogy 92, 1799–180631. Wang, Y., Xu, Y., Guth, K., and Kerrick, W. G. (1999) J. Muscle Res. Cell

Motil. 20, 645–65332. Laemmli, U. K. (1970) Nature 227, 680–68533. Jones, L. R., Zhang, L., Sanborn, K., Jorgensen, A. O., and Kelley, J. (1995)

J. Biol. Chem. 270, 30787–3079634. Sato, T. (1968) J. Electron Microsc. (Tokyo) 17, 158–15935. Groh, S., Marty, I., Ottolia, M., Prestipino, G., Chapel, A., Villaz, M., and

Ronjat, M. (1999) J. Biol. Chem. 274, 12278–1228336. Guo, W., Jorgensen, A. O., and Campbell, K. P. (1996) Soc. Gen. Physiol.

Ser. 51, 19–2837. Murray, B. E., and Ohlendieck, K. (1997) Biochem. J. 324, 689–69638. Hong, C. S., Ji, J. H., Kim, J. P., Jung, D. H., and Kim, D. H. (2001) Gene

(Amst.) 278, 193–19939. Goonasekera, S. A., Beard, N. A., Groom, L., Kimura, T., Lyfenko, A. D.,

Rosenfeld, A., Marty, I., Dulhunty, A. F., and Dirksen, R. T. (2007) J. Gen.Physiol. 130, 365–378

40. Wang, Y., Li, X. H., Duan, H. Z., Fulton, T. R., and Meissner, G. (2007)Biophys. J. (suppl.) 20a, Abstr. 262a

41. Lee, E. H., Song, D. W., Lee, J. M., Meissner, G., Allen, P. D., and Kim do,H. (2006) Biochem. Biophys. Res. Commun. 351, 909–914

42. Beard, N. A., Sakowska, M. M., Dulhunty, A. F., and Laver, D. R. (2002)Biophys. J. 82, 310–320

43. Zhang, L., Kelley, J., Schmeisser, G., Kobayashi, Y. M., and Jones, L. R.(1997) J. Biol. Chem. 272, 23389–23397

44. Zhang, L., Franzini-Armstrong, C., Ramesh, V., and Jones, L. R. (2001) J.Mol. Cell. Cardiol. 33, 233–247

45. Tijskens, P., Jones, L. R., and Franzini-Armstrong, C. (2003) J. Mol. Cell.Cardiol. 35, 961–974

46. Gyorke, I., Hester, N., Jones, L. R., and Gyorke, S. (2004) Biophys. J. 86,2121–2128

47. Yuan, Q., Fan, G. C., Dong, M., Altschafl, B., Diwan, A., Ren, X., Hahn,H. H., Zhao, W., Waggoner, J. R., Jones, L. R., Jones, W. K., Bers, D. M.,Dorn, G.W., 2nd,Wang, H. S., Valdivia, H. H., Chu, G., and Kranias, E. G.(2007) Circulation 115, 300–309

48. Sommer, J. R., High, T. D., and Taylor, I. (1994) in Proceedings of the 52ndAnnual Meeting of the Microscopy Society Of America (Bailey, G. W., andGarratt-Reed, A. J., eds) p. 192, San Francisco Press, Inc., San Francisco

49. Ito, K., Komazaki, S., Sasamoto, K., Yoshida, M., Nishi, M., Kitamura, K.,and Takeshima, H. (2001) J. Cell Biol. 154, 1059–1067

50. Komazaki, S., Ito, K., Takeshima, H., and Nakamura, H. (2002) FEBS Lett.524, 225–229

51. Paolini, C., Quarta, M., Nori, A., Boncompagni, S., Canato, M., Volpe, P.,Allen, P. D., Reggiani, C., and Protasi, F. (2007) J. Physiol. 583, 676–784

52. Takekura, H., Tamaki, H., Nishizawa, T., and Kasuga, N. (2003) J. MuscleRes. Cell Motil. 24, 439–451

Triadin Knockout and Skeletal Muscle Function

37874 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 282 • NUMBER 52 • DECEMBER 28, 2007

by guest on February 10, 2016http://w

ww

.jbc.org/D

ownloaded from

24

SUPPLEMENTAL DATA

Freeze-fracture analysis. Muscles were infiltrated in 30% (v/v) glycerol in H2O, frozen in liquid N2-

cooled propane, and then fractured, platinum shadowed and replicated in a Balzers BFA 400 apparatus

(Bal-tec AG Balzers Switzerland). Sections and replicas were observed using a Philips 410 microscope

(Philips Electron Optics, Mahwak, NJ). Images were recorded either on film or using a Hamamatsu

C4742-95 digital imaging system (Advanced Microscopy Techniques, Chazy, NY)

Fig. 1. Histograms display the number of occurrence of longitudinally oriented triads (expressed as

percentage) in fast and slow twitch muscles from wild type and triadin-null. The frequency of triad

orientations was counted in images from longitudinal sections covering 36 !m2 areas. Average data are

given in Table I.

Fig. 2. Tetrads formation in triadin-null muscle. Freeze–fracture replicas of plasmalemma from diphragm

muscle of E18 neonatal triadin-null mice showing DHPR particles arranged in clusters. Within the cluster

groups of up to four particles (tetrads) are depicted in circles.

Fig. 3. Dose-dependence of [3H]ryanodine binding to microsomal fraction of skeletal muscle. 40-50!g/ml

of SR membrane from wild type and triadin-null skeletal muscle were incubated with 2-35 nM

[3H]ryanodine in 0.5 M NaCl, 100!M CaCl2, 25 mM Hepes pH 7.4 for 1.5 h at 37 °C. Specific [

3H]-

ryanodine binding was plotted and fitted to a one site binding hyperbola. Linear Scatchard plot (Inset)

revealed that both, wild type and null membrane preparations, presented a single class of binding site with

near identical Kd. Data presented as mean ± SD, n=4.

25

Suppl. Fig. 1

26

Suppl. Fig.2

27

Suppl. Fig. 3

Todd Miller, Paul D. Allen and Claudio F. PerezKobayashi, Ying Wang, W. Glenn L. Kerrick, Anthony H. Caswell, James D. Potter, Xiaohua Shen, Clara Franzini-Armstrong, Jose R. Lopez, Larry R. Jones, Yvonne M.

Excitation-Contraction Coupling in Skeletal Muscle Homeostasis but Are Not Essential for2+Triadins Modulate Intracellular Ca

doi: 10.1074/jbc.M705702200 originally published online November 2, 20072007, 282:37864-37874.J. Biol. Chem. 

  10.1074/jbc.M705702200Access the most updated version of this article at doi:

 Alerts:

  When a correction for this article is posted• 

When this article is cited• 

to choose from all of JBC's e-mail alertsClick here

Supplemental material:

  http://www.jbc.org/content/suppl/2007/12/27/M705702200.DC1.html

  http://www.jbc.org/content/282/52/37864.full.html#ref-list-1

This article cites 50 references, 21 of which can be accessed free at

by guest on February 10, 2016http://w

ww

.jbc.org/D

ownloaded from