-
Notifications
You must be signed in to change notification settings - Fork 0
/
Copy pathCalcium_cardiac_cell.tex
3418 lines (2882 loc) · 157 KB
/
Calcium_cardiac_cell.tex
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
93
94
95
96
97
98
99
100
101
102
103
104
105
106
107
108
109
110
111
112
113
114
115
116
117
118
119
120
121
122
123
124
125
126
127
128
129
130
131
132
133
134
135
136
137
138
139
140
141
142
143
144
145
146
147
148
149
150
151
152
153
154
155
156
157
158
159
160
161
162
163
164
165
166
167
168
169
170
171
172
173
174
175
176
177
178
179
180
181
182
183
184
185
186
187
188
189
190
191
192
193
194
195
196
197
198
199
200
201
202
203
204
205
206
207
208
209
210
211
212
213
214
215
216
217
218
219
220
221
222
223
224
225
226
227
228
229
230
231
232
233
234
235
236
237
238
239
240
241
242
243
244
245
246
247
248
249
250
251
252
253
254
255
256
257
258
259
260
261
262
263
264
265
266
267
268
269
270
271
272
273
274
275
276
277
278
279
280
281
282
283
284
285
286
287
288
289
290
291
292
293
294
295
296
297
298
299
300
301
302
303
304
305
306
307
308
309
310
311
312
313
314
315
316
317
318
319
320
321
322
323
324
325
326
327
328
329
330
331
332
333
334
335
336
337
338
339
340
341
342
343
344
345
346
347
348
349
350
351
352
353
354
355
356
357
358
359
360
361
362
363
364
365
366
367
368
369
370
371
372
373
374
375
376
377
378
379
380
381
382
383
384
385
386
387
388
389
390
391
392
393
394
395
396
397
398
399
400
401
402
403
404
405
406
407
408
409
410
411
412
413
414
415
416
417
418
419
420
421
422
423
424
425
426
427
428
429
430
431
432
433
434
435
436
437
438
439
440
441
442
443
444
445
446
447
448
449
450
451
452
453
454
455
456
457
458
459
460
461
462
463
464
465
466
467
468
469
470
471
472
473
474
475
476
477
478
479
480
481
482
483
484
485
486
487
488
489
490
491
492
493
494
495
496
497
498
499
500
501
502
503
504
505
506
507
508
509
510
511
512
513
514
515
516
517
518
519
520
521
522
523
524
525
526
527
528
529
530
531
532
533
534
535
536
537
538
539
540
541
542
543
544
545
546
547
548
549
550
551
552
553
554
555
556
557
558
559
560
561
562
563
564
565
566
567
568
569
570
571
572
573
574
575
576
577
578
579
580
581
582
583
584
585
586
587
588
589
590
591
592
593
594
595
596
597
598
599
600
601
602
603
604
605
606
607
608
609
610
611
612
613
614
615
616
617
618
619
620
621
622
623
624
625
626
627
628
629
630
631
632
633
634
635
636
637
638
639
640
641
642
643
644
645
646
647
648
649
650
651
652
653
654
655
656
657
658
659
660
661
662
663
664
665
666
667
668
669
670
671
672
673
674
675
676
677
678
679
680
681
682
683
684
685
686
687
688
689
690
691
692
693
694
695
696
697
698
699
700
701
702
703
704
705
706
707
708
709
710
711
712
713
714
715
716
717
718
719
720
721
722
723
724
725
726
727
728
729
730
731
732
733
734
735
736
737
738
739
740
741
742
743
744
745
746
747
748
749
750
751
752
753
754
755
756
757
758
759
760
761
762
763
764
765
766
767
768
769
770
771
772
773
774
775
776
777
778
779
780
781
782
783
784
785
786
787
788
789
790
791
792
793
794
795
796
797
798
799
800
801
802
803
804
805
806
807
808
809
810
811
812
813
814
815
816
817
818
819
820
821
822
823
824
825
826
827
828
829
830
831
832
833
834
835
836
837
838
839
840
841
842
843
844
845
846
847
848
849
850
851
852
853
854
855
856
857
858
859
860
861
862
863
864
865
866
867
868
869
870
871
872
873
874
875
876
877
878
879
880
881
882
883
884
885
886
887
888
889
890
891
892
893
894
895
896
897
898
899
900
901
902
903
904
905
906
907
908
909
910
911
912
913
914
915
916
917
918
919
920
921
922
923
924
925
926
927
928
929
930
931
932
933
934
935
936
937
938
939
940
941
942
943
944
945
946
947
948
949
950
951
952
953
954
955
956
957
958
959
960
961
962
963
964
965
966
967
968
969
970
971
972
973
974
975
976
977
978
979
980
981
982
983
984
985
986
987
988
989
990
991
992
993
994
995
996
997
998
999
1000
\chapter{Cardiac cycle: calcium handling in cardiac cells}
\label{chap:calc-handl-card}
The purpose of modeling cardiac cells is to reach an adequate quantitative
understanding of the relationship between molecular function and the integrated
behavior of the cell at the tissue and whole-heart level in healthy and diseased
conditions. Typically, the proper regulations of ion concentrations in the
cells play important roles to this normal behavior. Ions can travel passively across the
membrane through ionic channels. The ionic gradient are tightly regulated by
active exchanges/pumps, and ion homeostasis is maintained via pump-leak model.
We have been studied these aspects in the previous chapters. From this chapter,
we will study how these components are assembled into a complex, yet
well-organized dynamical system; that is a cardiac cell.
Compared to nerve cells, modeling for the cardiac cells is far from complete,
especially at the whole-heart level. At the single cellular level, the modeling
with cardiac cells is very challenging as there is a significant difference in the
shape of action potential (AP), not only between that of cardiac cell and nerve
cell, but also between cardiac cells at different parts of the heart (nodal
cells, Purkinje fibre cells and myocardial ventricular cells).
\begin{framed}
An AP is the result of the global elevation of intracellular $\Ca$. The gain
of global elevation of $\Ca$ is $\sim$10x, i.e. from 0.1$\mu$M at
resting level to $1\mu$M. To model the action potential in
myocytes, in this chapter, we will study models built specifically
for cardiac cells at which $\Ca$ currents is of specifically
important. % Nevertheless, the early models
% still took HH model as the basis.
\end{framed}
Among different ions, in cardiac cell, calcium ions play a critical role to the
normal function of the heart, the contraction-excitation coupling. Calcium
is the major second mesenger in cardiac cells. The elevation of calcium
concentration is tightly coupled to membrane potential depolarization and the
force generation for contraction. Chap.
\ref{chap:models-ap-purkinje} to \ref{chap:ap_ventricular_myocyte_localcontrol}
will focus on the dynamics of calcium signaling at the cellular level, in which
\begin{itemize}
\item Chapter \ref{chap:models-ap-purkinje} will discuss AP in Purkinje
fibre cells. As we have seen in section \ref{sec:noble-model}, in Purkinjie
fiber, the shape of a cardiac AP is different from nerve AP. In
addition, the shape of the AP is different between pacemaker cells
and cardiac non-pacemaker cells, as shown in
Fig.~\ref{fig:nerve-cardiac-AP}.
\item Chapter \ref{chap:models-ap-nodal} will discuss AP in nodal cells
(SA and VA node). AP in AV nodal cells vary substantially.
\item Chapter \ref{chap:ap_ventricular_myocyte} will discuss AP in
myocardial cells. Even here, epicardial, midmyocardial and
endocardial cells have noticeable differences in action potential (AP) duration.
\item Chapter \ref{chap:sparkology-study-ca} discuss the modeling of $\Ca$
sparks, the elementary events in calcium dynamics. Then,
chapter~\ref{chap:ap_ventricular_myocyte_localcontrol} discuss new approaches
of whole-cell models that take into accounts the large number of CRUs in the
cell.
\item Chap. \ref{chap:force_calcium} focuses on force generation. We'll
discuss models at the whole-heart or tissue level in another part of the book (Part
\ref{part:whole-heart}).
\end{itemize}
\begin{framed}
There are two forces of modeling systems: (1) one wants all channels and effects
to be specified with a great detail, (2) one wants to neglect some that are
inconsequential to the final behavior of the model.
\end{framed}
In the previous chapters, we have learned models for different proteins
involving in regulating ion concentrations in the cell. These include ionic
channels, pumps/exchangers on the SL and SR membrane. However, this is not all.
The next step is to combine the models for the different subcellular componenets
into a working whole-cell and tissue level. This can provide an abundant of
information of understanding the biophysical basis of action potential (AP), the
changes in gene/protein expression level (represented by the change in the
maximum conductances or maximum transfer speed), or effects of gene mutations to
alterations of AP and calcium transient morphology, which can provide a
quantatively insights into the mechanism of pathological conditions, e.g.
arrhythmias.
Before going into whole-cell models, it's important to know that the cell is not
a homogeneous medium, but it has different cellular components, each one has
its own environment (e.g. ion concentrations -
Section.\ref{sec:ion-concentrations}), and they all involve, some how, in the
pathways of calcium signalling or excitation-contraction coupling (ECC) process.
By understanding the role of these components, we can have a good idea how
whole-cell models are created. This is the topic of this chapter.
\section{Calcium homeostasis}
In the cytoplasm, $\Ca$ extrusion from the cell is mediated by
\begin{enumerate}
\item PMCA pump:
\item NCX: while most of the residues is extruded out of the cell by
Na/Ca exchanger (NCX) in the submembrane space
\item SERCA pump: reuptake a large portion of $\Ca$
back into the NSR
\item uniporter: in the mitochondria by the electrophoretic uniporter -
Sect.\ref{sec:MCU}
\end{enumerate}
\section{Cardiac cycle: action potential in cardiac cells}
\label{sec:AP-cardiac}
\label{sec:cardiac-cycle}
\label{sec:ECC}
The cardiac cycle refers to the periodic propagation of electrical signal all
over the heart, which goes through two stages: {\bf diastole} and {\bf systole}.
During diastole, all four chambers of the heart are relaxed. Systole is
initiated when the electrical signal from the brain to the SA node
(Sect.\ref{sec:slow-response-ap}). The signal then propagated throughout the
atrial chambers, reaching the AV node; where the depolarization pauses for a
short time period (0.1 s in humans) to ensure completion of atrial systole.
After the delay in AV node, the signal propagate very fast through the His
bundle and then the bundle branches.
These branches give rise to thin filaments called Purkinje fibers, composed of
noncontractile cells that distribute the action potential to ventricular
myocytes and enable a uniformly contraction in the ventricle.
The concept of AP is given in Sect.\ref{sec:action-potential}.
In section \ref{sec:AP-neuron}, we have studied the neuronal
action potential (AP). In this section, we will go into details the AP
of cardiac cells.
\begin{figure}[hbt]
\centerline{\includegraphics[height=5cm,
angle=0]{./images/nodal-AP.eps}
,\includegraphics[height=5cm,
angle=0]{./images/cardiac-AP.eps} }
\caption{(A) Neuron AP; (B) Cardiac Purkinje AP}
\label{fig:nerve-cardiac-AP}
\end{figure}
Depending on types of ionic channels, and their densities, APs are different in
shapes between species (Sect.\ref{sec:detail-ap-different}) and between
different cell types in a single species, Fig.~\ref{fig:nerve-cardiac-AP}.
In nerve cells, the duration of AP is about 1ms. In skeletal muscle cells, the
duration of AP is about 2-5ms. In cardiac myocytes, the duration of AP is in the
range 200-400 ms
\footnote{\url{http://www.cvphysiology.com/Arrhythmias/A010.htm}}.
The long AP in cardiac myocytes is characterized by a longer plateau, as shown
in Fig. \ref{fig:cardiac_AP} that is maintained mostly by $\Ca$ ionic currents.
Quantities related to AP include transient rate, the duration of the AP, the
plateau range, the replorarizing duration ... Typically, a computational model
of cardiac cells need to replicate the AP of a specific species, with all of
such properties. Pathological conditions related to AP include the early after
depolarization (EAD) - Sect.\ref{sec:early-afterd-ead}, delayed after
depolarization (DAD) - Sect.\ref{sec:delay-afterd-dad}. We will learn the
quantitatively difference between AP in different cardiac cell types in
Part.\ref{part:compartmental_model}.
\begin{framed}
In cardiac cells, T-type $\Ca$ channels is important in initiating AP (T =
transient), while L-type $\Ca$ channels is important in sustaining an AP
(L=long). Because of their rapid kinetics, T-type is often found in
cells undergoing rhythmic electrical behavior (e.g. neuron cell bodies
involved in walking/breathing, pacemaker cells (SA node and AV node)
of the heart; while L-type channels are found in cardiac cells
(i.e. ventricular myocytes).
\end{framed}
% in which L-type $\Ca$ channels is the major canals.
% , there are two general types of cardiac
% AP: pacemaker AP (or ``slow response'' AP) and non-pacemaker AP (or ``fast
% response'' AP), as shown in Fig.~\ref{fig:cardiac_AP}(B)(C). Non-pacemaker APs
% are found through the heart, except for pacemaker cells. Here, we mainly
% concentrate on non-pacemaker APs
With different types of cardiac cells in the heart, there are two general types
of cardiac AP: non-pacemaker AP (or ``fast response'' AP) and pacemaker AP (or
``slow response'' AP), as shown in Fig.~\ref{fig:cardiac_AP}(B)(C).
\begin{enumerate}
\item Non-pacemaker APs (Sect.\ref{sec:fast-response-ap} are found through the
heart, except for nodal cells, with \ce{Ca^2+} influx prolonging the duration of AP to
produce a characteristic plateau phase.
Examples of non-pacemaker cells are {\it myocardial cells} (atria and
ventricles), and {\it Purkinje fibre cells}.
\item Pacemaker APs (Sect.\ref{sec:slow-response-ap}) are found in nodal tissues
within the heart, and \ce{Ca^2+} influx involves in the initial depolarization
phase of AP.
\end{enumerate}
\textcolor{red}{Here, we mainly concentrate on non-pacemaker APs}; though models
of pace-makers (e.d. nodal cells) are also described
(Chap.~\ref{chap:models-ap-nodal}). Examples of non-pacemaker cells are {\it
Atrial myocytes}, {\it ventricular myocytes}
(Chap.~\ref{chap:ap_ventricular_myocyte} and
Chap.~\ref{chap:ap_ventricular_myocyte_localcontrol}) and {\it Purkinje cells}
(Chap.~\ref{chap:models-ap-purkinje}). Fig.\ref{fig:ion_channel_roles_AP} shows
the roles of all types of ion channels at different phases of AP.
In muscle fibres, when the membrane is suddenly depolarized, the
potassium permeability of the membrane is increased; a phenomenon
known as {\bf K-activation} \citep{grundfest1961ime} or
{\bf delayed rectification}~\citep{hodgkin1949icu}. After the AP is
maintained for a few seconds, the increase in K permeability
diminished and disappears completely, i.e. the stage of K-inactivation
occur. To hold the potential not to fall down so fast, it is the
influx of \ce{Ca^2+} that leads to the CICR process. This keep the
membrane potential for a long enough time, and cause a plateau shape
in the AP.
% in which L-type \ce{Ca^2+} channels is the major canals.
\begin{figure}[htb]
\centerline{\includegraphics[height=4cm]{./images/action_potentials_compare.eps}}
\centerline{\includegraphics[height=5cm]{./images/pacemaker_AP.eps}, \includegraphics[height=5cm]{./images/ventricular_action_potential.eps}}
\caption{(A) compare between nerve cell AP vs. cardiac
non-pacemaker AP; Phases of a cardiac AP in (B) pacemaker cells,
(C) non-pacemaker cells}\label{fig:cardiac_AP}
\end{figure}
\subsection{``fast response'' AP (non-pacemaker AP)}
\label{sec:fast-response-ap}
The ``fast response'' AP is caused by the relatively changes in fast
\ce{Na+}, slow $\Ca$ and \ce{K+} conductances and currents. Once
the AP is initialized, the cell become insensitive to stimulus. In
other words, the duration of time comprising phase 0, 1, 2, and part
of phase 3 is the time within that a new AP cannot be initiated. This
period of time is termed {\bf effective refractory period} (ERP) or
{\bf absolute refractory period} (ARP) of the cardiac cell. In other
words, during this time, the stimulation by an adjacent cell to the
examining cell does not produce new, propagated AP. Near the end of
the activation impulse, the cell may be re-activated, but with a
stimulus stronger than normal. This period is called
{\bf relative refractory period}. In medicine, to prevent
{\it tachyarrythmias}, drugs are used to increase ERP to abolish
reentry currents.
\begin{figure}[hbt]
\centerline{\includegraphics[height=5cm]{./images/channel_activation_AP.eps},
\includegraphics[height=5cm]{./images/cardiac-current.eps}}
\caption{Activities of different types of channels during resting
vs. action potential in non-pacemaker cells}
\label{fig:channel-activity}
\end{figure}
Phases in a ``fast response'' AP, Fig.~\ref{fig:cardiac_AP}(C) or
Fig.~\ref{fig:channel-activity}:
\begin{itemize}
\item phase 4: the resting phase
At resting phase, some \ce{K+}-channels are still active, especially
I$_{\k1}$. In non-pacemaker cells, it has true resting potential (phase 4)
as it remains near the equilibrium potential for \ce{K+} ($E_{\ce{K+}} \approx
-90$mV).
\item phase 0: the sharp rise - {\bf AP upstroke} - corresponds to the
fast-inward of \ce{Na+} current $I_\na$, and to a lesser extent of
L-type $\Ca$ channels $I_\CaL$.
When the membrane depolarize to a threshold voltage, of about -70mV to -60mV,
there is a rapid depolarization that is caused by a transient
increase in intracellular [\ce{Na+}] through inward fast
\ce{Na+}-channels, as shown in Fig.~\ref{fig:channel-activity}.
This process changes the membrane potential away from $E_{\ce{K+}}$
(negative) and move to $E_{\ce{Na+}}$ (positive). Na/Ca exchanger is suspected
to play a role in abnormal trigger of AP in failing heart.
\item phase 1: first decay corresponds to the $\Na$-channel
inactivation and the activation of the $V_m$-dependent transient
outward $\K$ current $I_{to,1}$~\citep{sah2003}. A $V_m$-independent
$\Ca$-dependent chloride current $I_{to,2}$ may also contribute, yet
whether it's contributing to shape human AP is still
controversy~\citep{ORourke1996}. $I_{to}$ is found in most species (canine,
rabbit, human), but not in certain species (rat).
This {\bf notch} is apparent in ventricular myocytes (isolated from epi- and
mid- myocardial regions, but not from endocardium) of some mammalian species
(canine). During this time, the activation of L-type $\Ca$ channels also bring
a small amount of $\Ca$ which brings $V_m$ to a more depolarized state. This
is reflected by the shape of the AP dome. This amount of calcium is not enough
to initiate actin-myosin cross-bridge; yet it can trigger a larger amount of
$\Ca$ to be released from internal storage sarcoplasmic reticulum (SR/ER) via
RyR2 gating channels.
\item phase 2: a plateau sustained by a balance between $I_\CaL$ and
outward \ce{K+} channels (the rapid $I_\Kr$ and slow-activating
delayed rectifier $I_\Ks$
currents~\citep{horie1990,sanguinetti1990tcc} and the plateau $\K$
current $I_\Kp$~\citep{yue1988ncp}), \ce{Na+}/$\Ca$-exchanger
and \ce{Na+}/\ce{K+}-exchanger also play minor roles in this phase.
During this phase, the membrane potential is held as a high voltage
for a few hundreds of milliseconds, with low membrane conductance.
\begin{framed}
Phase 2 is very important. Thanks to this voltage hold, the cardiac
AP plays an important role in coordinating the contractility of the
heart~\citep{kleber2004bmc}.
\end{framed}
\item phase 3: second decay (rapid repolarization) corresponds to the
continue activating of $I_\Kr, I_\Ks$, and the inward rectifier $\K$
current ($V_m$-repolarization-activated) $I_{K1}$.
During this time, L-type channels are closed while \ce{K+} channels are still
open. Further, the repolarizing induces more \ce{K+} channels to open, pulling
the voltage back to the repolarized state, and even lower than the resting
potential. Thus, to recover back to resting potential, the cell need to
decrease the concentrations of $\Ca$ and $\Na$ using Na/K-ATPase pump,
Na/Ca-exchanger, and SERCA-ATPase pump.
\end{itemize}
\begin{framed}
The major difference between skeletal cells and cardiac cells is that:
influx of \ce{Na+} quickly depolarize the skeletal myocyte and trigger
calcium release from the SR. In cardiac myocytes, the release of
$\Ca$ from the SR, however, is induced by $\Ca$ influx into
the cell through a process called calcium-induced calcium-released
(CICR) which is studied in the next chapter. So, the influx of $\Ca$
is important in cardiac cells, but not in skeletal cells for an AP to
initiate.
\end{framed}
At resting phase, the resting potential is maintained mainly by
the activity of some \ce{K+}-channels. In non-pacemaker cells, it
has true resting potential (phase 4) as it remains near the
equilibrium potential for \ce{K+} (about -90mV).
When there is an electrical impulse, mainly due to the influx of
\ce{Ca^2+}, the membrane voltage is depolarized. This trigger the
opening of more low-voltage activated \ce{K+} channels. At the
threshold voltage, of about -70mV, there is a rapid depolarization
that is caused by a transient increase in intracellular [\ce{Na+}]
through fast \ce{Na+}-channels, as shown in
Fig.~\ref{fig:channel-activity}. This process change the membrane
potential away from $E_{\ce{K+}}$ (negative) and move to $E_{\ce{Na+}}$
(positive).
As shown in Fig.~\ref{fig:cardiac_AP}(C), the sharp rise (segment 0)
corresponds to the influx of \ce{Na+}. As the activity of \ce{K+}
channels is bi-stage. In other words, the high membrane voltage
deactivate the \ce{K+} channels. This causes a decay in membrane
voltage (segment 1). However, thanks to the release of \ce{Ca^2+} from
intracellular storage, via the CICR process, it held the membrane
potential as a high voltage for a few hundreds of milliseconds
(segment 2). With a repolarizing efflux of \ce{K+} ions, the membrane
voltage then decreases rapidly (segment 3). The repolarizing doesn't
stop at the resting potential but go to a more negative value, then
slowly increasing to the resting potential before a next AP can
occur. This last stage is known as the {\bf afterhyperpolarization}
(AHP) stage\footnote{more information is covered in Appendix A of
Computational Biology book}.
% whereas the two decays
% (segments 1 and 3) corresponds to the sodium-channel inactivation and
% repolarizing efflux of potassium ions, respectively. There is also an
% extended plateau (2) in which the membrane potential is held as a high
% voltage for a few hundreds of milliseconds. This is the result from
% the activation of voltage-gated \ce{Ca^2+} channels.
Thanks to this voltage hold, the cardiac AP plays an important role in
coordinating the contraction of the heart~\cite{kleber2004bmc}.
\begin{figure}[hbt]
\centerline{\includegraphics[height=6cm]{./images/channel_activation_AP.eps}}
\caption{Activities of different types of channels during resting
vs. action potential in non-pacemaker cells}
\label{fig:channel-activity}
\end{figure}
\begin{figure}[hbt]
\centerline{\includegraphics[height=5cm,
angle=0]{./images/time_fast_AP.eps}}
\caption{Time in fast AP: (1) depolarization, (2) absolute
refractory, (3) total refractory, (4) repolarization, (5)
automaticity cycle length (automaticity = the ability to initiate impulses)}
\label{fig:fast_AP}
\end{figure}
The ``fast response'' AP is caused by the relatively changes in fast
\ce{Na+}, slow \ce{Ca^2+} and \ce{K+} conductances and currents. Once
the AP is initialized, the duration of time comprising phase 0, 1, 2,
and part of phase 3 is the time within that a new AP cannot be
initiated, as shown in Fig.~\ref{fig:fast_AP}. This period of time is
termed {\bf effective refractory period} (ERP) or
{\bf absolute refractory period} (ARP) of the cardiac cell. In other
words, during this time, the stimulation by an adjacent cell to the
examining cell does not produce new, propagated AP. In medicine, to
prevent tachyarrythmias, drugs are used to increase ERP to abolish
reentry currents.
In cardiac cells, T-type \ce{Ca^2+} channels is important in
initiating AP, while L-type is important in sustaining an AP. Because
of their rapid kinetics, T-type is often found in cells undergoing
rhythmic electrical behavior (e.g. neuron cell bodies involved in
walking/breathing, pacemaker cells (at SA node and AV node) of the
heart; while L-type channels are found in myocardial cells.
The major difference between skeletal cells and cardiac cells is that:
influx of \ce{Na+} quickly depolarize the skeletal myocyte and trigger
calcium release from the SR. In cardiac myocytes, the release of
\ce{Ca^2+} from the SR is induced by \ce{Ca^2+} influx into the cell
through a process called calcium-induced calcium-released (CICR) which
is studied in the next chapter.
\subsection{``slow response'' AP (AV node or SA node)}
\label{sec:slow-response-ap}
The ``slow'' response AP occurs in pacemaker cells found in the
sino-atrial (SA) node. We don't observe a transient in crease in phase
0 of the AP at the AV node as it is not dependent on fast \ce{Na+}
channels as in non-pacemaker cells, but instead generated by the
influx of the calcium through slow-inward L-type calcium channels.
The shape of the ``slow'' AP observed in Fig.~\ref{fig:cardiac_AP}(B)
is the result of the fact that all the ``fast'' \ce{Na+} channels are
blocked as soon as the membrane potential is depolarized to
a threshold between -40mV to -30mV. When this occurs, the AP can still be
elicited, however, the inward current is carried by only the inward \ce{Ca^+}
influx\footnote{\url{http://www.cvphysiology.com/Arrhythmias/A006.htm}}. As
a result, we get a slower AP.
As the membrane change from resting potential (about -60mV in SA node), to about
-50mV, T-type $\Ca$ channels open which brings $\Ca$ into the cells. The inward
$\Ca$ current further depolarize the cell. When the membrane potential reaches
about -40mV, another type of $\Ca$ channel (L-type) opens, which can bring the
membrane potential to the desired threshold (end of phase 4) to trigger the rise
of potential to the peak (phase 0). During phase 0, $\Na$ current and $\Ca$
current via T-type decreases. When membrane potential reaches the peak, the
outward $\K$ current becomes active, and L-type $\Ca$ channels become
inactivated and close, which decreases inward current and increase outward
current. This helps bringing membrane potential downward to the resting
value (near the reversal potential for $\K$ current, i.e. about -96mV) in SA
node.
So, the changes in membrane potential is brought about by changes in $\Ca$ and
$\K$ mainly, and to a lesser extent by $\Na$ movement across the membrane
through ion channels. The rate can be modified by external factors: autonomic
nerves, hormones, drugs, ions and ischemia/hypoxia. The non-pacemaker cells can
change into a pacemaker cells under certain conditions. E.g.: if a cell becomes
hypoxic (low oxygen supply).
\begin{framed}
The primary pacemaker cells are those in sinoatrial node (SA node)
with rate of about 70-100 beats per minute. When cells in SA node are
dysfunction, then cells from atrioventricular node (AV node) - area
between the left atrial and right ventricle, within the atria septum -
will take the role, with rate of about 40-60 beats per minutes. If
cells from both SA node and AV node don't work, then cells (left and
right branches) of the Bundle of His, as well as from Purkinje fibres
are used, yet the rate is slower with 30-40 beats per minutes.
\end{framed}
%
% The shape of the ``slow'' AP observed in Fig.~\ref{fig:cardiac_AP}(B)
% is the result of the fact that all the ``fast'' \ce{Na+} channels are
% blocked as soon as the membrane potential is depolarized to
% -50mV. When this occurs, the AP can still be elicited, however, the
% inward current is carried by only the inward \ce{Ca^+}
% influx\footnote{\url{http://www.cvphysiology.com/Arrhythmias/A006.htm}}. As
% a result, we get a slower AP.
\subsection{AP in ventricular cells}
\label{sec:AP-ventricular-myocyte}
To do that, we need to know which ionic contribute to the AP, the
dynamics of individual types of channels. In the first half of this
book, we have discussed them already. In the second part of the book,
we will study in detail how they are combined to form a complex
system. This section will briefly review different ionic currents and
which one contributes to which phase.
The membrane potential of ventricular myocyte at rest is about -85mV.
In human, at 1Hz pacing, AP amplitude was 100mV \citep{li1999}, 132mV
\citep{li1998}, and 135mV \citep{nabauer1996}. Phase 0 upstroke velocity
$(dV_m/dt)_\max$ (Volt/sec) measured experimentally in human ventricular
subepicardium and midmyocardium were 228$\pm$20 (V/s) \citep{drouin1995},
446$\pm 46$ (V/s) \citep{pereon2000}.
After a delay of 3-9ms from the beginning of the stimuls, free
$[\Ca]_i$ begins to elevate quickly as a result of SR $\Ca$ release,
and reaches its peak after an addition of
8-19ms~\citep{cleeman1991}. The elevation of $[\Ca]_i$ is quite
homogeneous~\citep{takamatsu1989}.
\begin{framed}
What to adjust?
\begin{enumerate}
\item increase or decrease the time constants of SR calcium release
channel (e.g. from 2ms to 3.5ms) to adjust the delay to peak
\item adjust the maximal rate constant of calcium release from JSR
from, say 60 to 38 ms$^{-1}$, to obtain a peak calcium transient
value of 1.0$\mu$M during a normal AP.
\end{enumerate}
\end{framed}
% \subsection{Ionic currents}
% \label{sec:ionic-currents-3}
In cardiac ventricular cell membrane, there are mainly 6 ionic
currents\footnote{\url{http://www.cvrti.utah.edu/~quan/ep/paper.html}}:
\begin{enumerate}
\item $I_{Na}$: a fast sodium current
\item $I_{si}$: a slow inward current (old term used in
\citep{trautwein1973sic,beeler1977rap}; actually, is the calcium
current)
\item $I_{K}$: a time-dependent potassium current
\item $I_{K1}$: a time-independent potassium current
\item $I_{Kp}$: a plateau potassium current
\item $I_b$: a time-independent background current
\end{enumerate}
\begin{eqnarray}
\label{eq:396}
I_{ion} = I_{Na} + I_{si} + I_{K} + I_{K1} + I_{Kp} + I_{b}
\end{eqnarray}
We have learnt some of them in the previous chapters, and we will
learn the others in detail in the coming chapters.
For a single ion channel, we can assume the ionic current is related
to the voltage by the Ohm's law
\begin{eqnarray}
\label{eq:397}
I(t) = g(t,V_m)\times (V_m-E_{rev})
\end{eqnarray}
The conductance is determined by the maximum conductance $\overline{g}$ and
the fraction of open channel. Based on Hodgkin-Huxley model, the
fraction of open channel is determined by the hypothetical activation
variable $m$ and inactivation $h$ raised to an integer power.
\begin{eqnarray}
\label{eq:398}
g(t,V) = \overline{g} \times m(t,V)^i \times h(t,V)^j
\end{eqnarray}
with $i,j$ are positive integer; $m,h$ are assumed to obey the
first-order kinetics given as
\begin{eqnarray}
\label{eq:399}
\frac{dx}{dt} = \frac{x-x_\infty}{\tau}
\end{eqnarray}
with $i,j,x_\infty,\tau$ are determined experimentally. % The sections
% in this chapter cover different models in chronological order.
\subsection{AP in different species}
\label{sec:detail-ap-different}
\subsubsection{Canine}
\label{sec:canine}
\subsubsection{Rat}
\label{sec:rat}
\subsubsection{Guinea pigs}
\label{sec:guinea-pigs}
% \subsection{Types of cardiac AP and its Phases}
% \label{sec:phases-cardiac-ap}
% Nevertheless, they can be described in explicit
% phases.
% \begin{figure}[htb]
% \centerline{\includegraphics[height=4cm]{./images/action_potentials_compare.eps}}
% \centerline{\includegraphics[height=5cm]{./images/pacemaker_AP.eps}, \includegraphics[height=5cm]{./images/ventricular_action_potential.eps}}
% \caption{(A) compare between nervel cell AP vs. cardiac
% non-pacemaker AP; Phases of a cardiac AP in (B) pacemaker cells,
% (C) non-pacemaker cells}\label{fig:cardiac_AP}
% \end{figure}
%
% \subsubsection{``slow response'' AP (pacemaker)}
% \label{sec:slow-response-ap}
%
%
% \subsubsection{``fast response'' AP (non-pacemaker)}
% \label{sec:fast-response-ap}
\subsection{Heart rate}
In contrast to skeletal muscle, cardiac muscle can contract on its own in the
absence of neural or hormonal stimulation. This property is called {\bf
automaticity}, or {\bf autorhythmicity}
\footnote{\url{http://www.as.miami.edu/chemistry/2086/NEW-Chap20/NEW-Chapter
20_part2.htm}}. The conducting system includes:
\begin{itemize}
\item SA node (embeded in the posterior wall of the right atrium, near the
entrance of the superior vena cava)
\item AV node (at the junction between the atria and ventricles)
\item conducting cells:
\begin{enumerate}
\item conducting cells in the atria are found in {\it
internodal pathways} (to distribute the electrical signal to atrial muscle
cells as the impulse travels from the SA node to AV node). The connection between the AV node and
AV bundle is called {\bf bundle of His} (the only electrical connection
between the atria and the ventricles).
\item conducting cells in the ventricles includes those in the AV bundle and
the bundle branches, as well as {\it Purkinje fibers}, to distribute the
stimulus to the ventricular myocardium.
\end{enumerate}
\end{itemize}
Most of the cells in the conducting system are smaller than myocardial cells and
contains very few myofibrils. Electrical propgation before reaching the AV node,
it affects only the atria, as the fibrous skeleton isolates the atrium
myocardium from the ventricular myocardium. This takes about 50ms in human for
the stimulus to reach the AV node from SA node, then the atrium contraction
starts. The AV node cause a delay, i.e. the time to for the impulse to pass
through the AV node and enter AV bundle (about 100msec in human).
Then it follows a unidirectional propagation along the interventricular septum
within the AV bundle and enter the {\bf left/right bundle branches} (both
branches extend toward the apex of the heart, turn, and fan out deep to the
endocardial surface) before going into the {\bf Purkinje fibers}, and via the
moderator band, to the papillary muscles of the right ventricle. This phase
occurs very fast, about 25ms in human. The Purkinje fibers can conduct Ap very
fast (as fast as small myelinated axons). Finally, the impulse is distributed
throughout the ventricular myocardium; then the ventricular contraction begin,
starting from the apex of the heart (which also at about the time that the
atrial contraction completes). The entire process, from SA node to the complete
depolarization of the ventricular myocardium, in human takes about 225ms.
The papillary muscles begin contraction before the rest of the ventricular
myocytes as it receives the signal through the moderator band, rather than
through the Purkinje fibers. This contraction applies tension to the chordae
tendineae and brace the AV valves. This tension in the chordae tendineae
prevents the backflow of blood into the atria when the ventricle contracts.
As the contraction in ventricular myocytes start from the apex of the heart,
under electrical-field simulation, we often see a wave that begins at the apex
and spread toward the base.
\begin{framed}
The delay in AV node is important to give enough time for atrial contraction to
occur. Under testing condition, the AV node of a normal heart can conduct at a maximum rate of 230 AP
per minute (even if the SA node generates impulses at a faster rate). NOTE: When the
rate higher than about 180 beat per minute (bpm), the pumping efficiency start
to decrease. So, the rate 230 bpm can only occur when the conducting system is
damaged or stimulated by drugs.
The theoretical maximum limit of human ventricular myocytes pumping rate is
300-400bpm, depending on individual.
\end{framed}
As different region of cardiac cells in the conducting system have their own
rate of spontaneous depolarization, it is the cells with the fastest rate that
determine the whole-heart rate. SA node has the fastest rate: 80-100 per
minutes. However, cells from the AV node has the lower rate: 40-60 AP per
minutes at normal condition. Certain cells in the Purkinje fibers has a slower
rate: 20-40 AP per minute. As SA node reaches the threshold first, it then
define the heart rate. Thus, SA node contains {\bf pacemaker cells} (natural
pacemaker). If this region is damaged, the heart still can beat, yet with a
lower heart-rate.
\subsection{Signal-Force relation}
\label{sec:sign-force-relat}
Essentially, the heart is a pump that squeezes to pump blood to other body
organs. Ultimately, the cellular signal in the heart is aimed to control the
excitation-contraction properly. This process is coordinated by a tightly
controlled mechanism via cell signalling in which $\Ca$ play a major role, a
{\it second messenger}. Starting from the nerve system, the rhythmic electrical
signal is transmitted to the sinoatrial (SA) node. The signal is then
propagated to neighboring cells via gap junctions, into and through the right
and left atria, initiating atria contractions, until they converge at the
atrioventricular (AV) node.
The signal, in the form of depolarization wave, from the AV node of the heart is
conducted through cells, the Hist bundle and the Purkinje fiber conducting
system to initiate the electrical depolarization and contraction of myocardial cells of
both ventricles. The membrane potential changes during a cycle of
contraction-excitation is called action potential (AP), as depicted in
Fig.~\ref{fig:cardiac_AP}.
\begin{framed}
The rapid conduction in His-Purkinje network ensures almost uniform
and synchronous contraction of the ventricle. The slow conduction through AV
node determine and ensure the proper heart rate. The current flow through a
large number of ion channels and pumps underlies and coordinate these electrical
singals. The alteration of these critical balance of this multiple-current
pathway can lead to disruptive, often fatal, rhythm disturbances known as {\bf
cardiac arrhythmias}. Chapter \ref{chap:cardiac-diseases} will discuss in details.
\end{framed}
In an AP, the global transient increase of myoplasmic calcium leads to
a sequence of events in which contraction and excitation is the important one in
cardiac cells.
\begin{itemize}
\item contraction: The elevation of calcium $[\Ca]_i$ from 0.1 to $1\mu$M is the
result of the calcium release, either
from outside via SL or from internal calcium storage SR (Sect.\ref{sec:VICR_CICR}).
Part of the free $\Ca$ quickly bind to intracellular buffers, e.g. Calmodulin.
The other part diffuse to the myoplasm and bind to Troponin (Tpn) to induce the
contraction. The mechanism of $\Ca$ release and the fraction of $\Ca$ released
from different sources is discussed in Sect.\ref{sec:local_control}.
\item excitation: after the contraction, $\Ca$ need to be removed
from the cytosol to allow cell relaxation. The majority of $\Ca$ is
sequestered back into the SR during each heart beat via
{\bf SERCA pump} (sarcoplasmic/endoplasmic reticulum Ca-ATPase). For
fast sequestration, some will be sequestered into mitochondria
via $\Ca$ uniport and then being released out to be
extruded across the sarcolemma via {\bf Na-Ca exchangers}, e.g. in
rabbit and rat \citep{bassani1994rir}.
\end{itemize}
The fundamental unit of contraction in a cardiac muscle fiber is the {\bf
sarcomere} - the contractile apparatus between the {\bf Z-line} (or Z-disk).
However, there are recent evidences of the asymmetric between the left-right
of a sarcommere, suggesting that half-sarcommere is the true fundamental
unit~\citep{telley2006}. Modelling force-calcium signalling is described in
Chap.\ref{chap:force_calcium}.
% \section{Introduction}
% \label{sec:introduction-8}
\section{Ca2+-induced Ca2+-release (CICR)}
\label{sec:cicr}
The potential depolarization triggers the low-voltage-gated \ce{Na+}
channels, i.e. the fast \ce{Na+} current that facilitates the
transient increase in membrane potential to the peak. This happen in a
very short period of time compared to kinetics of other channels, so
in mathematical modeling, it's often modeled as an algebraic function
(rather than an ordinary differential equation).
The depolarized membrane potential, at a certain level, activate the
high-voltage-gated \ce{Ca^2+} channels, which yield a low-level influx of
\ce{Ca^2+} through $\Ca$ channels, mainly DHPR at T-tubules of the sarcolemma.
This is known as $I_{Ca,L}$ (Sect.~\ref{sec:L-type-Ca2+}). The \ce{Ca^2+} influx
trigger the opening of RyRs, which allow a larger calcium to be released from
the internal storage SR. This mechanism - calcium-induced calcium-release (CICR)
- is widely accepted since the discovery by Fabiato (1985) in skinned cardiac
cells, which subsequently observed in other cell types.
\begin{framed}
Calcium entries into the cardiac cell via the main gating L-type
calcium channels (LCC or DHPR). DHPR resides mainly in the
transverse-tubules (T-tubules), though a lesser amount not in the
subspace, which then given the name {\it non-junctional DHPR}.
T-tubule is an inward invagination of the sarcolemma, filled with
mucopolysaccharides, so that the potential depolarization can be considered
homogeneous at every point of the cytoplasm. Mucopolysaccharides aka
glycosaminoglycans are long chains of sugar molecules found throughout the
body, often in fluid and mucus.
\end{framed}
\subsection{Mechanism of $\Ca$ release: VICR vs. CICR}
\label{sec:VICR_CICR}
The electrical stimulus on the surface membrane leads to an AP causing the
depolarization along the surface and along the T-tubules.
In skeletal cells, the electrical depolarization can induce $\Ca$ release from
the underlying cisternae of the SR; thus $\Ca$ entry is not important for
initiating calcium release from SR. In cardiac cells, the depolarization trigger
the opening of $V_m$-gated L-type $\Ca$ channels (LCCs or aka dihydropyridine
receptor (DHPR)). The small influx of calcium via LCC then induces the larger
release of $\Ca$ from the terminal cisternae (and subsarcolemmal cisternae) of
the SR (and of the SL). So, the mechanism in skeletal muscle cells is VICR
(Voltage-induced $\Ca$ release) or depolarization-induced calcium release
(DICR), while that in cardiac cells is CICR ({\bf calcium-induced
calcium-release})~\citep{fabiato1975cic}.
\begin{framed}
{\bf Historical facts}:
In 1972, Bassingthwaighte \& Reuter postulated that the influx of $\Ca$ from
extracellular milieu {\it per se} is not high enough to trigger the
contractile of the heart~\citep{bassingthwaighte1972cme}. Thus, they proposed
that there must be some internal $\Ca$ storage; and the transarcolemmal $\Ca$
influx doesn't trigger the myofilaments directly, yet through the induction of
calcium release from this internal storage. Three years later, Fabiato-Fabiato
has confirmed this hypothesis on skinned cardiac cells (cells with SR removed
by microdissection)~\citep{fabiato1975cic} with the concept {\it
calcium-induced calcium release}, and later on on other species (human, dog,
cat, rabbit, rat, frog) ~\citep{fabiato1979cir}. Essentially, by their
experiments, there is no influx of $\Ca$, of any magnitude could directly
activated the myofilaments in skinned cardiac cells. Researchers also
identified that the only internal storage for calcium is the sarcoplasmic
reticulum (SR), not the mitochondria. Early literature reviews include
~\citep{ikemoto1977crs}, ~\citep{fabiato1983cir}.
\end{framed}
The strong buffering capacity in the cytosol, e.g. 100:1 \citep{bers1991ecc}
maintains a very low level of free calcium at rest, i.e. $10^{-7}$ M (or
$\approx 0.1\mu$M). The low resting $[\Ca]_{i,rest}$ compared to a much higher
concentration in the extracellular environment ($\approx 1.8$mM) facilitates a
large signal-to-background ratio $\Delta [\Ca]_i/[\Ca]_{rest}$ to be generated
just using a small amount of $\Ca$ added to the cytosol. The peak of average
free $[\Ca]_i$ during an AP is about 1$\muM$. Given the above buffering
capacity, it means the total cytosolic calcium to be released is about
100$\muM$ during systole. During relaxation, all this amount of released $\Ca$
need to be removed from the cytosol.
When $\Ca$ is globally released, a majority of free $\Ca$ then bind to the
$\Ca$-binding subunit of the thin filaments protein troponin to activate
contraction process. The cell senses the voltage signal by the change in calcium
concentration, causing it to contract or relax.
Since 1990s, the accepted mechanism of $\Ca$-induced $\Ca$ release (CICR) is a
local control rather than global (cell-wide) process~\citep{Cannell2011}. The
detail is discussed in Sect.\ref{sec:local_control}. Under this local-control
theory, calciums are released at thousands of sites known as calcium release
sites (units - CRU) (Sect.\ref{sec:cru_calcium_release_unit}). Each CRU operates
independently at normal condition. In addition, the slow diffusion of $\Ca$ in
the cytosol is another factor to facilitate the local increase of calcium
easily, creating local response, without having to raise global calcium.
This is achieved thanks to many $\Ca$ buffers in the cytoplasm, e.g. parvalbumin
and vitamin D-dependent $\Ca$ binding proteins. This is energetically important,
as only a small energy is required to remove this local transient.
\subsection{CICR - gradeness release: the importance of local control}
\label{sec:local_control}
% In the previous section, we also learned how $[\ce{Ca^2+}]_{i}$
% changes with ``on'' and ``off'' reactions in myocytes.
\cite{fabiato1983cir} hypothesized that CICR occurs through a process as
described as follows. The calcium influx increase the [free $\Ca$] in the
cytoplasm, which trigger the openings of $\Ca$-mediated gates on the SR
membrane.
Now we know these gates are SR transmembrane ion channel - RyRs - with type2 is
the major isoform in ventricular myocytes. CICR a high gain (or amplification),
and re-generative process which is supposed to give all-or-none behavior.
However, experimental data shown that contractile amplitude is a smoothly {\it
graded} function of $V_m$~\citep{new1972}. It was then hypothesized that the
global elevation of free $[\Ca]_i$ is the summation of all local elevations at
different CRUs. This is the essential content of {\bf local-control theory}.
Cell signalling performs its function through frequency and amplitude of the
signal. For better control, another dimension is added, i.e. the spatial. Local
control guarantees only a small amount of calcium can perform its signalling
purpose. Also, it can target to a particular local region in the cell. However,
as pointed by A.V. Hill, signaling just by means of diffusion is not fast enough
to trigger a homogeneous and instantaneous change in calcium
signalling~\citep{hill1910pea, hill1949}. Given that electrical potential can
progate extremely fast (time constant is 100$\mum$, compared to calcium
diffusion about 1$\mum$. So, the cell must provides some means to use the
electrical potential to facilitate instantaneous change in calcium signalling.
Using electron microscopy (EM), in cardiac cell, it shown that the T-tubule
(transverse-tubule) system, the deep invasion of the surface membrane found at
saromemre Z-lines that is found almost uniformly in the cardiac
cells~\citep{endo1964, porter1957}. So, T-tubules was suggested as a mechanism
to help bringing AP propagation quickly into the internal of the cell where it
can trigger $\Ca$ release from SR at almost simultaneously everywhere in the
cell. The microdomain at which $\Ca$ is released is called the dyadic subspace,
or calcium release site. The structure of a dyadic subspace is discussed in
detail in Sect.\ref{sec:cru_calcium_release_unit}.
\subsection{CICR - high 'gain'}
The concept macroscopic or 'whole-cell' {\bf gain} in EC coupling has received
much investigation \citep{wier2007a}. During a voltage-lamp, high gain refers to
the fact that $\Ca$ to be released from SR is much more than $\Ca$ entering the
cell via L-type $\Ca$ channels (LCC). Interestingly, $I_\CaL$ at negative
potentials is much more efficacious in triggering SR $\Ca$ release than at
positive potentials.
Thus, gain is always higher at negative potentials. So, gain decrease with
voltage increase that approach calcium reversal potential.
Macroscopic $I_\CaL$ is the ensemble of single channel currents $i_\ca$ through
thusands of individual channels. Given $N$ as the total number of LCC in a cell,
and $P_o$ is the opening probability, then
\begin{equation}
I_\CaL = N.P_o.i_\ca
\end{equation}
with $P_o$ is a function of $V_m$. Then, the question is which of the
two factors that control the EC gain: the single channel current $i_\ca$ or
the local $NP_o$ (i.e. at a sigle dyad). The impact of each of the two factors was studied
by \citep{altamirano2007}. Supprisingly, incease either $i_\ca$ or local $NP_o$
alone decrease EC gain, i.e. reducing either of them increase EC gain. Also, a
smaller $i_\ca$ at $V_\stim=+0$ mV may be a highly effective trigger of SR $\Ca$
release, and the redundancy of $\Ca$ channel openings at individual junctions
is critical to local control, i.e. reducing $NP_o$ has a stronger effect on EC
gain than does a decrease in $i_\ca$. So, most dyad need a single channel
opening and that can trigger $\Ca$ sparks, and more than one channel opening is
considerred redundant or wasted opening. The data then suggested that increasing
$i_\ca$ always decreases gain. So, at negative $V_\stim$, where there is less
channel opening during the stimulus, the gain is higher.
Different ways are being used to define
{\bf 'gain'}:
\begin{enumerate}
\item The peak SR $\Ca$ release flux divided by the peak flux of $\Ca$ across
the cell membrane \citep{wier1994lce}. Here, 'gain' is unitless. At +10mV, the
gain is about 12 in rat ventricle. However, as shown in
\citep{shannon2000pfs}, peak $J_\CaL$ and peak $J_\text{SR,rel}$ SR calcium
release are not at the same time.\footnotemark[1]
\item To study the effect of SR change, instead of using peak value, Shannon
et al. used integration \citep{shannon2000pfs}. Total calcium release divided
by the total calcium current, i.e. $\int J_\text{SR,rel}/\int J_\CaL$.
$\int J_\text{SR,rel}$ was obtained from total cytosolic calcium flux after
all other fluxes into and out of cytosol were accounted for. Here, gain is
about 10 at 'normal' SR calcium content of $\approx 0.1\mu$mol/L cytosol; and
as high as 50 at extremely high $[\Ca]_\SRT$.
\end{enumerate}
\footnotetext[1]{Old papers used $F$, rather than $J$ to indicate flux. For
consistency throughout the book, we use $J$ notation.}
\begin{framed}
\textcolor{red}{How to calculate flux}: Here, the flux is defined as the net
rate of movement of $\Ca$ ions into or out of the SR per litre of accessible
cytoplasm. Unit: $J$ (Molar/sec or M/s). From the fluorescence data, free
$[\Ca]_i$ is calibrated using proper formula (Sect.\ref{sec:calibrate_Fluo}).
\end{framed}
Net SR $\Ca$ flux was estimated about \ldots in guinea pig \citep{sipido1991}.
However, the actual SR $\Ca$ release flux ($J_\text{SR,rel}$) was \ldots in frog
skeletal muscle fibres \citep{rios1989}, and \ldots in intact cardiac cells
\citep{wier1994lce}. The formula being used is \citep{wier1994lce}.
\begin{equation}
J_\text{SR,rel} = \frac{d}{dt}[\Ca]_i - J_\text{SR,pump} - J_\text{SR,leak} -
J_{I_\ca} + \sum^N_{n=1} \frac{d}{dt}[\CaL]_n
\end{equation}
Here, the sign of $J$ is positive if the flux tends to increase $[\Ca]_i$, and
the units is chosen as mM/s. So, $J_\text{SR,pump}$ is negative in sign, while
$J_\text{SR,leak}, J_{I_\ca}$ are positive in sign,
Fig.\ref{fig:Wier1994_Fig1}(I).
The total $\Ca$ influx via LCC ($J_{I_\ca}$), which can be measured by
eliminating $\Na$ current, NCX, $\K$ currents \citep{balke1994}.
\citep{shannon2000pfs} blocked NCX by using 0 Na inside and outside the cell;
blocked mitochondrial influx by using mitochondrial uptake inhibitor RU360
inside the pipette. The transient of $\Ca$ is achieved by using voltage-clamp
(with 200ms pulse), with internal perfusion with $\Ca$ indicator. The holding
potential was -40mV, and pulsed potential was ranging from -30mV to +80mVs (e.g.
-30, -20, 0, +20, +40, +60, +80mV). The peak $J_\text{SR,rel}$ achieved at a
clamp pulse potential of 0mV. Apparently, 'gain' of the system, defined as
$J_\text{SR,rel(max)}/J_{I_\ca(\max)}$ vary with membrane potential. As shown in
Fig.\ref{fig:Wier1994_Fig1}(II), the gain decline with more positive voltage
pulse (from -20mV to +20mV). At +10mV, the gain is about 12.
\begin{figure}[hbt]
\centerline{\includegraphics[height=7cm]{./images/Wier1994_Fig1.eps},\includegraphics[height=7cm]{./images/Wier1994_Fig3.eps}}
\caption{ (I) Simulation result in rat (from top to bottom): $[\Ca]_i$, flux
through LCC, flux through SR, and flux through SR pump. (adapted from Fig.1
\citep{wier1994lce}). The continuation line at the lowest figure indicate a
constant leak into the cytoplasm $J_\text{SR,leak}$; (II) The average peak
flux (mean$\pm$S.E.M for 9 cells): (A) $J_\text{SR,rel},J_{I_\ca}$, (B)
normalized of (A); (C) the gain in rat ventricular myocyte is calculated using
$J_\text{SR,rel}/J_{I_\ca}$}
\label{fig:Wier1994_Fig1}
\end{figure}
In addition to voltage-dependency of gain, gain can be changed with pathology or
under different experimental conditions \citep{gomez1997,altamirano2007}. Gain
is expected to increase with a higher SR $\Ca$ load, giving the same $I_\ca$
current \citep{shannon2000pfs}. However, to answer the question whether the
relation is linear or nonline, \citep{shannon2000pfs} found out that during
anormal twitch, the fractional release is about 43\% of SR calcium, and about
4\% during low SR calcium load, and 60\% at maximal SR calcium load. So, they
hypothesized that fractional release was a steeply nonlinear function of
$[\Ca]_\SRT$ and $[\Ca]_\SR$. In other words, the fractional release is not the
simple function of $[\Ca]_\SRT$, but also the release process itself.
SR calcium level was measured by using caffeine application
\citep{ginsburg1998}. The level of $[\Ca]_\SRT$ appears to reach a maximal level
of $\sim 100-120\mu$mol/L cytosol (given that cytosolic volume is about 14 times
larger than SR volume, so maximum SR level is about 14000-16800$\muM$).
As SR calcium content never depleted during the course, \citep{shannon2000pfs}
suggested a measurement to evaluate the maximum 'SR calcium depletion' during a
twitch
\begin{equation}
\text{SR calcium depletion} = \frac{\text{initial SR calcium} -
\text{ minimum SR calcium}}{\text{initial SR calcium}} \times 100
\end{equation}
So, given initial calcium is 96$\mu$mol/L cytosol, and minimum SR calcium
content is 53$\mu$mol/L cytosol; the SR calcium depletion is 45\%. Based on the
fact that no SR release was detected at SR loads below about 50\%, it's
suggested that this is the threshold for a 'shut-off' signals for SR calcium
release. Also, it's suggested that both extra SR calcium and intra SR calcium
(lumenal calcium) both affect to the gating of RyR2s.