A Rossi PHD Thesis Final 2
A Rossi PHD Thesis Final 2
A Rossi PHD Thesis Final 2
By
of the
Doctor of Philosophy
Supervised by
University of Rochester
Rochester, NY
2009
ii
DEDICATION
To my parents, for free dinners, clean laundry, and most importantly, their
CURRICULUM VITAE
The author was born in Elmira, New York on March 4, 1980. She attended
Nazareth College of Rochester from 1998 to 2002, where she received the Dean’s
Scholar Award. In 2001, she held an internship with Corning Incorporated in the
different formulations of polymer coatings for optical fibers. In 2002, she graduated
Magna Cum Laude with a Bachelor of Science degree in Biology. Shortly thereafter,
in the Fall of 2002, she began graduate studies at the University of Rochester School
of Medicine and Dentistry. She pursued her research in Pharmacology under the
direction of Professor Robert T. Dirksen and received the Master of Science degree
from the University of Rochester in 2005. In 2006, she was awarded a Biophysical
Society Student Travel Grant to support her attendance and presentation at the 50th
Annual Biophysical Society Meeting. From 2006 to 2008, she received additional
training and financial support from a National Institute of Dental and Craniofacial
Chicago, Illinois.
PUBLICATIONS
Rossi AE, Dirksen RT. Sarcoplasmic Reticulum: The Dynamic Calcium Governor of
Durham WJ*, Aracena-Parks P*, Long C*, Rossi AE, Goonasekera SA,
RT, Hamilton SL. RyR1 S-Nitrosylation Underlies Environmental Heat Stroke and
Reviews 2009;37(1):29-35.
2009;20(3):1058-67.
ACKNOWLEDGEMENTS
There are several deserving individuals who I would like to thank for making
advisor Dr. Robert T. Dirksen for allowing me the freedom to pursue research outside
of the lab “box” and pioneer new techniques in the laboratory. His mentoring has
career in research, which I appreciate more than I can express. Furthermore, I extend
my thanks to the members of my thesis committee: Dr. Paul. Brookes, Dr. Shey-
Shing Sheu, and Dr. David Yule, for helpful input and guidance on my project,
research. I express my gratitude to Dr. Werner Melzer at the University of Ulm for
kindly allowing me to use the automated detection software for calcium spark
detection. Also, I would like to thank Dr. Susan Hamilton at Baylor College of
Medicine for the knock-in mice and for graciously hosting my visit to her laboratory.
Pennsylvania for helpful technical discussions and sharing her expertise. Most
importantly, I thank Drs. Simona Boncompagni and Feliciano Protasi at the Gabriele
of Simona’s time and effort on the electron microscopy analysis, and for allowing me
to present some of her work. Additionally, this research would not have been possible
without the financial support from the National Institutes of Health Grants AR044657
vi
Linda Groom and Sara Geitner for all they do, day-to-day, to keep the laboratory
running. In particular, I thank Linda for helping to pass the time between
experiments, exchanging notes about the previous night’s reality TV shows. Also, I
am grateful to all current and former Dirksen laboratory members for their productive
scientific discussions, in particular, Dr. John Lueck, for his support and interest in
for sharing frustrations and successes in research, and for reminding me to relax and
have fun. Finally, I am grateful to my entire family for their love and support
throughout my graduate career and my life, especially my sisters Julie and Johanna,
parents for teaching me the value of hard work and the power of prayer and positive
thinking. Thank you for your encouragement. Most of all, to my twin, Marie, thank
you for being my toughest critic, my biggest fan, and my best friend.
vii
ABSTRACT
release units (CRUs), or triads, the structures formed by the close apposition of
(i.e. Ca2+ sparks). In this way, a structural association of mitochondria with the triad
disposition during skeletal muscle development. To this end, the following research
young (0.5-1 month of age) and adult (2-4 months of age) mice using both confocal
targeting is a highly coordinated and developmentally regulated process and suggests that
mitochondrial calcium uptake and function are altered during muscle development.
CRU targeting, mitochondrial calcium uptake was measured in FDB fibers from
young (1 month of age) and adult (4 months of age) mice by confocal imaging of
in triadic and longitudinal mitochondria was significant in fibers isolated from both 1
and 4 month-old mice. The increase in mitochondrial free calcium persisted for the
duration of tetanic stimulation and returned to resting levels within twenty minutes
after stimulation. Interestingly, the global calcium chelator EGTA had varied effects
calcium uptake in 4 month triadic mitochondria. Taken together, these results suggest
that mitochondrial calcium uptake was permitted by the global calcium and provide
originating from the SR, mitochondria, and/or the structural interaction between them
could lead to muscle dysfunction and disease. Accordingly, the final goal of this
from a mouse model of the muscle disorders Malignant Hyperthermia (MH) and
mitochondria.
TABLE OF CONTENTS
DEDICATION .............................................................................................................. ii
ACKNOWLEDGEMENTS .......................................................................................... v
FOREWORD .............................................................................................................. xx
INTRODUCTION ........................................................................................................ 1
1.1. RATIONALE................................................................................................... 58
Development ........................................................................................... 62
1.3.2. JC-1 Emission Distribution and Magnitude are Altered During Postnatal
Development ........................................................................................... 63
2.1 RATIONALE.................................................................................................... 75
Stimulation .............................................................................................. 82
RELEASE ....................................................................................................... 94
3.1. RATIONALE................................................................................................... 95
4.2.4. Time Lapse Imaging and Analysis of Superoxide Flashes ..................... 119
CONCLUSIONS....................................................................................................... 131
LIST OF TABLES
LIST OF FIGURES
Figure I.1. Structural Features of Key Calcium Regulatory Proteins of the Skeletal
Figure I.3. Mitochondria are Precisely Localized Adjacent to the Calcium Release
Development ................................................................................................... 70
Positioning ...................................................................................................... 71
Figure 1.3. Mitochondrial Distribution and Membrane Potential are Altered During
Stimulation ...................................................................................................... 91
Figure 3.1. Osmotic Shock-Induced Ca2+ Sparks and Bursts in FDB Fibers During
Figure 3.3. Ca2+ Spark Amplitude Peaks at 1 Month During Postnatal Development . 113
Figure 3.4. Osmotic Shock-Induced Ca2+ Spark Frequency per CRU Density
Figure 3.5. Osmotic Shock Induces Swelling of T-Tubules and Caveolae ............. 115
Figure 4.1 JC-1 Emission Ratio is Altered During Postnatal Development in Y522S
Figure 4.2. Mitochondrial Disposition and Sarcomere Spacing in Adult FDB Fibers . 128
ABBREVIATIONS
DH, dehydrogenase
F, fluorescence
WT, wild-type
FOREWORD
Portions of the text from the Introduction, as well as Figures I.1 and I.2, were
reprinted with permission of John Wiley & Sons, Inc. from the following
301703.html):
Additional text from the Introduction and Figures I.3, I.4, and I.5 were
permissions/journals.html):
author. Figures 1.1 and 1.3 were adapted from the following publication with
copyright permission granted by the American Society for Cell Biology (outlined
at http://www.molbiolcell.org/misc/ifora.shtml):
and 4. The data in Figures I.3B, I.3C, 1.2 and 3.5 and Tables 1.1 and 3.2 were
Figure 4.3 was reproduced from the following publication with copyright permission
Heat Stroke and Sudden Death in Y522S RyR1 Knock-in Mice. Cell
2008;133(1):53-65.
1
INTRODUCTION
2
Skeletal muscle contraction, at the most basic level, requires calcium and
ATP, and thus, is under the control of two major organelles, the sarcoplasmic
reticulum (SR) and mitochondria. The SR is the primary regulator of muscle calcium
cycling, mediating calcium storage, release, and reuptake. Equally important are the
mitochondria, the major source of cellular ATP. Although mitochondria are adjacent
isolation. Therefore, the purpose of this work is to determine the degree of structural
governor in muscle that provides an automatic feedback control for altering and
regions of the SR and the transverse tubular membrane, as well as envelopment of the
myofibrils by the longitudinal SR, uniquely position this organelle for controlling
calcium release and reuptake during muscle contraction. Given its central role in
excitation, and a slower means of calcium reuptake and maintained calcium storage
storage, release, and reuptake are balanced by the concerted action of three major
and IP3 receptors) for calcium release; and 3) sarco(endo)plasmic reticulum calcium-
the level and calcium binding capacity of proteins within the SR. A number of
different acidic calcium binding proteins participate in calcium buffering and help
maintain SR free calcium at levels around 1 mM.93 The most important protein in
coupling.21
4
Calsequestrin (CSQ)
MacLennan and colleagues.158 Initial studies found that the protein is highly acidic,
40 µM) and high capacity (~40 mol calcium/mol protein at pH 7.5) calcium binding.
Thus, the protein was given the name “calsequestrin” for its ability to sequester
cardiac) have been identified and differ in the number of acidic residues in their
calsequestrin has fewer acidic residues in its C-terminal tail than the cardiac
isoform.21 Fast- and slow-twitch skeletal muscle express the skeletal isoform,10,28,68
whereas the cardiac isoform is expressed in both cardiac muscle and slow-twitch
adult muscle.10,190,204 While at the protein level, the switch from the cardiac to
sequences of amino acids to form a monomer with a hydrophilic core. Each domain
has a disk-like shape with a four-stranded β-sheet at its hydrophobic core and two
hydrophilic α-helices on each side of the sheet. The structure of each domain is
hydrophilic side chains form crevices that are similar to the redox centers of
of which are in the C-terminus of the monomer. These acidic residues are important
for calcium binding. Rather than the presence of discrete calcium binding sites, pairs
of acidic residues create a net surface charge to which calcium binds.248 As calcium
concentrations in the low micromolar range (10 µM) cause collapse of the monomer’s
not membrane bound;87 it resides in the lumen of the SR. Calsequestrin is anchored
to the junctional face of the SR membrane through interactions with junctin, triadin,
and RyR1 (Figure I.2).266 Junctin and triadin binding involves an aspartate-rich
the SR membrane, a number of different triadin splice variants have also been
identified.168,230,243 The bulk of the full-length triadin protein is located in the lumen
of the SR,132 where it interacts with calsequestrin, junctin, and RyR1. Multiple Lys-
Glu-Lys-Glu (KEKE) motifs located in the luminal portion of triadin are required for
long, positively charged luminal C-terminus (Figure I.1C). Like triadin, the luminal
structure of junctin binds both calsequestrin and RyR1.124 By binding to both RyR1
and calsequestrin, triadin, and junctin anchor calsequestrin and thereby, increase
calcium buffering close to sites of SR calcium release (Figure I.2). In addition, the
calcium release, and thus, skeletal muscle contractility. The means by which
calcium storage protein has recently been appreciated. In native SR, HRC
(699–862 amino acids) exists as a multimer of five or more subunits.222 The primary
terminal cysteine-rich region (Figure I.1D).112 The central region is responsible for
both calcium binding and interactions with other junctional proteins. Specifically, the
concentration of acidic residues in the central region may provide charge on the
HRC also binds to triadin, perhaps signifying a greater role for this molecule
than simply increasing the number of calcium binding sites in the SR. The repeats in
the central region are necessary for the interaction of HRC with triadin.147 The
histidine-rich region contains two different types of repeats, A and B, which are
responsible for binding to the KEKE motifs in the luminal portion of triadin. In both
8
types of repeats, a string of acidic amino acids are flanked by histidine residues. The
number of acidic residues varies between types A and B. Similarly, the number of
terminal histidine residues differs; type A ends with a single residue, whereas type B
ends with multiple histidines.114 More than 4.5 repeats are required for triadin
binding, although the types of repeats do not affect the strength of interaction. In
addition, the interaction between HRC and triadin is calcium sensitive, such that the
will be to determine whether the dual functions of HRC and calsequestrin (calcium
and triadin binding) are redundant, or if the two proteins exhibit unique roles in
Junctate
protein in the junctional SR (Figure I.1E).237 Junctate is a 298 amino acid integral SR
membrane protein (33 kDa) with structural homology to junctin. Junctate is derived
from alternative splicing of the genes that encode both junctin and aspartyl β-
junctin, the first 23 of which are located in the cytoplasm. After a single-pass
transmembrane domain, the rest of the protein is located in the lumen of the SR. The
with an apparent kD of 217 µM, primarily within the negatively charged luminal C-
terminus of the protein.237 Thus, like calsequestrin and HRC, junctate is a high-
the calcium storage/release capacity of the SR. However, unlike calsequestrin and
HRC, the calcium binding sites of junctate are tethered close to sites of calcium
abundant in skeletal muscle compared with pancreas, heart, brain, and kidney, it is
muscle.237
Sarcalumenin (SAR)
calcium storage close to sites of calcium release. However, this does not preclude an
important role for calcium binding proteins within nonjunctional regions of the SR.
moderate-affinity (kD of 300 µM) calcium binding protein that is primarily located in
10
longitudinal SR. The full-length amino acid sequence of SAR (436 acidic amino
colocalizes with SERCA, suggesting that the protein not only contributes to calcium
buffering within the longitudinal SR, but may also influence calcium reuptake.146
These conclusions are supported by recent results obtained from SAR knockout
mice.262
been associated with a human skeletal muscle myopathy. However, in a few families,
mutation involves the substitution of a negatively charged aspartic acid residue for a
histidine at position 307.138 This aspartate residue is located in the negatively charged
C-terminus of the protein, and mutation of this residue to histidine may interfere with
been linked to three different nonsense mutations that each lead to the introduction of
11
release in skeletal muscle, it would not be surprising if future work were to identify
signal (i.e. a calcium transient). This signal is then harnessed to do work by driving
the terminal cisternae of the SR.91 In adult skeletal muscle, RyR1 proteins, or feet,
are arranged in ordered arrays within the SR terminal cisternae on either side of the
transverse tubule (T-tubule). Within these RyR1 arrays, every other calcium release
channel is associated with a “tetrad” of four DHPRs in the T-tubule membrane. The
line, though its specific arrangement with respect to the myofibrils is largely
sections, start to form ordered arrays in the junctional gap between the SR and
sarcolemma. Assembly of the RyR1 arrays provides the structural requirement for
represent early stages in formation of the EC coupling apparatus. At this point, clear
and exhibiting a longitudinal orientation, are the foundations of the future T-tubular
these two membrane systems results in an increase in the formation of robust SR-T-
13
tubule junctions, which are enriched in RyR1 arrays with alternately apposed DHPR
tetrads.88,90 Finally, the formation of the fully developed triad involves reorientation
along the junction between the A- and I-bands (A-I) of the sarcomere in adult
to the A-I junction, the speed of the contractile response increases, and the time from
the start of the action potential to the beginning of recorded muscle tension
tubular DHPRs and RyR1s in the terminal SR, guaranteeing the fidelity2 of EC
coupling. Essentially, the tight triad junction ensures that each action potential
triggers calcium release via a defined signaling interaction between DHPR and RyR1
calcium release, while retrograde coupling (RyR1 to DHPR) reflects the influence of
Junctophilin
the triad structure which is essential for proper coupling between the DHPR and
Junctophilin MORN motifs are thought to be responsible for binding to the T-tubular
phospholipids.227 Four separate junctophilin subtypes (JP-1, JP-2, JP-3, and JP-4)
have been identified. JP-1 and JP-2 are both expressed in skeletal muscle,227 although
the function of JP-1 has been investigated in greater detail (Figure I.1H). JP-1 is
coupling, and thus results in reduced force generation during muscle contraction.120
15
membrane in skeletal, but not cardiac, muscle (Figure I.1I).270 JP-45 also has a single
and interacts with both calsequestrin and the C-terminus of the DHPR α1s-subunit
(Figure I.2),4 suggesting that this protein may regulate calcium release during EC
coupling and store-operated calcium entry during tetanus and fatigue.4 In fact, more
recently JP-45 has been found to be important for functional expression of the
DHPR.5
channel and as a voltage sensor that initiates calcium release through RyR1 during
subsequent to the initial action potential release less calcium than the first due to
stores are fully depleted to keep constant the amount of calcium released.192 In a
sense, the DHPR can be considered the ultimate regulator of calcium release during
EC coupling.
16
The skeletal muscle DHPR is comprised of five different subunits: α1S, β1a,
α2/δ, and γ. The pore-forming α1S-subunit contains four homologous internal repeats
transmembrane segments S5 and S6 of each repeat form the pore of the channel, and
the fourth transmembrane segment of each repeat contains a series of positive charges
II-III loop of the DHPR α1S-subunit and, as yet, undefined regions of the cytosolic
DHPR α1S-subunit46,255 and the C-terminal region of the DHPR β1a-subunit to RyR1
calmodulin binding domains located in both RyR1 and the C-terminal region of the
RyRs were first identified based on their high-affinity binding to the plant
these channels are poorly selective cation channels. They possess weak selectivity
among divalent cations or among monovalent cations, and only modest selectivity for
divalent over monovalent cations. Calcium competes with Mg2+, Na+, and K+ for
charge carriers231). This high conductance permits rapid and substantial calcium
Three RyR isoforms with distinct tissue distributions have been identified:
RyR1 is primarily expressed in skeletal muscle; RyR2 is the major isoform found in
cardiac muscle; and RyR3 is found in small amounts in a wide variety of tissues.86
This work focuses on the skeletal muscle RyR isoform, RyR1 (Figure I.1G). The
The human RyR1 protomer contains 5038 amino acids and has a molecular weight of
~565 kDa. The C-terminal one-fifth of each RyR1 protomer contains the
transmembrane segments that anchor the protein in the terminal SR. The most recent
approximately four-fifths of the protein is located in the space between the SR and T-
tubule and interacts with the DHPR and a variety of other junctional regulatory
proteins.86
calcium; low calcium (1-10 µM) activates calcium release, whereas high calcium (1-
rarely ever reaches millimolar levels, the effect of high calcium on RyR1 activity is
through mechanism, in which calcium released from the SR binds to the activation
and inactivation sites on the cytoplasmic face of the channel240; and 2) a luminal
calcium sensor that increases RyR1 channel activity when SR calcium reaches
luminal calcium is unknown, but appears to involve both direct effects of calcium on
inhibiting release channel activity at normal SR calcium levels (~1 mM), and this
bind the RyR in a calcium-independent manner and prime the channel for activation.
calsequestrin with triadin and junctin are calcium sensitive, such that extreme low (<
10 µM) and high (> 4 mM) levels of luminal calcium promote a conformational
change in the protein that weakens its interaction with triadin and junctin.266 Thus,
this relieves the inhibitory influence of the quaternary complex on calcium release.102
However, even small alterations in luminal SR calcium levels during release may
19
normal activation of release by the voltage sensor. In fact, DHPRs can activate RyRs
over a broad range of SR calcium levels192 and thus fully deplete SR calcium,
ability of the voltage sensor to fully deplete the SR is likely mediated by long-range
allosteric effects by which luminal calcium levels alter inhibition of release channel
RyR1 Regulators
number of soluble molecules (e.g., Mg2+, ATP, calcium, and reactive oxygen species)
and associated proteins (e.g., FKBP12, calmodulin, homer, and PKA) also regulate
the activity of the SR calcium release channel. Cytoplasmic ATP activates whereas
cytoplasmic Mg2+ inhibits channel activity.62,63,217 Both the long and short form of
the homer protein activate calcium release through RyR1.251 Homer multimerizes at
its C-terminal coiled-coil and leucine zipper domains260 to form a junctional signaling
of the fully open and closed states of the channel,122 controlling channel gating
20
coupling.14,16,228
the precise role of the calmodulin interaction with RyR1 in modulating calcium
the channel (residues 1-2401 and 3120-4475) that are major sites of S-nitrosylation
in the SR transmembrane redox potential do not directly alter the amount of calcium
calmodulin kinase II (CaMKII), the effects of which are still highly controversial.
suggested to trigger FKBP12 dissociation from the release channel and subsequent
loss of FKBP12 regulation.199 However, other investigators have found that PKA
binds with high affinity to the C-terminal pore-forming region of the channel.41
(~100 µM) abolish ion conduction through the channel.269 Although not a
that stimulates calcium release from the SR by binding to the release channel.215
calcium release through RyR1 and RyR2, but not RyR3, through interactions with the
by binding to the ion conduction pore of the channel.156 Given its highly cationic
nature, it is possible that, in addition to its effect of blocking the channel pore,
ruthenium red may also alter channel regulation through interactions with calcium
Calcium Influx
operated calcium entry by forming a complex with the IP3 receptor and the transient
23
prolonged or repetitive depolarization but not store depletion.52 Further, ECCE does
not require either Stim-1 or Orai-1, but does require both the voltage sensor (DHPR)
for activation and RyR1. It is clear, then, that SOCE and ECCE are controlled by
distinct molecular complexes.153 The dependence of both SOCE and ECCE on RyR1
between multiple triadic proteins simultaneously controls both calcium release during
pathways.
ATPase (PMCA) pumps and Na+/Ca2+ exchangers balances calcium influx during EC
coupling in cardiac muscle, these calcium transport processes play only a minor role
calcium reuptake controls the rate of muscle relaxation in skeletal muscle and is
There are five primary isoforms of SERCA encoded by three separate genes:
isoform found in fast-twitch skeletal muscle. The last two exons of ATP2A1 are
also results in the production of two C-terminal splice variants (SERCA2a and
skeletal, and neonatal skeletal muscle. SERCA2b is found in both nonmuscle tissues
and some endothelial cells.258,259 Despite significant structural and primary sequence
25
similarities, the tissue-specific expression patterns of each isoform suggests that the
The crystal structure of SERCA1a, a 994 amino acid (110 kDa) monomeric
protein located in the membrane of both terminal cisternae and longitudinal SR,160
segments (M1-M10), with the bulk of the protein, including the N- and C-termini,
located in the cytoplasm (Figure I.1J). The structure can be divided into three parts: a
a stalk domain.233 The cytoplasmic head is subdivided into three distinct domains: 1)
the actuator (A) domain (residues 1-50 and 131-238); 2) the nucleotide binding (N)
domain (residues 360-604); and 3) and the phosphorylation (P) domain (residues 330-
359 and 605-737) (Figure I.1J). The actuator domain facilitates engagement of the
gating mechanism and thereby regulates calcium binding and release.235 The γ-
moiety binds to a site in the N domain.235 The luminal linkers between the
transmembrane segments are short except for the loop between M7 and M8. A
number of residues are essential for coordinating the cooperative binding of two
closely spaced calcium binding sites (5.7 Å apart) in the transmembrane region of the
protein. Site I is formed by Asn768, Glu771, Thr799, Asp800, Glu908, and the
(Glu309, Asn 796, and Asp800) and three main-chain oxygen atoms from M4.167
Two calcium ions bind with high affinity to sites I and II on the cytoplasmic side of
26
the protein at the beginning of the reaction cycle and are ultimately released from the
luminal side of the protein following ATP hydrolysis. For each ATP molecule
hydrolyzed per reaction cycle, two calcium ions from the cytoplasm are pumped into
the SR lumen in exchange for two or three H+ ions.263 In this way, SERCA1a utilizes
the energy stored in ATP to clear myoplasmic calcium during muscle relaxation.
Sarcolipin (SLN)
thought to inhibit SERCA1 activity by decreasing its affinity for calcium and
lowering its Vmax.185 SERCA1 is also inhibited by NO, while its calcium affinity
Specifically, SAR interacts with SERCA1, and SR vesicles isolated from skeletal
in skeletal muscle are similar in wild-type and SAR-deficient mice, the interaction of
SAR with SERCA may serve to increase protein stability and reduce SERCA
27
for relaxation.262
I.2. MITOCHONDRIA
All major muscle fiber types (red, white, and intermediate) exhibit pairs of slender
mitochondria encircling the myofibrils within the I-band on either side of the Z-
line.186 In addition, mitochondria are also found to cluster under the sarcolemma and
(Figure I.3A and 244), consistent with their close apposition to triads (Figure I.3B and
also run the length of the sarcomere, forming a longitudinal lattice between the
myofibrils.244
cell types, including smooth muscle cells, pancreatic acini, and hepatocytes.92 For
hepatic rough ER, and the isolated organelles retain a structural interaction similar to
short (5 nm) synthetic linker to enhance the connection between the two
compartments. Local calcium signaling between the two organelles was augmented
Similarly, close association and tight coupling between mitochondria and the
metabolic coupling during normal muscle activity, and also possibly influence
fatigability and muscle damage during eccentric exercise. Recently, a physical link
between the outer mitochondrial membrane (OMM) and the face of the junctional SR
opposite to that of the RyR1 feet has been suggested to anchor mitochondria to the
during muscle contraction and shortening, but also provide a structural basis for local
ATP production.
Following the power stroke, ATP binding to myosin causes dissociation of the
myosin head from actin, which is required for subsequent iterations of the crossbridge
cycle. In addition, ATP also provides the energy used to drive calcium reuptake by
the SERCA pumps during contractile relaxation. In fact, the majority of ATP
of ATP and pyruvate per glucose. Following glycolysis, the matrix enzyme complex
acetyl-CoA that is then funneled into the tricarboxylic acid (TCA) cycle (Figure I.4).
The series of TCA cycle reactions produce reducing equivalents NADH and
FADH2.34 These products enter the electron transport chain (ETC) at the NADH-
located within the inner mitochondrial membrane (IMM). Transfer of electrons from
NADH to complex I results in accumulation of NAD+ which feeds back into the TCA
cycle for continued NADH generation. Complex I and II transfer the electrons from
important to note that a small percentage of electrons flowing through the respiratory
chain can escape at complex I and the Q cycle. These free electrons combine with
molecular oxygen to produce superoxide anions which can then be converted to other
reactive oxygen species (ROS) (Figure I.4). In the final step, electrons are shuttled
the final electron acceptor.34 Electron transfer at complexes I, III, and IV is balanced
(Figure I.4). This process generates an electromotive proton gradient used to drive
ATP synthesis from adenosine diphosphate (ADP) and inorganic phosphate (Pi) by
31
the F1F0-ATPase (complex V).34 The sum of reactions from the beginning of the
ATP generation between the different types of skeletal muscle underscores the critical
Specifically, slow-twitch muscle fibers are highly enriched in mitochondria, are more
and generates ATP at a much lower rate than OXPHOS, fast-twitch glycolytic fibers
ATP generation.2 While the cellular mechanisms of skeletal muscle fatigue are
complex, a decline in cellular ATP levels during repetitive muscle activation clearly
decreases during fatigue, the stimulatory effect of ATP on calcium release and
activity, mitochondrial adenine nucleotide transport, and flux through the OXPHOS
during the TCA cycle are stimulated by elevations in matrix free calcium (Figure I.4).
First, the phosphatase that converts inactive pyruvate dehydrogenase phosphate to the
mitochondria reveal that Ca2+ activates the phosphatase by decreasing its Km for
limiting for flux through the TCA cycle, calcium stimulation of these enzymes
the ATP synthetic capacity of the F1F0-ATPase.229 In this way, mitochondrial calcium
uptake during EC coupling would serve to stimulate aerobic ATP production in order
33
to help keep pace with increased ATP consumption associated with crossbridge
mitochondrial ATP production has been documented in cultured cells and isolated
fibers.
these discrete regions, coupled with the highly negative inner mitochondrial
membrane potential (-180 mV), provide a sufficient driving force for calcium uptake.
Calcium uptake occurs via the calcium uniporter (CaUP), a rapid calcium uptake
mode (RaM), and the mitochondrial ryanodine receptor (mRyR)(see Figure I.4 and
34
). In cardiac myocytes, rapid transient increases in mitochondrial calcium during
uptake. This uptake is blocked by inhibition of the CaUP with ruthenium red, but not
in slow- and fast-twitch skeletal muscle fibers during both electrical stimulation and
caffeine exposure.212 However, since mitochondria are associated with the SR on the
34
side opposite to sites of RyR1-mediated calcium release (Figure I.3C), this most
tunneling between organelles. These data support the presence of a highly localized
striated muscle. Indeed, Rudolf et al.203 provided striking evidence for efficient
twitch glycolytic fibers. In addition, even when mitochondrial calcium uptake during
tetanus and fatigue is detectable, its magnitude represents only a minor fraction of the
total calcium release, precluding a major role for mitochondria in shaping the global
calcium buffers present in the mitochondrial matrix, even low levels of mitochondrial
coupling.
35
metabolism coupling. Local SR calcium release events, termed calcium sparks, were
sparks provide brief elevations in local calcium release that last on the order of a few
tens of milliseconds and exhibit a spatial width of ~1 µm.131 While calcium sparks
are not observed in mammalian skeletal muscle fibers under basal conditions, they are
suppression of calcium sparks in intact adult mammalian skeletal muscle under basal
adult skeletal muscle cells correlates with an increase in calcium spark frequency,131
Similarly, calcium spark activity following strenuous exercise may involve a physical
calcium buffering in spark suppression has been neither supported nor disproven.
Certainly, evidence presented in the previous section indicates that mitochondria are
TCA cycle substrates diminishes calcium spark activity.117 Moreover, the onset of
accelerates the onset and increases the frequency of calcium sparks,118 whereas ROS
muscle.
accessory proteins (e.g. calmodulin and FKBP12) and increases calcium activation
and the probability of RyR1 channel opening,81,111 calcium sparks in muscle may be
unmasked under conditions that increase the oxidative environment around the
CRU.118,166 Mitochondria are both intimately associated with the CRU (Figure I.3)
superoxide ions at complex I and during the Q cycle, this ROS production is balanced
mitochondria play a critical role in maintaining proper balance of the cellular redox
potential. Also, given their localization at the triad (Figure I.3), mitochondria may
influence calcium spark activity by regulating the local redox environment of the
release by ensuring that the local redox environment around the CRU is sufficiently
reduced. On the other hand, calcium spark activity is increased under conditions that
overwhelm this local redox control mechanism and promote RyR1 oxidation either
suggesting that antioxidants may correct defects in local calcium release and increase
fatigue resistance.2
The findings presented clearly indicate that proper calcium storage, release,
and reuptake are essential for normal physiological functioning of muscle excitation,
contraction, and relaxation. This conclusion is made all the more evident by the fact
disorders (Malignant Hyperthermia, Central Core Disease, and Brody Disease) result
from dysfunction in the control and coordination of different aspects of these three
critical processes.
I.3.1. Fatigue
generation) with muscle activity. Several major mechanisms that cause fatigue have
been identified, most of which originate in the muscle itself and are a direct result of
are energetically expensive processes. In fact, the energy demand of contraction can
accommodate these sudden changes, muscle utilizes creatine phosphate (CrP). CrP is
generated from creatine and MgATP in a reaction catalyzed by creatine kinase. This
39
with the onset of activity, initially maintain a stable level of ATP. However, CrP
generate CrP. Therefore, cellular ATP levels will eventually fall with continued
and their accumulation leads to muscle acidosis, which is classically associated with
fatigue. At the same time, ATP hydrolysis results in an increase in the concentration
of Pi and ADP, as well as Mg2+.2 More importantly, these metabolic changes during
Pi.2
It is also worth noting that free radical and ROS/RNS production are
superoxide in muscle.2 Superoxide can then react with NO· formed by nNOS
The contribution of elevated levels of free radicals and ROS/RNS to muscle fatigue
40
calcium reuptake.2
However, fatigue can arise from a breakdown at any point of the EC coupling
process, not only calcium release and reuptake. For example, changes in membrane
excitability due to alterations in Na+, K+, and/or Cl- fluxes, and subsequent action
potential failure, can result in fatigue. In addition, a decline in performance can occur
appearing over a period of hours to days. Additionally, muscle injury can activate
damage and initiation of muscle repair.196 Free radicals generated during contraction
production.
direction and nature of triads following downhill treadmill running. In addition, other
reduction in SR Ca2+ available for release during successive contractions.40 Over the
long term, SR and T-tubular membrane damage and defective SR Ca2+ reuptake result
in an elevation of myoplasmic Ca2+. Indeed, peak muscle damage occurs 1-3 days
proteases.50
42
state results as the muscle strives to replenish and maintain levels of ATP, ultimately
mechanism as the underlying defect in MH. First, in vitro contracture tests conducted
identified a single point mutation in the RyR1 gene that results in the substitution of a
Cys residue for Arg615.94 This is the only mutation known to cause MH in pigs.161
Third, SR calcium release channels isolated from skeletal muscle of MHS porcine
three main clusters of mutations in RyR1 are linked to MHS in humans: mutations in
43
MH mutations in the N-terminal and central regions of the RyR1 protein are thought
to disrupt a critical interdomain interaction that normally acts to stabilize the resting
calcium levels, increased reactive nitrogen species (RNS) generation, and subsequent
Although linkage to the RyR1 gene is shown for more than 50% of all MH
kindreds, evidence also suggests that other MH-susceptible gene loci exist, including
different point mutations of a highly conserved arginine residue in the III–IV linker of
the skeletal muscle DHPR α1s-subunit (R1086H and R1086C) have been confirmed
for the 1q31 locus.173,177 Similar to MH mutations in RyR1, the R1086H mutation in
the DHPR also increases the sensitivity of the SR calcium release mechanism to
activation by both caffeine and voltage.255 The finding that MH manifests from point
44
mutations in both the skeletal muscle DHPR and RyR1, two key proteins of the EC
muscle EC coupling.
amorphous “cores” or regions in skeletal muscle that are devoid of mitochondria and
oxidative enzyme activity. In most cases, CCD has been linked to mutations in the
RyR1 gene, and these mutations are clustered in the same three regions of the RyR1
protein as those found for MH. Although certain mutations in RyR1 lead to the
coincidence of MH and CCD, some RyR1 mutations result in only increased MHS or
CCD in the apparent absence of MHS. RyR1 mutations in regions 1 and 2 that lead
leak that results in a net depletion of SR calcium stores.12,178,232,236 However, the vast
majority of CCD mutations in RyR1 are found in the C-terminal part of the protein
putative pore-lining region of the channel and appear to disrupt calcium activation79
and permeation through the channel. Because CCD mutations in the pore region of
the channel reduce calcium release during EC coupling in the absence of an effect on
effects of MH and CCD mutations in RyR1 on resting calcium, store content, and
calcium release during EC coupling are controversial. In general, mutations that lead
MH and CCD coincidence arises from RyR1 mutations that cause uncompensated SR
thus reducing calcium release during EC coupling by altering both the amount of
calcium available for release and the effect of luminal calcium on the release
mechanism. In addition, for RyR1 mutations that lead to both MH alone and
voltage), and thus, are also susceptible to uncontrolled calcium release upon exposure
to volatile anesthetics. By contrast, mutations within the pore region of RyR1 that
result in CCD in the absence of increased MHS do not increase release channel
Brody disease, first described in the literature in 1969 by Dr. Irwin A. Brody,
legs, arms, and eyelids are typically affected, and episodes are exacerbated in the
cold. The selective defect of a “relaxing factor”32 was later identified as a mutation in
the ATP2A1 gene, which encodes the skeletal muscle SERCA isoform, SERCA1.184
have been identified. It was initially believed that defects in sarcolipin regulation of
separate Brody families.163 Both point mutations and premature stop codons in
inheritance. In most cases, the mutations result in deleted functional domains that
total SERCA or SERCA1 protein levels was found between Brody disease and
stimulated ATPase activity and a significant increase in the time required for return to
in skeletal muscle of patients with Brody disease.23 Thus, SERCA1 defects result in
47
the inability of SERCA activity to keep pace with calcium release during repetitive
activity, calcium extrusion via PMCA pumps and Na+/Ca2+ exchangers, as well as
I.4. OVERVIEW
calcium levels. This control involves a complex and highly regulated balance of
calcium storage, release, and reuptake coordinated by the activity of and interactions
between a diverse set of SR calcium regulatory proteins (Figure I.2). Under basal
(calsequestrin, HRC, junctate, and SAR) set resting myoplasmic calcium levels (~50-
100 nM) well below that required for myofilament activation and also establish a
10,000-fold calcium gradient across the SR membrane. During muscle excitation, the
stimulates rapid and dynamic release of calcium stores through activation of RyR1
release channels located in the terminal cisternae of the SR. As local calcium levels
decrease within regions of release, changes in luminal calcium act to self-limit further
restoration of low resting calcium levels are then achieved primarily through ATP-
driven calcium reuptake into the SR via SERCA1 activity. In this way, the SR in
calcium gradient across the SR that is used to drive calcium release and muscle
contraction during excitation, and provides for efficient SR calcium reuptake and
the calcium storage, release, or reuptake functions of the SR calcium governor can
CRUs in adult mammalian skeletal muscle situates these organelles to interact with
the SR and affect local calcium signaling at the triad. Specifically, I hypothesize that
localized SR calcium release likely by controlling the local redox environment of the
powerful local control mechanism for integrating calcium release/reuptake and ATP
utilization during muscle contraction with ATP production and skeletal muscle
50
bioenergetics. Potentially, the efficiency of this local control mechanism would rely
organization. Therefore, the first aim of this research was to establish mitochondrial
localization in flexor digitorum brevis (FDB) fibers from young (2-4 weeks) and adult (2-
4 months) mice using confocal microscopy. The expectation was that mitochondria
mitochondrial calcium uptake should occur in adult muscle, with fully matured CRUs
intimately associated with mitochondria. In support of this, the studies cited earlier
to triads. Yet, published literature does not address how mitochondrial calcium
uptake varies with mitochondrial-SR colocalization. For this reason, the second aim
51
necessary for maximal calcium spark suppression, as well. In order to test this
assumption, the last aim of this work utilized an osmotic shock paradigm to induce
calcium sparks in intact young and adult FDB fibers. CRU-targeted mitochondria
sparks more efficiently than those not localized to CRUs would support a privileged
Figure I.1. Structural Features of Key Calcium Regulatory Proteins of the Skeletal
K) Sarcolipin.
53
Calsequestrin, triadin, junctin, junctate, RyR1, JP-1, JP-45, SERCA, and sarcolipin
are depicted using the schematics shown in Figure I.1. For simplicity, sarcalumenin
and HRC proteins are shown as gray four-point stars and white triangles, respectively,
and SERCA is shown only in the longitudinal SR. Potential interactions between
RyR1 and the α1s- and β1a-subunits of the skeletal muscle DHPR, and the C-terminus
Figure I.3. Mitochondria are Precisely Localized Adjacent to the Calcium Release
A) Confocal image of a single flexor digitorum brevis (FDB) skeletal muscle fiber
obtained from an adult mouse (4 months of age) stained with the mitochondrial-selective
dye tetramethylrhodamine ethyl ester (TMRE). TMRE fluorescence along the line of
periodicity of ~2 µm (inset). B and C) Representative low (B) and high (C) resolution
electron micrographs of FDB muscle fibers obtained from an adult mouse (4 months of
age). Mitochondria (open arrows) are aligned adjacent to the triad (arrowheads), on
the ATP synthetic activity of the F1F0-ATPase (complex V). Low levels of superoxide
anions (reactive oxygen species, ROS) are generated as byproducts of electrons being
passed to O2 at complex I and during the Q cycle (Q). OAA, oxaloacetate; Ac-CoA,
acetyl coenzyme A; CS, citrate synthase; Acon, aconitase; α-KG, α-ketoglutarate; MDH,
redox environment and local Ca2+ spark activity of the Ca2+ release unit (CRU). SR,
1.1. RATIONALE
excitation contraction (EC) coupling and the subsequent effects on the physiology of
determined the origins and maturation of this interorganelle coupling, which is the
foundation for any proposed functional interaction between the two. The goal of the
and T-tubule components of the CRU, mitochondria would lack the proper
disposition in young fibers but would gradually progress towards the precise
1.2. METHODS
URSMD. Single acutely dissociated flexor digitorum brevis (FDB) muscle fibers
59
were isolated from young (0.5 and 1 month old) and adult (2 and 4 month old)
C57/Bl6 mice as described previously.152 Briefly, FDB muscles were dissected from
mouse hind limbs in Ringer’s solution (in mM: 146 NaCl, 5 KCl, 2 CaCl2, 1 MgCl2,
shaking in a water bath for 45 minutes at 37°C. Single FDB fibers were then
liberated by trituration with glass pipettes of decreasing bore size. Fibers were plated
on glass coverslips and allowed to settle for 1-2 hours before dye loading.
Membrane Potential
fibers obtained from 0.5 and 4 month-old mice were simultaneously loaded with 100
sequentially excited using 543 nm (8X attenuation, 605/75 nm emission) and 488 nm
(4X attenuation, 515/30 nm emission) lasers, respectively. All images (512x512, 0.09
Confocal microscope (30 µm pinhole) equipped with a SuperFluor 40X 1.3 NA oil
objective.
development, FDB fibers obtained from 0.5, 1, 2, and 4 month-old mice were loaded
60
with 1.5 µM JC-1 in Ringer’s solution for 20 minutes at 37°C. With the focal plane
positioned at the center of the fiber, JC-1 was excited using a 488 nm laser (8X
aggregate (red) and JC-1 monomer (green) fluorescence was subsequently created
offline. A 20 µm line along the longitudinal axis of the fiber was drawn to generate a
mean XY ratio profile with X values representing fiber width and Y values
representing the longitudinally averaged fluorescence ratio. Because each line length
varied with fiber width, all X values were normalized to the longest X value, yielding
a percent of the fiber width for each profile. The ratio values were binned and
comparison between different fibers, images were recorded using identical laser
power, photomultiplier sensitivity, and were processed using identical values for
contrast and brightness. Images were processed and analyzed using NIH ImageJ and
deconvolution was applied to all images for display purposes. Statistical significance
determined with Kruskal-Wallis one-way ANOVA on ranks with Dunn’s post test.
regardless of ∆ψm.59 TMRE and MTG were sequentially excited using 543 nm (8X
61
FDB muscles were fixed in situ with fixative solution (3.5% glutaraldehyde in
0.1 M NaCaCo buffer, pH 7.4) at room temperature. Small bundles of fixed muscle
fibers/cells from FDB muscles were then post-fixed in 2% osmium tetroxide (OsO4)
in the same buffer for 2h and block-stained in aqueous saturated uranyl acetate.
Specimens were then dehydrated in graded ethanol and acetone, infiltrated with Epon
812-acetone (1:1) mixture for 2 h, and then embedded in Epon. Ultrathin sections
(approximately 30-40 nm) and semi-thin sections (100 nm) were cut in Leica Ultracut
R (Leica Microsystem, Austria) using a Diatome diamond knife (Diatome Ltd. CH-
2501 Biel, Switzerland) and stained in 4% uranyl acetate and lead citrate. All
sections were examined with a FP 505 Morgagni Series 268D electron microscope
(Philips) at 60kV equipped with Megaview III digital camera and Soft Imaging
System (Germany).29
CRU and mitochondrial density (and their position relative to the sarcomeres)
that were randomly collected from longitudinal sections of internal fiber areas. In
each EM image, the number of triads and mitochondria was determined, as well as
individual mitochondrion extended from one band to the other, it was counted in
62
both. CRUs and mitochondria were marked and counted in each micrograph and the
1.3. RESULTS
localization in single isolated FDB muscle fibers at early and late stages of postnatal
development (0.5 and 4 months). In each case, fibers were co-loaded with di-8-
selective fluorescent probe. Mitochondria in FDB fibers from 0.5 month-old mice are
green signal). On the other hand, a primitive T-tubule system has begun to develop
even in young 0.5 month-old fibers, as evidenced by fibers exhibiting double rows of
transverse di-8-ANEPPS staining, consistent with the known location of the T-tubule
network at the A-I junction in mammalian muscle (Figure 1.1A and B, red signal).
mitochondria in 0.5 month old FDB fibers. The absence of significant transverse
MTG and di-8-ANEPPS colocalization suggests that mitochondria are not well-
localized to the triad junction at this stage of postnatal development. Indeed, electron
are clustered beneath the sarcolemma and within the myofibrillar space (Figure 1.2A)
not typically observed in FDB fibers from adult (4 months-old) mice (Figure 1.1C
1.3.2. JC-1 Emission Distribution and Magnitude are Altered During Postnatal
Development
skeletal muscle development is also observed in FDB fibers stained with the
membrane potential across each of the developmental time points studied, in addition
adult muscle fibers, puncta of red JC-1 fluorescence were clearly visible, though
64
young fibers exhibited more puncta than adult fibers, and red JC-1 fluorescence was
typically more concentrated adjacent to the sarcolemma (Figure 1.3A, left). We also
found that the average ratio of JC-1 aggregate (red) to JC-1 monomer (green)
fluorescence (Figure 1.3A, right) was significantly (p < 0.001) greater in FDB fibers
isolated from 0.5 and 1 month-old mice (1.21 ± 0.02 and 1.58 ± 0.01 at 50% fiber
width, respectively) compared to that observed for fibers from 2 and 4 month-old
mice (0.82 ± 0.01 and 0.74 ± 0.01 at 50% fiber width, respectively) (Figure 1.3B).
The decrease in JC-1 emission ratio throughout postnatal development likely reflects
old fibers (1.5 ± 0.03, at 50% fiber width) was significantly (p < 0.001) higher than in
1.4. DISCUSSION
To date, most skeletal muscle structural studies have focused on two major
areas: 1) muscle cytoarchitecture, including the filament systems (e.g. actin, myosin,
sarcotubular system.82,90 Certainly, the focus on the former is due to its role in
placed on the sarcotubular system is best described in a review published in the early
authors write, “A nerve impulse passing along the T-tubules affects some as yet
unknown change in the SR membranes (the T-tubule does not open directly into the
lateral cisternae) that causes these membranes to lose some of the Ca2+ they have
obvious, therefore, that both the T-system and the SR system of skeletal muscle cells
have vitally important physiological roles, and the sequence of development of these
current study can be explained in the same context. Mitochondria are postulated to
play a significant role in calcium cycling at the triad during skeletal EC coupling
The close juxtaposition of mitochondria and triads in adult striated muscle has
mitochondrial disposition in skeletal muscle in the greatest detail. All muscle types
(red, white, and intermediate) exhibit pairs of mitochondria on either side of the Z-
line encircling the myofibrils. Consistent with these original findings,186 I find close
juxtaposition of mitochondria and triads at the A-I band junction in FDB fibers
isolated from adult (2-4 months of age) mice, with limited numbers of mitochondria
development. To begin, 0.5 and 1 month were chosen for study as transitional time
points during which SR and T-tubule membranes develop and mature.82,90 Though
triads are present in some neonatal skeletal muscle,82 they are few in number and are
“stretched” by subjecting the muscle fibers to osmotic shock, indicating that the
pairing (Table 1.1), concurrent with an increase in tethers, suggests that tethers help
recently identified the protein which stabilizes and maintains close endoplasmic
Thus, MFN2 is a strong candidate molecule for the skeletal muscle mitochondrial-SR
tethers. At the very least, the identification of a “tethering protein” will contribute
(e.g. FDB) with fast oxidative (e.g. extensor digitorum longus, EDL) and slow-twitch
(e.g. soleus) muscle may reveal possible mechanisms. Given that close
follow the same progression as in the FDB examined in this study. However,
68
longitudinal columns of mitochondria are common in red muscle,186 which also has
the greatest mitochondrial density of all muscle types. Perhaps, differences in the
function during muscle excitation are likely to differ significantly during skeletal
muscle development. For example, I found that the relative mitochondrial membrane
targeting to CRUs lags behind triad maturation. Thus, the degree of mitochondrial
calcium uptake, ATP production, and reactive oxygen species (ROS) generation
during muscle contraction are likely to depend on both development and maturation
with no overlapping regions (at 14,000 X), all from internal areas of the fibers.
70
Development
Confocal images of representative FDB fibers obtained from 0.5 (A and B) and 4 (C
and D) month-old mice stained for surface and T-tubule membranes (red, 10 µM di-
Positioning
month FDB fibers. EM labeling key: arrowhead, sarcolemma (SM); solid arrow, z-
small open arrows; open stars, locations between myofibrils missing mitochondria.
72
73
A) JC-1 confocal image overlay (J-aggregate [red] and JC-1 monomer [green]
right) in representative FDB fibers obtained from 0.5, 1, 2, and 4 month old mice. B)
Representative JC-1 ratio image (left) used to calculate mean (± SE) JC-1 ratio spatial
profiles (right) in FDB fibers obtained from 0.5 (filled circles), 1 (open circles), 2
(filled triangles), and 4 (open triangles) month-old mice. To determine the JC-1
profile for each ratio image, a 20 µm long line of interest was drawn across the
averaged ratio. Data from each population was normalized, binned and averaged as
described in the Methods section. 1.21 ± 0.02, n = 109 (0.5 m); 1.58 ± 0.01, n = 43 (1
m); 0.82 ± 0.01, n = 49 (2 m); and 0.74 ± 0.01, n = 75 (4 m) at 50% fiber width. *p <
Green FM (MTG) ratio spatial profiles of FDB fibers obtained from 1 (open circles)
and 2 (filled triangles) month-old mice. 1.5 ± 0.03, n = 113 (1 m); and 1.3 ± 0.02, n =
2.1 RATIONALE
block mitochondrial calcium uptake, disrupt mitochondrial function, and alter the
2.2. METHODS
Single acutely dissociated FDB muscle fibers were isolated from young (1
month old) and adult (4 month old) C57/Bl6 mice as described in Section 1.2.1.
76
Fibers were plated in a modified Ringer’s solution (in mM: 140 NaCl, 4 KCl, 1
tissue culture dishes and allowed to settle for 20 minutes before dye loading.
of Mag-Fluo-4 loaded fibers was excited at 488 nm. Fluorescence emission at 517
4/AM. Green (MTG, Fluo-4) and red (Rhod-2, TMRE) fluorophores were
sequentially excited using 488 nm (16X attenuation, 515/30 nm emission) and 543nm
(8X attenuation, 605/75 nm emission) lasers, respectively. Time series (11 frames,
256x256 pixels/frame, 0.195 µm/pixel) were acquired using a Nikon Eclipse C1 Plus
77
Confocal microscope (30 µm pinhole) equipped with a SuperFluor 40X 1.3 NA oil
evoked calcium release were elicited with high frequency (8V, 5 ms, 100 Hz)
stimulation using an extracellular agar-plug electrode placed near the fiber of interest.
To enable comparison between different fibers, images were recorded using identical
laser power, photomultiplier sensitivity, and were processed using identical values for
contrast and brightness. Image sequences were processed and analyzed offline using
along the longitudinal axis of the fiber was drawn to generate a mean XY ratio profile
with X values representing fiber length and Y values representing the longitudinally
averaged fluorescence ratio. Values for profile peaks (I-band fluorescence) and
respective values at rest. It is worth noting that image acquisition was staggered with
the beginning of each tetanus to minimize movement artifact, though the stimulus
train did continue through acquisition of the Rhod-2 images. Thus, Rhod-2 images
provided details of both mitochondrial calcium following the tetanus and the
profiles were centered over the A-band and reported as a ratio to A-band signal.
78
Wallis one way ANOVA on ranks with Dunn’s post test. For comparison of
fluorescence in the presence and absence of EGTA, a Mann-Whitney rank sum test
2.3. RESULTS
muscle fibers from young (1 month of age) and adult (4 months of age) mice using
the high-affinity (in vitro Kd ~570 nM) calcium indicator Rhod-2. Rhod-2 has
calcium uptake because its cationic nature directs dye accumulation to mitochondria.
However, several published studies of striated muscle cells report a significant level
calcium. In these studies, selective mitochondrial loading was attained by: 1) cold
the loading procedures, FDB fibers exhibited a significant level of cytosolic Rhod-2
was loaded into intact FDB skeletal muscle fibers at room temperature for a limited
79
At rest, Rhod-2 fluorescence in both young and adult FDB fibers exhibited
alternating with a brighter I-band signal (Figure 2.1). In the young fibers, Rhod-2
were clearly visible in the MTG channel. Additionally, the resting I-band Rhod-2
signal in some fibers did not colocalize with MTG fluorescence (data not shown),
diffuse I-band signal. Agents to collapse the membrane potential (e.g. FCCP, CCCP)
to inhibit mitochondrial calcium uptake were not employed to gauge the contribution
have been reported to increase resting myoplasmic calcium and alter electrically-
train of high frequency stimulation (500 ms, 100 Hz) produced a uniform increase in
A-band fluorescence (Figure 2.1 and 2.4B and D) with a heterogenous increase in I-
band fluorescence (Figures 2.1 and 2.4A). Specifically, the Rhod-2 signal in certain
80
significant mitochondrial calcium uptake. In other areas, there was a parallel increase
in both I-band and A-band fluorescence, but the I-band signal remained diffuse.
following a single high frequency train (Figure 2.4A). Calcium uptake in intermittent
well. Thus, subsequent experiments in both young and adult FDBs utilized a high
SR calcium release, and all measurements were made in areas of discrete Rhod-2
signal.
of global calcium with the low affinity indicator Mag-Fluo-4 revealed that a single
high frequency train evoked rapid SR calcium release, with an initial rise time of a
few milliseconds (3.86 ± 0.02 ms) (Figure 2.2B). There was a summation of the first
few pulses during the train such that peak calcium levels (1.26 ± 0.03 ∆F/F0) (Figure
2.2C) were reached within a few tens of milliseconds and remained high for the
following the end of the stimulation train (Figure 2.2B). Subsequent tetani elicited
largely the same response as the first train, with a tendency for decreased peak
amplitude after ~15-20 tetani (Figure 2.2D). Importantly, there was no significant
81
increase or decrease in baseline fluorescence from the first to the last tetanus (Figure
2.2E).
accumulated calcium following the first tetanus (Figure 2.4A and C). The increase in
I-band/A-band fluorescence after the first tetanus was similar between both young
and adult fibers (1.17 ± 0.04 and 1.19 ± 0.04, respectively). There was a trend for
continued increase in mitochondrial calcium uptake with the next five to ten tetani,
although the increases were not significantly different from the initial rise after the
first tetanus. Regardless, the signal remained elevated throughout tetanic stimulation
and at 1 minute following the last stimulus train (Figure 2.4A and C). On the other
hand, the Rhod-2 I-band/A-band ratio had returned near baseline levels by 20 minutes
(1.04 ± 0.01 and 1.06 ± 0.02 for 1 and 4 month fibers, respectively). Moreover, Fluo-
myoplasmic calcium was back to resting levels at 1 minute following the last tetanus
(Figure 2.3A and B). In addition, we know from the Mag-Fluo-4 studies that
myoplasmic calcium rapidly returns to baseline following each tetanus (Figure 2.2B).
Taken together, this data supports the conclusion that the increase in I-band
fluorescence relative to A-band fluorescence was also observed after a single tetanus
(1.22 ±0.09) (Figure 2.4A and E). The change in longitudinal fluorescence followed
82
a similar time course as I-band fluorescence, remaining high for the length of the
stimulation protocol and decreasing gradually to baseline following the final tetanus
(Figure 2.4E). The magnitude of change with tetanic stimulation was similar between
triadic and longitudinal mitochondria. However, direct comparisons were not made
between the two as clear longitudinal mitochondria were rarely in the same focal
young or adult FDB fibers during the tetanic stimulation protocol (Figure 2.3A and
C). In other words, mitochondrial calcium uptake in my experiments did not induce a
TMRE dye and response to a single twitch at the end of the tetanic stimulation
protocol indicated that both the mitochondrial and sarcolemmal potentials were intact
membrane potential would have escaped detection with the relatively slow and
stimulation, Rhod-2 fluorescence was measured in fibers loaded with the calcium
83
chelator EGTA-AM. EGTA has relatively slow kinetics, and thus, is unable to buffer
calcium within the microdomain (i.e. nanometers from the calcium release source). If
the mitochondrial calcium uptake mechanism is located closer than a couple hundred
experiments, one criterion for choosing fibers to image was twitch with a single, low
µM EGTA-AM, the peak myoplasmic calcium level after a single tetanus was
significantly reduced (1.26 ± 0.03 ∆F/F0 and 0.75 ± 0.02 ∆F/F0 in the absence and
presence of EGTA, respectively) (Figure 2.2C). The initial rise phase and the decay
Similar to resting fibers in the absence of EGTA, both young and adult fibers
exhibited transverse rows of diffuse I-band fluorescence. Within the first tetanus,
mitochondrial calcium uptake peaked after the first few tetani, remained elevated for
the duration of tetanic stimulation, and gradually decreased to resting levels after the
last tetanus (Figure 2.4C). In young fibers, mitochondrial calcium uptake in neither
triadic mitochondria (Figure 2.4C) nor longitudinal mitochondria (Figure 2.4E) was
triadic mitochondria of adult fibers after the first tetanus was significantly (p < 0.01)
lower in the presence of EGTA (1.19 ± 0.04 and 0.96 ± 0.03, respectively). This
decrease was moderately significant (0.05 > p > 0.01) following subsequent trains.
84
2.4. DISCUSSION
mitochondria in FDB fibers did not take up detectable calcium. Therefore, I utilized
significant mitochondrial calcium uptake. Given that the typical nerve impulse
mitochondrial calcium uptake in intact murine FDB fibers.140 In their study, single
intervals, and finally at 2.5 s intervals until tetanic force declined to 40% of rest (up to
during and after the fatiguing stimulation. Yet, the authors reported that
upon this study, I began stimulating fibers with 350 ms trains at 70 Hz at 2.5 s
calcium within the first few trains. Ultimately, after some optimization of the
uptake, I set stimulation at 500 ms, 100 Hz trains given every 2.5 s for a total of 40
and the published study are minor. Consequently, this experimental variation cannot
rest. On the other hand, Lännergren et al. described defined transverse rows of Rhod-
signal is not unlike the discrete transverse mitochondrial fluorescence I see after
tetanic stimulation. Conceivably, variations between the mouse strains used or the
(see Section 1.2.1) could result in the differing Rhod-2 signal at rest. Regardless, the
data suggests that mitochondrial resting free calcium was elevated in the Lännergren
tetanic train, but not a single twitch, implies that sustained SR calcium release is
length and timing. Essentially, this would enable selective titration of the duration
and magnitude of SR calcium release. This type of systematic study will provide
young versus adult triadic mitochondria. In the present study, no such differences
calcium release would provide evidence for (or against) a threshold calcium for
mitochondrial calcium uptake. If a threshold exists, it may differ between young and
postnatal development.
would, instead, indicate that mitochondria are responding to global calcium. The
supports the former. However, the trend in adult fibers for reduced mitochondrial
global calcium. Perhaps, simply dampening myoplasmic calcium with a low EGTA
signaling during EC coupling. In fact, the true test of highly privileged mitochondrial
calcium uptake following SR release is in the inability of the fast calcium buffer
allow it to confine the calcium microdomain, chelating calcium within a few tens of
mitochondrial calcium uptake could reveal important information about the proximity
mitochondria within the same muscle fiber would provide more definitive proof of
calcium uptake, whereas only BAPTA would effectively reduce triadic mitochondrial
longitudinal and triadic mitochondria, I was unable to resolve differences between the
mitochondria prevented direct comparisons within the same fiber in this limited set
data set. Not to mention, the longitudinal clusters were often concentrated beneath the
young and adult fast-twitch skeletal muscle fibers accumulate calcium with tetanic
oxidative (e.g. extensor digitorum longus) and slow-twitch muscle fibers, which are
in adult muscle. These studies, when paired with experimental use of calcium
chelators and high resolution ultrastructure studies, will provide more information
different fibers types will give insight into the importance of mitochondrial calcium
fluorescence in FDB skeletal muscle fibers obtained 4 month-old mice before and
after electrical stimulation. B) MTG (left) and Rhod-2 (right) fluorescence along the
(left) at 2.5 s intervals for a total of 40 tetani (right). Filled squares indicate the time of
in FDB fibers obtained from 1 month-old mice in the absence (upper) and presence
transients in the absence and presence of EGTA. D and E) Mean (± SE) peak
fluorescence (∆F/F0) (D) and resting fluorescence (F0) (E) of each tetanic train.
91
Stimulation
FDB fibers obtained from 1 (left) and 4 (right) month-old mice before tetanic
stimulation; after 1 and 40 tetani; and 1 and 20 minutes following the last (40th)
from 1 (upper) and 4 (lower) month-old mice before tetanic stimulation; after 1, 5,
10, and 40 tetani; and 1 and 20 minutes following the last (40th) tetanus. Bars: 5 µm.
Normalized A-band fluorescence (Ft/Ft=0) (B) and Mean (± SE) normalized triadic
stimulation at 1 (circles) and 4 (triangles) months in the absence (filled symbols) and
RELEASE
95
3.1. RATIONALE
Accordingly, the present study compared the incidence and properties of osmotic
3.2. METHODS
Single acutely dissociated FDB muscle fibers were isolated from young (0.5
and 1 month old) and adult (2 and 4 month old) C57/Bl6 mice as described in Section
1.2.1. Fibers were plated in a modified Ringer’s solution (in mM: 140 NaCl, 4 KCl, 1
96
glass cover slips and allowed to settle for 20 minutes before dye loading.
fibers were immersed in hypertonic Ringer (in mM: 140 NaCl, 4 KCl, 1 MgSO4, 50
CaCl2, 5 NaHCO3, 10 Glucose, 10 HEPES; pH 7.4, ~450 mOsm). With the focal
plane centered along fiber periphery, a series of line scans (1024 lines, 47.7 µm/line,
2.12 s/line) parallel to the long axis of the fiber (i.e. spanning multiple sarcomeres)
were acquired. To enable comparison between different fibers, images were recorded
using identical laser power, photomultiplier sensitivity, and were processed using
identical values for contrast and brightness. Images were analyzed offline using
Detection limits were set broadly to identify events with amplitudes (A) ranging from
0.2 to 6 ∆F/F, full width at half-maximal amplitude (FWHM) from 0.5 to 50 µm, rise
time (R) from 0.5 to 100 ms, and full duration at half-maximal amplitude (FDHM)
from 0.5 to 1000 ms. Spark mass was calculated as A*1.206*FWHM3, according to
3.3. RESULTS
in an Age-Dependent Manner.
facilitate inhibition of local calcium release at the triad. To test this hypothesis in
single intact FDB fibers, the incidence and properties of calcium sparks, revealed by
skeletal muscle development (0.5, 1, 2, and 4 months) that were shown in Chapter 1
exposure to hypertonic Ringer, FDB fibers of all ages exhibited spontaneous local
calcium release events (Figure 3.1). Activity persisted for the hour of
the fibers. Because the external solution contained calcium, I cannot exclude the
possibility that some sparks were in fact calcium influx through membrane tears
caused by osmotic shock. However, the quantal nature of the observed events argues
against membrane tears. Furthermore, triads are found within a few microns of the
sarcolemma (Figure 3.5, Table 3.2) and in the few number of line scans where a clear
Fluo-4 banding pattern was observed, events did occur in the expected triadic position
(~20-30 ms) local calcium release events (classic Ca2+ sparks) were observed (Figure
3.1A and B) with time courses similar to the calcium sparks identified in other
98
(>100 ms), analogous to so-called calcium “bursts”250 were also detected (Figure
3.1A and C). To examine a broad range of events, an upper limit was set on both rise
The events were segregated into short (rise time < 30 ms and/or FDHM < 100
ms) and long (rise time > 30 ms and/or FDHM > 100 ms) events. The percentage of
short events was greatest at 0.5 month (61% of total events) and smallest at 1 month
(41%). The total population of events was split 50/50 between short and long at 2
and 4 months (Figure 3.2A). However, the parameters varied with age within the
population of short events to the same degree as within the population of long events.
detected events, and all events from here on will be generally referred to as “Ca2+
sparks”.
Analysis of average (± SE) Ca2+ spark amplitude, spatial width, and mass are
presented in Figure 3.3. Mean peak Ca2+ spark amplitude was significantly reduced
in 0.5 month fibers (0.34 ± 0.01 ∆F/F) compared to events observed in fibers isolated
from 1 (0.41 ± 0.01 ∆F/F), 2 (0.40 ± 0.01 ∆F/F), and 4 (0.41 ± 0.01 ∆F/F) month-old
mice, which exhibited similar peak amplitudes (Figure 3.3A, Table 1). Likewise,
events at 1, 2, and 4 months exhibited similar FDHM (48.26 ± 2.67 ms, 47.26 ± 2.08
ms, and 49.62 ± 1.83 ms, respectively) which were greater than that observed at 0.5
months (33.66 ± 1.94 ms) (Figure 3.2C, Table 1). A similar trend (an increase during
early postnatal development) was not observed for either Ca2+ spark rise time or
99
FWHM. Rather, rise time and FWHM peaked at 1 month and then were reduced
thereafter (Table 3.1). Specifically, rise time at 1 month (38.22 ± 1.04 ms) was
significantly longer than 0.5, 2, and 4 month events (29.54 ± 0.79 ms, , 33.28 ± 0.71
ms, and 33.87 ± 0.54 ms, respectively) (Figure 3.2B, Table 3.1). In addition, the
FWHM was greater at 1 month (1.48 ± 0.02 µm) than that observed at 0.5, 1 or 2
months (1.35 ± 0.02 µm, 1.36 ± 0.02 µm, and 1.30 ± 0.01 µm, respectively) (Figure
greatest at 1 month (1.8 ± 0.2 µm3, 2.7 ± 0.2 µm3, 2.1 ± 0.1 µm3, and 1.8 ± 0.1 µm3
Finally, raw Ca2+ spark frequency followed the identical pattern of both
amplitude and FDHM, being lowest at 0.5 month (76.12 ± 3.72 events/min) (Figure
3.4A). Higher Ca2+ spark frequencies observed in FDB fibers of 1, 2 and 4 month-old
mice (115.02 ± 9.05 events/min, 110.14 ± 5.70 events/min, and 107.81 ± 4.27
events/min, respectively) (Figure 3.4A, Table 3.1 ) most likely reflect an increase in
the CRU number and density that occurs during postnatal development (Tables 1.1
determined from EM analysis (Table 1.1), normalized Ca2+ spark frequency was
3.1). Fibers isolated from 0.5 month-old mice exhibited the greatest frequency of
was a clear linear correlation between the tether frequency29 and Ca2+ spark frequency
FDB fibers isolated from 0.5 month-old mice upon exposure to hypertonic Ringer
(Figure 3.5C) indicate that osmotic shock treatment used to trigger Ca2+ sparks does
revealed no gross ultrastructural changes in the interior of the fiber following osmotic
shock fibers of either 0.5 or 4 month FDB fibers in the interior of the fiber (data not
shown). On the other hand, osmotic shock did induce minor changes in the
subsarcolemmal region (within 1 µm of the surface membrane) (Figure 3.5A and B).
(T-tubules) and caveolae just beneath the surface membrane, the extent of which was
variable within a given field of view. Importantly, osmotic shock treatment produced
similar changes in FDB fibers from both 0.5 and 4 month old mice. The only major
difference between the two ages was a higher percentage of fibers with
1.29
101
3.4. DISCUSSION
spontaneous, spatially restricted calcium release events have been detected in many
different cell types including intact smooth muscle181 and amphibian skeletal
cardiac and amphibian skeletal muscle represent elementary units of the global
calcium from single calcium spark properties is made possible by the homogeneity of
sparks and the relatively high frequency of spontaneous events in cardiac and
skeletal muscle are extremely rare,60 though they can be unmasked by sarcolemmal
exercise. Currently, there are three proposed mechanisms to explain the absence or
receptor isoform, RyR1. On the other hand, amphibian skeletal muscle expresses
102
both RyR1 (RyRα) and RyR3 (RyRβ) isoforms, suggesting that RyR3 accounts for
sparks are readily detectable in mammalian diaphragm, which expresses both RyR1
expression during the transition of myotubes into myofibers, though Ca2+ sparks are
also observed in RyR3-null myotubes,254 indicating that RyR3 supports, but is not
skeletal muscle may ensure exclusive control of SR calcium release by the voltage-
sensor (i.e. DHPR), thereby preventing spontaneous calcium release events in resting
skeletal muscle cells that lack RyR3. In amphibian muscle, voltage-induced calcium
release (VICR) initiated by the DHPR can activate lone RyRs that are not coupled to
amphibian skeletal muscle. On the other hand, CICR is not a significant means of
coupling, and thus, spontaneous calcium release is rare. Additional evidence that the
types, the T-tubular network lacks the structure of adult muscle. Further, osmotic
muscle may physically disrupt triads, and thus, the tight DHPR-RyR1 inhibitory
calcium sparks in adult mammalian skeletal muscle, does not compromise DHPR-
RyR1 coupling. Thus, while the DHPR-RyR1 interaction likely contributes to spark
the DHPR plays a critical role in reducing the rate of spontaneous Ca2+ spark
generation.268
sparks in mammalian skeletal muscle. Most recently, Isaeva et al.117,118 proposed that
calcium sparks. Their initial studies clearly demonstrated that Ca2+ spark frequency
respiration by providing substrates for the tricarboxylic acid (TCA) cycle reversibly
mitochondrial matrix calcium (Ca2+ marks) elicited by calcium sparks have been
104
possible that mitochondrial calcium uptake directly inhibits junctional CICR and thus
mechanisms for spark inhibition must be considered. For example, mitochondria may
indirectly regulate SR calcium release via their capacity to generate ATP and ROS.
redox potential, and Ca2+ spark development in permeabilized skeletal muscle fibers.
In their study, the onset of spark activity following permeabilization was slowest in
poor glycolytic fibers. Mitochondrial substrates and ROS scavengers also delayed the
domain that facilitates both direct calcium buffering of sparks by mitochondria and
105
predicted to exhibit a more limited capacity to suppress calcium sparks due to their
more remote distance from sites of release. Indeed, I find a progressive decrease in
spark frequency per CRU density from 0.5 to 4 months, a time during which we know
that both the number of mitochondria and percent of mitochondria localized to the I-
band increases (see Chapter 1). The decrease in the propensity for a given CRU to
For instance, the increase in raw frequency of events from 0.5 to 1 month reflects a
burst in CRU density during this time. Further, higher mean peak spark amplitude
and longer FDHM at 1, 2, and 4 months likely results from CRU maturation. That
since the greatest overall triadic structural changes occur between 0.5 and 1 month.
Thus, at a certain point of development (~1 month), CRUs are, in effect, equivalent in
month postnatal stage represents a critical transitional time point. Specifically, rise
time is slowest and both FWHM and spark mass are maximal at 1 month.
targeting, resulting in high amplitude and long events with a large spatial spread.
106
Yet, spark frequency per CRU continues to decrease in parallel with increased
increases further from 2 to 4 months, Ca2+ sparks are more spatially restricted and
less frequent than that observed in younger fibers. It will be important for future
mitochondrial and Ca2+ spark location in the same cell. For example, do areas of
repeated spark activity correspond to areas that lack mitochondrial targeting? Is there
That is, do briefer sparks occur in regions containing triadic mitochondria, whereas
Clearly, my results pose many intriguing questions with regard to the relationship
Therefore, mitochondrial calcium uptake and respiration are likely to vary during
107
mitochondrial function triggered by osmotic shock that might alter calcium spark
with osmotic shock, future work utilizing ROS indicators could be used to test for
Osmotic shock-induced calcium sparks may also involve factors other than
observed following osmotic shock hint at other possible mechanisms of calcium spark
interaction between caveolin-3 and nNOS (neuronal nitric oxide synthase) results in a
nitrosylation could result, which has recently been shown to sensitize RyR1 to
positioned to counter this insult, they may not be sufficient to suppress sparks
altogether, particularly at the periphery where caveoli are located. More importantly,
inhibition could provide insight into more physiological local ROS/RNS signaling
between mitochondria and the SR and the role of this process in some forms of muscle
disease.
109
All acquired X-t line scan images were processed with CaSparks automated detection
densities reported in Table 1.1. * p < 0.01 for comparisons with 4 months.
110
Measurements for a single set electron micrograph of FDB fibers isolated in isotonic
solution from 0.5 and 4 month-old mice. Mitochondria and CRUs quantified as in
Section 1.2.3.
111
Figure 3.1. Osmotic Shock-Induced Ca2+ Sparks and Bursts in FDB Fibers During
Postnatal Development
A) Representative X-t line scan images of Ca2+ sparks in 0.5, 1, 2, and 4 month FDB
fibers. Images were processed using a Gaussian smoothing function and median
marked event in A. B) Short events with time courses less than 50 ms. C) Long
events with time courses greater than 50 ms. Amplitude, rise time, and FDHM (A, R,
in an Age-Dependent Manner
A) Percent of short (rise time < 30 ms and/or FDHM < 100 ms, black) and long (rise
time > 30 ms and/or FDHM > 100 ms, white) events. B) Mean (± SE) rise time of
total population of events (short and long). C) Mean (± SE) FDHM of total
population of events (short and long). Statistical significance determined with Fisher
Exact text (A) and Kruskal-Wallis one way ANOVA on ranks with Dunn’s post test
Figure 3.3. Ca2+ Spark Mass Peaks at 1 Month During Postnatal Development
A) Mean (± SE) peak Ca2+ spark amplitude. B) Mean (± SE) FWHM. C) Mean (±
SE) Ca2+ spark mass calculated as: amplitude*1.206*FWHM3. * p < 0.01 compared
with 4 months.
114
Figure 3.4. Osmotic Shock-Induced Ca2+ Spark Frequency per CRU Density
Mean (± SE) raw frequency of Ca2+ sparks (A) and Ca2+ spark frequency normalized
to CRU density (see Table 1.1) (B). * p < 0.01 compared with 4 months. C)
Correlation between normalized Ca2+ spark frequency and tether density (0.8
tethers/100 µm2, 4.0 tethers/100 µm2, 10.6 tethers/100 µm2, and 15.6 tethers/100 µm2
A and B) Electron micrographs (longitudinal sections) of 0.5 (A) and 4 (B) month FDB
fibers immersed in isotonic (upper) or hypertonic (lower) solution before fixation. Left
panels depict subsarcolemmal regions. Right panels depict T-tubule profiles within
single 4 month FDB fiber before (left) and after (right) exposure to hypertonic solution.
116
DISEASE
117
4.1. RATIONALE
bioenergetics and redox control. Defects in this local control mechanism could lead
to excessive mitochondrial ROS production that disrupts the cellular redox state in a
associated with Central Core Disease (CCD) reflects defective SR calcium release
potential in skeletal muscle fibers from knock-in mice heterozygous for a mutation
(Y522S) in RyR1 that, in humans, gives rise to Malignant Hyperthermia (MH) and
CCD coincidence. The expectation was that mitochondrial integrity (i.e. localization,
mice, as a result of the aberrant calcium release and elevated ROS/RNS production
4.2. METHODS
the URSMD (Section 1.2.1). Single acutely dissociated FDB muscle fibers were
isolated from young (0.5 and 1 month-old) and adult (2 and 4 months-old)
Membrane Potential
membrane potential during postnatal development, FDB fibers obtained from 0.5, 1,
2, and 4 month-old mice were loaded with 1.5 µM JC-1 in Ringer’s solution for 20
minutes at 37°C. Fibers were imaged and analyzed offline as described in Chapter 1.
with Dunn’s post test or a Mann-Whitney rank sum test for pairwise comparison
where appropriate.
adult (8 month) wild-type (WT) and YS mice were loaded with the potentiometric
focal plane positioned at the center of the fiber, TMRE was excited using a 543 nm
image (n = 4). A 5 µm line along the longitudinal axis of the fiber was drawn to
length and Y values representing the longitudinally averaged fluorescence ratio. The
hyaluronidase (15 µl, 0.4 U/µl). One hour later, 80 µg of MT-cpYFP in a total
was then electroporated using electrodes placed perpendicular to the long axis of the
muscle. Electroporation parameters were 100 V/cm, 20-ms duration, and 20 pulses
delivered at 1 Hz. One-week later, single muscle fibers from electroporated FDB
excited using a 488 nm (32X attenuation, 515/30 nm emission) laser. Time series
(100 total frames, ~1 frame/s) were acquired using a Nikon Eclipse C1 Plus Confocal
microscope (30 µm pinhole) equipped with a SuperFluor 40X 1.3 NA oil objective
(Nikon Instruments Inc.). Each acquired image was 256 pixels x 256 pixels (0.2
µm/pixel).
120
4.3. RESULTS
localization in single FDB muscle fibers isolated from heterozygous Y522S (YS)
interior of the fiber. Conversely, the majority of mitochondria in FDB fibers isolated
On the other hand, JC-1 ratios of YS fibers differed significantly from WT,
both in the pattern of change throughout development and the mean values (Figure
4.1B). In contrast to WT fibers, the average ratio of JC-1 aggregate (red) to JC-1
monomer (green) fluorescence (Figure 4.1A, right) was the highest in FDBs isolated
from 2 month-old mice (1.37 ± 0.01 at 50% fiber width), followed by 1 (1.20 ± 0.02),
0.5 (1.05 ± 0.02), and 4 months,(0.84 ± 0.01) (Figure 4.1B). Additionally, the mean
121
JC-1 ratios of 0.5 and 1 month YS fibers were smaller than the corresponding value
observed in WT fibers (1.21 ± 0.02 and 1.58 ± 0.01 for 0.5 and 1 month,
respectively), whereas the ratios of 2 and 4 month YS fibers were greater than WT
fibers (0.82 ± 0.01 and 0.74 ± 0.01 for 2 and 4 months, respectively).
of adjacent I-bands in single isolated FDB muscle fibers from 8 month-old WT and
YS mice (Figure 4.2A and B). With this approach, a predominance of areas lacking
yield information about sarcomeric spacing. TMRE was chosen as the mitochondrial
TMRE fluorescence was quite heterogenous, though the degree of heterogeneity was
the same for both genotypes. For example, TMRE often exhibited punctuate
fluorescence and a number of regions were void of TMRE signal. Several areas of
122
each fiber were sampled at random, including these areas of “irregular” fluorescence,
middle 50th percentile of values falls within the range of normal sarcomere length (~2
µm). The mean I-band-to-I-band distance was not significantly different between WT
and YS fibers (1.89 ± 0.02 µm and 1.87 ± 0.01 µm, respectively) (Figure 4.2C).
the I-band, positioned adjacent to CRUs in both WT and YS fibers (Figure 4.3).
a regular rounded or slightly elongated shape, with defined lamellar cristae and an
WT fibers exhibited membrane myelin figures and disrupted cristae (Figure 4.3A,
panel 4). However, swollen mitochondria with a translucent matrix (Figure 4.3B,
panels 2-4) were frequently observed in YS muscle along with normal mitochondria
cristae. Damaged external membranes and vacuole formation were common, as well
(Figure 4.3, panels 2-4). It is possible that these structural abnormalities result in
both WT and YS fibers, time lapse imaging revealed transient increases in MT-
cpYFP fluorescence with a fairly large amplitudes (1.4 F/F0 and 1.5 F/F0,
respectively) (Figure 4.4B and C). The flashes were relatively slow, lasting ~ 10 s.
Yet, the flash in the WT fiber appeared to originate from a single mitochondrion,
detection of superoxide flashes in single isolated FDB fibers and suggest that
fibers of YS mice.
4.4. DISCUSSION
within muscle fibers that are devoid of mitochondrial enzyme activity. Though the
as the primary cause of the disease. CCD mutations in RyR1 fall into one of two
terminal/central mutations, also associated with MH, which result in leaky channels
with enhanced sensitivity to activation by both voltage and RyR1 agonists.76 Chelu et
al. have generated a knock-in mouse model of one particular N-terminal mutation
extensively.48,81 Mice heterozygous for the YS mutation in RyR1 survive and breed,
mutation renders RyR1 more sensitive to activation by ligands (e.g. caffeine) and the
elevated temperatures, additional calcium release through the sensitized channel feeds
mitochondria.
125
could potentially accumulate more calcium than in WT, altering respiration and other
membrane potential I observe between WT and YS FDB fibers (compare Figure 1.3B
activity.34 Further, calcium could directly stimulate neuronal nitric oxide synthase
(nNOS), generating NO·, which would indirectly increase ROS by inhibiting complex
IV, and could generate RNS by reacting with ROS. As a result, elevated ROS and
RNS levels in heterozygous tissue could contribute to RyR1 modification and directly
processes such as contraction and SR calcium reuptake via SERCA. To maintain the
functional integrity of other mitochondria and preserve gross muscle function, the
126
give rise to the focal regions of damaged mitochondria seen in electron micrographs
and the network superoxide flash detected with the MT-cpYFP imaging (Figure
RyR1 mutations that result in MH/CCD coincidence. Mutations that result in CCD
alone, as discussed previously, have the opposite effect on RyR1, namely inhibiting
calcium permeation. Yet, in both cases, central cores are a common feature. This
suggests that the causative factor in core development may be any type of disruption
in calcium homeostasis, rather than the absolute calcium levels of the myoplasmic
resulting in a decrease in ROS production. In this case, the decrease in ROS would
imbalance of calcium, as well as ROS and ATP production are likely to lead to
cores.
127
Figure 4.1 JC-1 Emission Ratio is Altered During Postnatal Development in Y522S
Knock-In Mice
A) JC-1 confocal image overlay (J-aggregate [red] and JC-1 monomer [green]
fluorescence, left) and false-colour JC-1 ratio images (J-aggregate/JC-1 monomer, right)
Y522S knock-in mice. B) Mean (± SE) JC-1 ratio spatial profiles of FDB fibers obtained
from 0.5 (filled circles), 1 (open circles), 2 (filled triangles), and 4 (open triangles)
month-old mice. 1.05 ± 0.02, n = 26 (0.5 m); 1.20 ± 0.02, n = 30 (1 m); 1.37 ± 0.01, n
= 31 (2 m); 0.84 ± 0.01, n = 30 (4 m) at 50% fiber width. *p < 0.01 compared with 4
Figure 4.2. Mitochondrial Disposition and Sarcomere Spacing in Adult FDB Fibers
A) Confocal images of single adult (8 months-old) FDB fibers from obtained from
wild-type (WT, upper) and heterozygous Y522S knock-in (YS, lower) mice stained
for nuclei (blue, 2.5 µg/ml Hoechst 34580) and mitochondria (red, 10 nM TMRE).
B) TMRE fluorescence along the lines of interest marked in (A). Arrow heads denote
percentile; left box boundary, 25th percentile; solid line, median; dotted line, mean;
right box boundary, 75th percentile; right whisker, 90th percentile. Break in X-axis
from 3 to 8 microns. 1.89 ± 0.02 µm, n = 3442 measurements (WT); and 1.87 ± 0.01
Knock-In Mice
fibers obtained from 2 month-old wild-type (WT, A) and heterozygous Y522S knock-
in (YS, B) mice. A) Panels 1-4: Typical WT mitochondria are regularly shaped with
dense matrices and lamellar cristae. B) Panel 1: Regularly shaped mitochondria are
membranes (filled arrows), and vacuoles (filled stars) are also readily observed.
130
cpYFP (left) that has been costained for mitochondria with MitoTracker® Red
superoxide flashes in WT (B) and heterozygous Y522S (YS) knock-in (B) FDB
CONCLUSIONS
132
The close apposition of mitochondria and CRUs in adult skeletal muscle has
and mitochondria has not been investigated in any great detail. Accordingly, this
negative membrane potential than in adult FDB fibers, suggesting that mitochondrial
FDBs are exclusively positioned adjacent to CRUs on either side of the Z-line.
In their location adjacent to CRUs at the A-I band junction, mitochondria are
aptly positioned for sequestering calcium release from the SR during EC coupling.
Indeed, this work clearly demonstrates that triadic mitochondria in FDB fibers
133
development (see Chapter 2). Likely, this reflects a through-space coupling rather
than direct “calcium tunneling”212 because mitochondria are positioned on the side of
the SR terminal cisternae opposite to RyR1 feet (see Chapter 1). We have reported
the average distance between RyR1 feet and the outer mitochondrial membrane
close to the effective working distance for the slow calcium chelator, EGTA.179
Therefore, the tendency for EGTA to reduce mitochondrial calcium uptake observed
in adult FDB fibers is not surprising. Additional experiments may reveal significance
mitochondrial calcium uptake was not observed during a single twitch. Moreover,
following SR calcium release calls into question the source of mitochondrial calcium
uptake and whether close SR-mitochondrial juxtaposition is essential for the process.
In short, determining the conditions under which mitochondria take up calcium and
the dependence upon mitochondrial position with respect to triads will have important
ATP production, and thus, lessen the energy deficit during times of vigorous muscle
activity. Yet, previous studies have shown that the magnitude of mitochondrial
both rely heavily upon oxidative phosphorylation. Therefore, future studies require
with mitochondrial inhibition of local SR calcium release. Further, the data provides
clues into the possible mechanism of calcium spark unmasking by osmotic shock,
namely the disruption of caveolae may trigger downstream redox signaling events.
Consequently, osmotic shock could prove to be a useful method for the study of local
potentially lead to excessive mitochondrial ROS production that disrupts the cellular
muscle cores lacking mitochondrial enzyme activity, are characteristic of the skeletal
muscle disorder Central Core Disease (CCD). However, the mechanistic link
understood. Therefore, the final section of the present work began by characterizing
there are significant areas of swollen and damaged mitochondria. The focal nature of
ionic and metabolic homeostasis, likely triggered by the aberrant calcium release (or
leak) from mutant release channels. Clearly, defining the precise mechanism by
demands further investigation and will provide important information regarding the
BIBLIOGRAPHY
6. Andrade FH, Reid MB, Allen DG, Westerblad H. Effect of hydrogen peroxide
and dithiothreitol on contractile function of single skeletal muscle fibres from
the mouse. J Physiol 1998;509:565-575.
12. Avila G, Dirksen RT. Functional effects of central core disease mutations in
the cytoplasmic region of the skeletal muscle ryanodine receptor. J Gen
Physiol 2001;118:277-290.
14. Avila G, Lee EH, Perez CF, Allen PD, Dirksen RT. FKBP12 binding to RyR1
modulates excitation-contraction coupling in mouse skeletal myotubes. J Biol
Chem 2003;278:22600-22608.
15. Avila G, O'Connell KM, Dirksen RT. The pore region of the skeletal muscle
ryanodine receptor is a primary locus for excitation-contraction uncoupling in
central core disease. J Gen Physiol 2003;121:277-286.
16. Avila G, Dirksen RT. Rapamycin and FK506 reduce skeletal muscle voltage
sensor expression and function. Cell Calcium 2005;38:35-44.
18. Balog EM, Gallant EM. Modulation of the sarcolemmal l-type current by
alteration in SR Ca2+ release. Am J Physiol 1999;276:C128-C135.
19. Baylor SM. Calcium sparks in skeletal muscle fibers. Cell Calcium
2005;37:513-530.
20. Beard NA, Sakowska MM, Dulhunty AF, Laver DR. Calsequestrin is an
inhibitor of skeletal muscle ryanodine receptor calcium release channels.
Biophys J 2002;82:310-320.
21. Beard NA, Laver DR, Dulhunty AF. Calsequestrin and the calcium release
channel of skeletal and cardiac muscle. Prog Biophys Mol Biol 2004;85:33-
69.
22. Beard NA, Casarotto MG, Wei L, Varsanyi M, Laver DR, Dulhunty AF.
Regulation of ryanodine receptors by calsequestrin: Effect of high luminal
Ca2+ and phosphorylation. Biophys J 2005;88:3444-3454.
23. Benders AA, Veerkamp JH, Oosterhof A, Jongen PJ, Bindels RJ, Smit LM,
Busch HF, Wevers RA. Ca2+ homeostasis in Brody's disease. A study in
skeletal muscle and cultured muscle cells and the effects of dantrolene an
verapamil. J Clin Invest 1994;94:741-748.
139
25. Beurg M, Ahern CA, Vallejo P, Conklin MW, Powers PA, Gregg RG,
Coronado R. Involvement of the carboxy-terminus region of the
dihydropyridine receptor α1a subunit in excitation-contraction coupling of
skeletal muscle. Biophys J 1999;77:2953-2967.
26. Beurg M, Sukhareva M, Ahern CA, Conklin MW, Perez-Reyes E, Powers PA,
Gregg RG, Coronado R. Differential regulation of skeletal muscle l-type Ca 2+
current and excitation-contraction coupling by the dihydropyridine receptor
α subunit. Biophys J 1999;76:1744-1756.
27. Bhat MB, Zhao J, Zang W, Balke CW, Takeshima H, Wier WG, Ma J.
Caffeine-induced release of intracellular Ca2+ from chinese hamster ovary
cells expressing skeletal muscle ryanodine receptor. Effects on full-length and
carboxyl-terminal portion of Ca2+ release channels. J Gen Phys 1997;110:749-
762.
30. Brandl CJ, deLeon S, Martin DR, MacLennan DH. Adult forms of the Ca 2+
ATPase of sarcoplasmic reticulum: Expression in developing skeletal muscle.
J Biol Chem 1987;262:3768-3774.
31. Brandt NR, Caswell AH, Wen SR, Talvenheimo JA. Molecular interactions of
the junctional foot protein and dihydropyridine receptor in skeletal muscle
triads. J Membr Biol 1990;113:237-251.
34. Brookes PS, Yoon Y, Robotham JL, Anders MW, Sheu SS. Calcium, ATP,
and ROS: A mitochondrial love-hate triangle. Am J Physiol Cell Physiol
2004;287:C817-C833.
35. Brown LD, Rodney GG, Hernandez-Ochoa E, Ward CW, Schneider MF. Ca2+
sparks and t tubule reorganization in dedifferentiating adult mouse skeletal
muscle fibers. Am J Physiol Cell Physiol 2007;292:C1156-C1166.
38. Buck E, Zimanyi I, Abramson JJ, Pessah IN. Ryanodine stabilizes multiple
conformational states of the skeletal muscle calcium release channel. J Biol
Chem 1992;267:23560-23567.
41. Callaway C, Seryshev A, Wang JP, Slavik KJ, Needleman DH, Cantu C, Wu
Y, Jayaraman T, Marks AR, Hamilton SL. Localization of the high and low
affinity [3H]ryanodine binding sites on the skeletal muscle Ca2+ release
channel. J Biol Chem 1994;269:15876-15884.
42. Campbell KP, Knudson CM, Imagawa T, Leung AT, Sutko JL, Kahl SD,
Raab CR, Madson L. Identification and characterization of the high affinity
[3H]ryanodine receptor of the junctional sarcoplasmic reticulum Ca 2+ release
channel. J Biol Chem 1987;262:6460-6463.
47. Catterall WA. Structure and regulation of voltage-gated Ca2+ channels. Annu
Rev Cell Dev Biol 2000;16:521-555.
48. Chelu MG, Goonasekera SA, Durham WJ, Tang W, Lueck JD, Riehl J, Pessah
IN, Zhang P, Bhattacharjee MB, Dirksen RT, Hamilton SL. Heat- and
anesthesia-induced malignant hyperthermia in an RyR1 knock-in mouse.
FASEB J 2006;20:329-330.
50. Chen W, Ruell PA, Ghoddusi M, Kee A, Hardeman EC, Hoffman KM,
Thompson MW. Ultrastructural changes and sarcoplasmic reticulum Ca2+
regulation in red vastus muscle following eccentric exercise in the rat. Exp
Physiol 2007;92:437-447.
51. Cheng H, Lederer WJ, Cannell MB. Calcium sparks: Elementary events
underlying excitation-contraction coupling in heart muscle. Science
1993;262:740-744.
52. Cherednichenko G, Hurne AM, Fessenden JD, Lee EH, Allen PD, Beam KG,
Pessah IN. Conformational activation of Ca2+ entry by depolarization of
skeletal myotubes. Proc Nat Acad Sci 2004;101:15793-15798.
53. Cheung A, Dantzig JA, Hollingworth S, Baylor SM, Goldman YE, Mitchison
TJ, Straight AF. A small-molecule inhibitor of skeletal muscle myosin II. Nat
Cell Biol 2002;4:83-88.
54. Ching LL, Williams AJ, Sitsapesan R. Evidence for Ca2+ activation and
inactivation sites on the luminal side of the cardiac ryanodine receptor
complex. Circ Res 2000;87:201-206.
142
56. Chu A, Sumbilla C, Inesi G, Jay SD, Campbell KP. Specific association of
calmodulin-dependent protein kinase and related substrates with the junctional
sarcoplasmic reticulum of skeletal muscle. Biochem 1990;29:5899-5905.
57. Clark KA, McElhinny AS, Beckerle MC, Gregorio CC. Striated muscle
cytoarchitecture: An intricate web of form and function. Annu Rev Cell Dev
Biol 2002;18:637-706.
58. Close RI. Dynamic properties of mammalian skeletal muscles. Physiol Rev
1972;52:129-197.
59. Collins TJ, Berridge MJ, Lipp P, Bootman MD. Mitochondria are
morphologically and functionally heterogeneous within cells. EMBO J
2002;21:1616-1627.
66. Csernoch L. Sparks and embers of skeletal muscle: The exciting events of
contractile activation. Pflugers Arch 2007;454:869-878.
70. Davies KJ, Quintanilha AT, Brooks GA, Packer L. Free radicals and tissue
damage produced by exercise. Biochem Biophys Res Comm 1982;107:1198-
1205.
76. Dirksen RT, Avila G. Altered ryanodine receptor function in central core
disease: Leaky or uncoupled Ca2+ release channels? Trends Cardiovasc Med
2002;12:189-197.
77. Dirksen RT, Avila G. Distinct effects on Ca2+ handling caused by malignant
hyperthermia and central core disease mutations in RyR1. Biophys J
2004;87:3193-3204.
78. Du GG, Sandhu B, Khanna VK, Guo XH, MacLennan DH. Topology of the
Ca2+ release channel of skeletal muscle sarcoplasmic reticulum (RyR1). Proc
Natl Acad Sci USA 2002;99:16725-16730.
79. Du GG, Khanna VK, Guo X, MacLennan DH. Central core disease mutations
R4892W, I4897T and G4898E in the ryanodine receptor isoform 1 reduce the
Ca2+ sensitivity and amplitude of Ca2+-dependent Ca2+ release. Biochem J
2004;382:557-564.
80. Dulhunty AF, Laver D, Curtis SM, Pace S, Haarmann C, Gallant EM.
Characteristics of irreversible ATP activation suggest that native skeletal
ryanodine receptors can be phosphorylated via an endogenous camkii.
Biophys J 2001;81:3240-3252.
83. Feng W, Liu G, Allen PD, Pessah IN. Transmembrane redox sensor of
ryanodine receptor complex. J Biol Chem 2000;275:35902-35907.
84. Feng W, Tu J, Yang T, Vernon PS, Allen PD, Worley PF, Pessah IN. Homer
regulates gain of ryanodine receptor type 1 channel complex. J Biol Chem
2002;277:44722-44730.
85. Fessenden JD, Perez CF, Goth S, Pessah IN, Allen PD. Identification of a key
determinant of ryanodine receptor type 1 required for activation by 4-chloro-
m -cresol. J Biol Chem 2003;278:28727-28735.
86. Fill M, Copello JA. Ryanodine receptor calcium release channels. Physiol Rev
2002;82:893-922.
145
93. Fryer MW, Stephenson DG. Total and sarcoplasmic reticulum calcium
contents of skinned fibres from rat skeletal muscle. J Physiol 1996;493:357-
370.
94. Fujii J, Otsu K, Zorzato F, de Leon S, Khanna VK, Weiler JE, O'Brien PJ,
MacLennan DH. Identification of a mutation in porcine ryanodine receptor
associated with malignant hyperthermia. Science 1991;253:448-451.
102. Gyorke I, Hester N, Jones LR, Gyorke S. The role of calsequestrin, triadin,
and junctin in conferring cardiac ryanodine receptor responsiveness to luminal
calcium. Biophys J 2004;86:2121-2128.
104. Hennig R, Lomo T. Firing patterns of motor units in normal rats. Nature
1985;314:164-166.
109. Hidalgo C, Bull R, Marengo JJ, Perez CF, Donoso P. SH oxidation stimulates
calcium release channels (ryanodine receptors) from excitable cells. Biol Res
2000;33:113-124.
111. Hidalgo C, Donoso P, Carrasco MA. The ryanodine receptors Ca2+ release
channels: Cellular redox sensors? IUBMB Life 2005;57:315-322.
112. Hofmann SL, Goldstein JL, Orth K, Moomaw CR, Slaughter CA, Brown MS.
Molecular cloning of a histidine-rich Ca2+-binding protein of sarcoplasmic
reticulum that contains highly conserved repeated elements. J Biol Chem
1989;264:18083-18090.
113. Hollingworth S, Peet J, Chandler WK, Baylor SM. Calcium sparks in intact
skeletal muscle fibers of the frog. J Gen Physiol 2001;118:653-678.
114. Hong S, Kim TW, Choi I, Woo JM, Oh J, Park WJ, Kim dH, Cho C.
Complementary DNA cloning, genomic characterization and expression
analysis of a mammalian gene encoding histidine-rich calcium binding
protein. Biochim Biophys Acta 2005;1727:188-196.
115. Hwang SY, Wei J, Westhoff JH, Duncan RS, Ozawa F, Volpe P, Inokuchi K,
Koulen P. Differential functional interaction of two vesl/homer protein
isoforms with ryanodine receptor type 1: A novel mechanism for control of
intracellular calcium signaling. Cell Calcium 2003;34:177-184.
118. Isaeva EV, Shkryl VM, Shirokova N. Mitochondrial redox state and Ca2+
sparks in permeabilized mammalian skeletal muscle. J Physiol 2005;565:855-
872.
121. Jackson MJ, Edwards RH, Symons MC. Electron spin resonance studies of
intact mammalian skeletal muscle. Biochim Biophys Acta 1985;847:185-190.
124. Jones LR, Zhang L, Sanborn K, Jorgensen AO, Kelley J. Purification, primary
structure, and immunological characterization of the 26-kda calsequestrin
binding protein (junctin) from cardiac junctional sarcoplasmic reticulum. J
Biol Chem 1995;270:30787-30796.
1
25. Jung C, Martins AS, Niggli E, Shirokova N. Dystrophic cardiomyopathy:
Amplification of cellular damage by Ca2+ signalling and reactive oxygen
species-generating pathways. Cardiovasc Res 2008;77:766-773.
128. Kim E, Shin DW, Hong CS, Jeong D, Kim dH, Park WJ. Increased Ca 2+
storage capacity in the sarcoplasmic reticulum by overexpression of hrc
(histidine-rich Ca 2+ binding protein). Biochem Biophys Res Comm
2003;300:192-196.
130. Klein MG, Cheng H, Santana LF, Jiang YH, Lederer WJ, Schneider MF. Two
mechanisms of quantized calcium release in skeletal muscle. Nature
1996;379:455-458.
149
131. Klein MG, Schneider MF. Ca2+ sparks in skeletal muscle. Prog Biophys Mol
Biol 2006;92:308-332.
132. Knudson CM, Stang KK, Jorgensen AO, Campbell KP. Biochemical
characterization of ultrastructural localization of a major junctional
sarcoplasmic reticulum glycoprotein (triadin). J Biol Chem 1993;268:12637-
12645.
134. Kobayashi YM, Alseikhan BA, Jones LR. Localization and characterization of
the calsequestrin-binding domain of triadin 1: Evidence for a charged α-strand
in mediating the protein-protein interaction. J Biol Chem 2000;275:17639-
17646.
136. Krause KH, Milos M, Luan-Rilliet Y, Lew DP, Cox JA. Thermodynamics of
cation binding to rabbit skeletal muscle calsequestrin. Evidence for distinct
Ca2+- and Mg2+-binding sites. J Biol Chem 1991;266:9453-9459.
139. Lamb GD, Cellini MA, Stephenson DG. Different Ca2+ releasing action of
caffeine and depolarisation in skeletal muscle fibres of the rat. J Physiol
2001;531:715-728.
141. Lannergren J, Bruton JD. Mitochondrial Ca2+ in mouse soleus single muscle
fibres in response to repeated tetanic contractions. Adv Exp Med Biol
2003;538:557-562.
143. Laver DR, Baynes TM, Dulhunty AF. Magnesium inhibition of ryanodine-
receptor calcium channels: Evidence for two independent mechanisms. J
Membr Biol1997;156:213-229.
144. Laver DR, O'Neill ER, Lamb GD. Luminal Ca2+-regulated Mg2+ inhibition of
skeletal RyRs reconstituted as isolated channels or coupled clusters. J Gen
Physiol 2004;124:741-758.
145. Leberer E, Charuk JH, Green NM, MacLennan DH. Molecular cloning and
expression of cdna encoding a lumenal calcium binding glycoprotein from
sarcoplasmic reticulum. Proc Natl Acad Sci USA 1989;86:6047-6051.
146. Leberer E, Timms BG, Campbell KP, MacLennan DH. Purification, calcium
binding properties, and ultrastructural localization of the 53,000- and 160,000
(sarcalumenin)-dalton glycoproteins of the sarcoplasmic reticulum. J Biol
Chem 1990;265:10118-10124.
147. Lee HG, Kang H, Kim DH, Park WJ. Interaction of HRC (histidine-rich Ca2+-
binding protein) and triadin in the lumen of sarcoplasmic reticulum. J Biol
Chem 2001;276:39533-39538.
148. Lee JM, Rho SH, Shin DW, Cho C, Park WJ, Eom SH, Ma J, Kim dH.
Negatively charged amino acids within the intraluminal loop of ryanodine
receptor are involved in the interaction with triadin. J Biol Chem
2004;279:6994-7000.
150. Lindsay AR, Manning SD, Williams AJ. Monovalent cation conductance in
the ryanodine receptor-channel of sheep cardiac muscle sarcoplasmic
reticulum. J Physiol 1991;439:463-480.
151. Ludolph DC, Konieczny SF. Transcription factor families: Muscling in on the
myogenic program. FASEB J 1995;9:1595-1604.
151
152. Lueck JD, Mankodi A, Swanson MS, Thornton CA, Dirksen RT. Muscle
chloride channel dysfunction in two mouse models of myotonic dystrophy. J
Gen Physiol 2007;129:79-94.
154. Lynch PJ, Tong J, Lehane M, Mallet A, Giblin L, Heffron JJ, Vaughan P,
Zafra G, MacLennan DH, McCarthy TV. A mutation in the
transmembrane/luminal domain of the ryanodine receptor is associated with
abnormal Ca2+ release channel function and severe central core disease. Proc
Natl Acad Sci USA 1999;96:4164-4169.
155. Lytton J, MacLennan DH. Molecular cloning of cdnas from human kidney
coding for two alternatively spliced products of the cardiac Ca2+-ATPase
gene. J Biol Chem 1988;263:15024-15031.
159. MacLennan DH, Yip WH, Iles GH, Seeman P. Isolation of sarcoplasmic
reticulum proteins. Cold Spring Harbor Symp Quant 1972:37469-37478.
160. MacLennan DH, Brandl CJ, Korczak B, Green NM. Amino-acid sequence of
a Ca2+ + Mg2+-dependent ATPase from rabbit muscle sarcoplasmic reticulum,
deduced from its complementary DNA sequence. Nature 1985;316:696-700.
163. MacLennan DH. Ca2+ signalling and muscle disease. Eur J Biochem
2000;267:5291-5297.
152
164. MacLennan DH, Loke J, Karpati G. Diseases associated with ion channel and
ion transporter defects: Brody disease and brody syndrome. Molecular and
structural basis of skeletal muscle diseases. Basel; ISN Neuropath Press 2002.
p 103-105.
165. Marengo JJ, Hidalgo C, Bull R. Sulfhydryl oxidation modifies the calcium
dependence of ryanodine-sensitive calcium channels of excitable cells.
Biophys J 1998;74:1263-1277.
166. Martins AS, Shkryl VM, Nowycky MC, Shirokova N. Reactive oxygen
species contribute to Ca2+ signals produced by osmotic stress in mouse
skeletal muscle fibres. J Physiol 2008;586:197-210.
171. McArdle A, van der Meulen J, Close GL, Pattwell D, Van Remmen H, Huang
TT, Richardson AG, Epstein CJ, Faulkner JA, Jackson MJ. Role of
mitochondrial superoxide dismutase in contraction-induced generation of
reactive oxygen species in skeletal muscle extracellular space. Am J Physiol
Cell Physiol 2004;286:C1152-C1158.
173. McCarthy TV, Quane KA, Lynch PJ. Ryanodine receptor mutations in
malignant hyperthermia and central core disease. Hum Mut 2000;15:410-417.
153
174. McCormack JG, Halestrap AP, Denton RM. Role of calcium ions in
regulation of mammalian intramitochondrial metabolism. Physiol Rev
1990;70:391-425.
175. Meier PJ, Spycher MA, Meyer UA. Isolation and characterization of rough
endoplasmic reticulum associated with mitochondria from normal rat liver.
Biochim Biophys Acta 1981;646:283-297.
180. Neher E. Vesicle pools and Ca2+ microdomains: New tools for understanding
their roles in neurotransmitter release. Neuron 1998;20:389-399.
181. Nelson MT, Cheng H, Rubart M, Santana LF, Bonev AD, Knot HJ, Lederer
WJ. Relaxation of arterial smooth muscle by calcium sparks. Science
1995;270:633-637.
184. Odermatt A, Taschner PE, Khanna VK, Busch HF, Karpati G, Jablecki CK,
Breuning MH, MacLennan DH. Mutations in the gene-encoding SERCA1, the
fast-twitch skeletal muscle sarcoplasmic reticulum Ca2+ ATPase, are
associated with Brody disease. Nat Gen 1996;14:191-194.
188. Pacher P, Thomas AP, Hajnoczky G. Ca2+ marks: Miniature calcium signals
in single mitochondria driven by ryanodine receptors. Proc Natl Acad Sci
USA 2002;99:2380-2385.
189. Pan Z, Yang D, Nagaraj RY, Nosek TA, Nishi M, Takeshima H, Cheng H,
Ma J. Dysfunction of store-operated calcium channel in muscle cells lacking
MG29. NatCell Biol 2002;4:379-383.
190. Park KW, Goo JH, Chung HS, Kim H, Kim DH, Park WJ. Cloning of the
genes encoding mouse cardiac and skeletal calsequestrins: Expression pattern
during embryogenesis. Gene 1998;217:25-30.
191. Posterino GS, Cellini MA, Lamb GD. Effects of oxidation and cytosolic redox
conditions on excitation-contraction coupling in rat skeletal muscle. J Physiol
2003;547:807-823.
192. Posterino GS, Lamb GD. Effect of sarcoplasmic reticulum Ca2+ content on
action potential-induced Ca2+ release in rat skeletal muscle fibres. J Physiol
2003;551:219-237.
193. Postma AV, Denjoy I, Hoorntje TM, Lupoglazoff JM, Da Costa A, Sebillon P,
Mannens MM, Wilde AA, Guicheney P. Absence of calsequestrin 2 causes
severe forms of catecholaminergic polymorphic ventricular tachycardia. Circ
Res 2002;91:e21-e26.
155
196. Proske U, Allen TJ. Damage to skeletal muscle from eccentric exercise. Exerc
Sport Sci Rev 2005;33:98-104.
197. Protasi F. Structural interaction between ryrs and dhprs in calcium release
units of cardiac and skeletal muscle cells. Front Biosci 2002;7:d650-d658.
208. Sencer S, Papineni RV, Halling DB, Pate P, Krol J, Zhang JZ, Hamilton SL.
Coupling of RyR1 and l-type calcium channels via calmodulin binding
domains. J Biol Chem 2001;276:38237-38241.
210. Shin DW, Ma J, Kim DH. The asp -rich region at the carboxyl-terminus of
calsequestrin binds to Ca2+ and interacts with triadin. FEBS Lett
2000;486:178-182.
212. Shkryl VM, Shirokova N. Transfer and tunneling of Ca2+ from sarcoplasmic
reticulum to mitochondria in skeletal muscle. J Biol Chem 2006;281:1547-
1554.
213. Shuaib A, Paasuke RT, Brownell KW. Central core disease. Clinical features
in 13 patients. Medicine 1987;66:389-396.
220. Stiber J, Hawkins A, Zhang ZS, Wang S, Burch J, Graham V, Ward CC, Seth
M, Finch E, Malouf N, Williams RS, Eu JP, Rosenberg P. Stim1 signalling
controls store-operated calcium entry required for development and contractile
function in skeletal muscle. NatCell Biol 2008;10:688-697.
221. Stromer MH, Goll DE, Young RB, Robson RM, Parrish FC, Jr. Ultrastructural
features of skeletal muscle differentiation and development. J Anim Sci
1974;38:1111-1141.
222. Suk JY, Kim YS, Park WJ. Hrc (histidine-rich Ca2+ binding protein) resides in
the lumen of sarcoplasmic reticulum as a multimer. Biochem Biophys Res
Comm 1999;263:667-671.
223. Sun J, Xu L, Eu JP, Stamler JS, Meissner G. Classes of thiols that influence
the activity of the skeletal muscle calcium release channel. J Biol Chem
2001;276:15625-15630.
224. Sweeney HL. The importance of the creatine kinase reaction: The concept of
metabolic capacitance. Med Sci Sports Exerc 1994;26:30-36.
228. Tang W, Ingalls CP, Durham WJ, Snider J, Reid MB, Wu G, Matzuk MM,
Hamilton SL. Altered excitation-contraction coupling with skeletal muscle
specific fkbp12 deficiency. FASEB J 2004;18:1597-1599.
229. Territo PR, Mootha VK, French SA, Balaban RS. Ca(2+) activation of heart
mitochondrial oxidative phosphorylation: Role of the F0F1-ATPase. Am J
Physiol Cell Physiol 2000;278:C423-C435.
231. Tinker A, Lindsay AR, Williams AJ. A model for ionic conduction in the
ryanodine receptor channel of sheep cardiac muscle sarcoplasmic reticulum. J
Gen Physiol 1992;100:495-517.
241. Trollinger DR, Cascio WE, Lemasters JJ. Selective loading of rhod 2 into
mitochondria shows mitochondrial Ca2+ transients during the contractile cycle
in adult rabbit cardiac myocytes. Biochem Biophys Res Comm 1997;236:738-
742.
247. Wang J, Best PM. Inactivation of the sarcoplasmic reticulum calcium channel
by protein kinase. Nature 1992;359:739-741.
160
248. Wang S, Trumble WR, Liao H, Wesson CR, Dunker AK, Kang CH. Crystal
structure of calsequestrin from rabbit skeletal muscle sarcoplasmic reticulum.
Nat Struct Biol 1998;5:476-483.
251. Ward CW, Feng W, Tu J, Pessah IN, Worley PK, Schneider MF. Homer
protein increases activation of Ca2+ sparks in permeabilized skeletal muscle. J
Biol Chem 2004;279:5781-5787.
252. Warren GL, Ingalls CP, Lowe DA, Armstrong RB. Excitation-contraction
uncoupling: Major role in contraction-induced muscle injury. Exerc Sport Sci
Rev 2001;29:82-87.
255. Weiss RG, O'Connell KM, Flucher BE, Allen PD, Grabner M, Dirksen RT.
Functional analysis of the r1086h malignant hyperthermia mutation in the
DHPR reveals an unexpected influence of the iii-iv loop on skeletal muscle
EC coupling. Am J Physiol Cell Physiol 2004;287:C1094-C1102.
256. Westhoff JH, Hwang SY, Scott DR, Ozawa F, Volpe P, Inokuchi K, Koulen
P. Vesl/homer proteins regulate ryanodine receptor type 2 function and
intracellular calcium signaling. Cell Calcium 2003;34:261-269.
260. Xiao B, Tu JC, Petralia RS, Yuan JP, Doan A, Breder CD, Ruggiero A,
Lanahan AA, Wenthold RJ, Worley PF. Homer regulates the association of
group 1 metabotropic glutamate receptors with multivalent complexes of
homer-related, synaptic proteins. Neuron 1998;21:707-716.
264. Zable AC, Favero TG, Abramson JJ. Glutathione modulates ryanodine
receptor from skeletal muscle sarcoplasmic reticulum. Evidence for redox
regulation of the Ca2+ release mechanism. J Biol Chem 1997;272:7069-7077.
267. Zhang Y, Fujii J, Phillips MS, Chen HS, Karpati G, Yee WC, Schrank B,
Cornblath DR, Boylan KB, MacLennan DH. Characterization of cDNA and
genomic DNA encoding SERCA1, the Ca2+-ATPase of human fast-twitch
skeletal muscle sarcoplasmic reticulum, and its elimination as a candidate
gene for Brody disease. Genomics 1995;30:415-424.
269. Zimanyi I, Buck E, Abramson JJ, Mack MM, Pessah IN. Ryanodine induces
persistent inactivation of the Ca2+ release channel from skeletal muscle
sarcoplasmic reticulum. Mol Pharmacol 1992;42:1049-1057.